Tectonic Evolution, Collision, and Seismicity of Southwest Asia: In Honor of Manuel Berberian's Forty-Five Years of Research Contributions 9780813725253

"This volume, published in honor of Manuel Berberian, brings together a wealth of new data, analyses, and frontier

121 2 66MB

English Pages [680]

Report DMCA / Copyright

DOWNLOAD FILE

Polecaj historie

Tectonic Evolution, Collision, and Seismicity of Southwest Asia: In Honor of Manuel Berberian's Forty-Five Years of Research Contributions
 9780813725253

Citation preview

Tectonic Evolution, Collision, and Seismicity of Southwest Asia: In Honor of Manuel Berberian’s Forty-Five Years of Research Contributions

edited by

Rasoul Sorkhabi University of Utah Energy & Geoscience Institute Salt Lake City, Utah 84108, USA

Special Paper 525 3300 Penrose Place, P.O. Box 9140

Boulder, Colorado 80301-9140, USA

2017

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084697/spe525-00.pdf by University of Wisconsin Madison user

Copyright © 2017, The Geological Society of America (GSA), Inc. All rights reserved. Copyright is not claimed on content prepared wholly by U.S. government employees within the scope of their employment. Individual scientists are hereby granted permission, without fees or further requests to GSA, to use a single figure, a single table, and/or a brief paragraph of text in other subsequent works and to make unlimited photocopies of items in this volume for noncommercial use in classrooms to further education and science. Permission is also granted to authors to post the abstracts only of their articles on their own or their organization’s Web site providing that the posting cites the GSA publication in which the material appears and the citation includes the address line: “Geological Society of America, P.O. Box 9140, Boulder, CO 80301-9140 USA (http://www.geosociety.org),” and also providing that the abstract as posted is identical to that which appears in the GSA publication. In addition, an author has the right to use his or her article or a portion of the article in a thesis or dissertation without requesting permission from GSA, provided that the bibliographic citation and the GSA copyright credit line are given on the appropriate pages. For any other form of capture, reproduction, and/or distribution of any item in this volume by any means, contact Permissions, GSA, 3300 Penrose Place, P.O. Box 9140, Boulder, Colorado 80301-9140, USA; fax +1-303-357-1070; [email protected]. GSA provides this and other forums for the presentation of diverse opinions and positions by scientists worldwide, regardless of their race, citizenship, gender, religion, sexual orientation, or political viewpoint. Opinions presented in this publication do not reflect official positions of the Society. Published by The Geological Society of America, Inc. 3300 Penrose Place, P.O. Box 9140, Boulder, Colorado 80301-9140, USA www.geosociety.org Printed in U.S.A. GSA Books Science Editors: Richard A. Davis Jr. and Christian Koeberl Library of Congress Cataloging-in-Publication Data Names: Sorkhabi, Rasoul B., 1961– editor. | Geological Society of America. Title: Tectonic evolution, collision, and seismicity of southwest Asia : in honor of Manuel Berberian’s forty-five years of research contributions / edited by Rasoul Sorkhabi. Description: Boulder, Colorado : The Geological Society of America, 2017. | Series: Special paper ; 525 | Includes bibliographical references. Identifiers: LCCN 2017001865 | ISBN 9780813725253 (pbk.) Subjects: LCSH: Geology, Structural–Middle East. | Plate tectonics–Middle East. | Seismology–Middle East. | Geology–Middle East. | Berberian, M. Classification: LCC QE634.M53 T43 2017 | DDC 551.80956–dc23 LC record available at https://lccn.loc.gov/2017001865 Cover: Zard Kuh (“Yellow Mountain” in Persian) in the High Zagros of Iran with peaks as high as 4200 m exhibits majestic landscapes in the Zagros Mountains. This photo showing a summit of 3972 m exposes northeast–dipping Jurassic and Cretaceous sediments of the Tethys Ocean— the Bangestan-Khami Group of carbonates. Photo courtesy of Dr. Mohammad Fakhari.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084697/spe525-00.pdf by University of Wisconsin Madison user

Contents

1. Tectonic evolution, collision, and seismicity of southwest Asia: An introduction . . . . . . . . . . . . . . 1 Rasoul Sorkhabi 2. Manuel Berberian: An appreciation and bibliography of his lifelong contribution to geoscience . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17 Rasoul Sorkhabi PART I. ZAGROS OROGEN AND IRANIAN PLATEAU 3. Development of geological perceptions and explorations on the Iranian Plateau: From Zoroastrian cosmogony to plate tectonics (ca. 1200 BCE to 1980 CE) . . . . . . . . . . . . . . . . 25 Manuel Berberian 4. Tehran: An earthquake time bomb. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87 Manuel Berberian and Robert S. Yeats 5. Archaeological and architectural evidence of historical seismic activity along the Zāgros Main Recent fault at Borujerd (western Iranian Plateau); the 1316 C.E. earthquake . . . . . . . . 171 Manuel Berberian, Mohammad Moqaddas, and Ahmad Kabiri 6. Kinematics of the Great Kavir fault inferred from a structural analysis of the Pees Kuh Complex, Jandaq area, central Iran . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 213 Sasan Bagheri, Razieh Madhanifard, and Foruzan Zahabi 7. Mid-ocean-ridge to suprasubduction geochemical transition in the hypabyssal and extrusive sequences of major Upper Cretaceous ophiolites of Iran . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 229 Morteza Khalatbari Jafari, Hassan A. Babaie, and Mohammad Elyas Moslempour 8. The geodynamic significance of the correlation of the Khoy ophiolites in northwest Iran with ophiolites in southeast Turkey . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 291 Morteza Khalatbari Jafari and Hassan A. Babaie 9. Tectono-stratigraphic evidence for the opening and closure of the Neotethys Ocean in the southern Sanandaj-Sirjan zone, Iran . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 319 Mohammad Reza Sheikholeslami 10. The lower–middle Cambrian transition and the Sauk I-II unconformable boundary in Iran, a record of late early Cambrian global Hawke Bay regression . . . . . . . . . . . . . . . . . . . . . . . . . . 343 Yaghoob Lasemi and Hadi Amin-Rasouli

iii

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084697/spe525-00.pdf by University of Wisconsin Madison user

iv

Contents 11. Biostratigraphy of acritarchs and chitinozoans in Ordovician strata from the Fazel Abad area, southeastern Caspian Sea, Alborz Mountains, northern Iran: Stratigraphic and paleogeographic implications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 367 Mohammad Ghavidel-Syooki 12. Controls on the sequence stratigraphic architecture of the Neogene Zagros foreland basin . . . 399 Mortaza Pirouz, Guy Simpson, Sebastien Castelltort, Georges Gorin, and Abbas Bahroudi 13. Size distribution and controls of landslides in the Zagros mountain belt (Iran) . . . . . . . . . . . . . 423 Neda Ghazipour and Guy Simpson PART II. THE CAUCASUS REGION AND ANATOLIA 14. Aspects of the seismotectonics of Armenia: New data and reanalysis . . . . . . . . . . . . . . . . . . . . . 445 Arkadi Karakhanyan, A. Arakelyan, A. Avagyan, and T. Sadoyan 15. Archaeoseismological studies at the Pambak-Sevan-Syunik fault system, Armenia . . . . . . . . . . 479 Arkadi Karakhanyan, R. Badalyan, A. Harutyunian, A. Avagyan, H. Philip, V. Davtyan, G. Alaverdyan, K. Makaryan, and M. Martirosyan 16. Tethyan evolution and continental collision in Georgia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 501 Shota Adamia, A. Chabukiani, T. Chkhotua, O. Enukidze, N. Sadradze, and G. Zakariadze 17. Postcollisional tectonics and seismicity of Georgia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 535 Shota Adamia, V. Alania, N. Tsereteli, O. Varazanashvili, N. Sadradze, N. Lursmanashvili, and A. Gventsadze 18. The evolution of the Intra-Pontide suture: Implications of the discovery of late Cretaceous– early Tertiary mélanges. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 573 Kenan Akbayram, A.M. Celâl Şengör, and Ercan Özcan 19. New synthesis of the Izmir-Ankara-Erzincan suture zone and the Ankara mélange in northern Anatolia based on new geochemical and geochronological constraints. . . . . . . . . . . . 613 Ender Sarıfakıoğlu, Yildirim Dilek, and Mustafa Sevin

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084697/spe525-00.pdf by University of Wisconsin Madison user

The Geological Society of America Special Paper 525

Tectonic evolution, collision, and seismicity of southwest Asia: An introduction Rasoul Sorkhabi* University of Utah, Energy & Geoscience Institute, Salt Lake City, Utah 84108, USA

The whole southern border of Eurasia advances in a series of great folds towards Indo-Africa; these folds lie side by side in closely syntactic arcs, and for long distances they are overthrust to the south against the Indo-African table-land… This circumstance distinctly indicates that the folding of the uppermost part of the Earth’s mass is, under certain conditions, only the expression of a forced adaptation… A great part of this folding is of recent age, or has been continued into very recent times; it is not certain that the movement has ended.

—Eduard Suess (1904), The Face of the Earth, vol. 1, p. 596–597

WHAT’S IN A NAME?

2012), a group of historians and geographers have shown that the term “Middle East,” despite its wide usage, is a purely geopolitical concept devoid of any solid scientific foundation. For one thing, the term is a remainder of the colonial Anglo-Eurocentric perspective that divided the East (Orient) into the Far East, the Near East, and the Middle East. Moreover, there is no consensus among historians and geographers as to how to define the boundaries of the “Middle East.” Over the past century, different authors have included various parts of the vast region from Morocco in North Africa as far east as Afghanistan in their definitions of “the Middle East” (see, for example, see the comparison of “Middle East” maps complied by Bonine, 2012). For these reasons, several eminent geographers (e.g., Cressey, 1960; Dudley, 1962; Brice, 1966; East et al., 1971) have suggested the use of the term “southwest Asia” akin to “southeast Asia.” Likewise, the United Nations does not recognize the terms Near, Far, or Middle East; indeed, the UN Statistics Division (https:// unstats.un.org/) divides Asia and Africa into geographic divisions such as northern Africa, western Asia, southern Asia, and so forth. Indeed, as the terms “Near East” and “Far East” are rarely used today (except for the names of art exhibitions and collections), it is expected that the term “Middle East” will also

The word “Asia,” popularized in Greek and Roman geography, is probably derived from the Akkadian ashu, meaning sunrise or east. Similarly, the Greek term “Anatolia” for Asia Minor means “sunrise or east” (www.etymonline.com). Asia is the world’s largest continent accounting for 30% of the Earth’s land surface (150 million km2); it is the most populous continent, accounting for 60% of the world’s population (7.4 billion in 2016; http://worldpopulationreview.com/continents/asia-population/). Asia is also the Earth’s loftiest topographic feature—thanks to the still active orogenesis that has produced the high plateaus of Tibet, Iran, and Anatolia fringed by spectacular mountain ranges. This high topography, in turn, provides a watershed for large and long rivers as well as vast alluvial plains that have supported agriculture and urban life for millennia (Sengör, 1997). This GSA Special Paper, which grew out of a topical session at the 2013 Geological Society of America Annual Meeting, covers tectonic evolution and seismicity in southwest Asia (Figs. 1, 2, and 3). The region discussed in this volume is often called the “Middle East.” However, we have avoided this misleading term. In a recent book entitled Is There a Middle East? (Bonine et al.,

*[email protected] Sorkhabi, R., 2017, Tectonic evolution, collision, and seismicity of southwest Asia: An introduction, in Sorkhabi, R., ed., Tectonic Evolution, Collision, and Seismicity of Southwest Asia: In Honor of Manuel Berberian’s Forty-Five Years of Research Contributions: Geological Society of America Special Paper 525, p. 1–16, doi:10.1130/2017.2525(01). © 2017 The Geological Society of America. All rights reserved. For permission to copy, contact [email protected].

1

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084703/spe525-01.pdf by University of Wisconsin Madison user

Figure 1. Geographic map of southwest Asia including Iran, Caucasus, Turkey and the Arabian Peninsula, and the surrounding seas (Caspian Sea, Black Sea, Mediterranean, Red Sea, Gulf of Aden, Gulf of Oman, and Persian Gulf). U.A.E—United Arab Emirates.

2 R. Sorkhabi

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084703/spe525-01.pdf by University of Wisconsin Madison user

Figure 2. Seismic activity of southwest Asia (January 2000–November 2017) showing the epicenters of earthquakes with magnitudes greater than 4.5 (earthquake data from the U.S. Geological Survey). The present-day plate boundaries are also shown. U.A.E—United Arab Emirates.

Introduction

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084703/spe525-01.pdf by University of Wisconsin Madison user

3

Figure 3. Simplified geologic map of southwest Asia, showing the distribution of Proterozoic, Paleozoic, Mesozoic, and Cenozoic rocks. Major faults are also shown. U.A.E—United Arab Emirates.

4 R. Sorkhabi

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084703/spe525-01.pdf by University of Wisconsin Madison user

Introduction fade away as geopolitical notions alter and geographic perspectives are better focused. Southwest Asia includes the Arabian Peninsula to the south and the mountainous Anatolian-Iranian plateaus that merge with the Caucasus “Knot” in the north. The region has a long history of human settlement and culture closely associated with natural resources and environmental changes. The Neolithic culture as well as the age of metals (copper, bronze, and iron) began in southwest Asia earlier than other parts of the world. Similarly, the Agricultural Revolution, earliest cities, writing, and mathematics first appeared in the Fertile Crescent of this region. Geologically, southwest Asia is a convergent zone between the Arabian and Eurasian tectonic plates and is characterized by active deformation and seismicity. It is this aspect of the region that is the focus of this GSA Special Paper. The physical geography and resources of southwest Asia were explored and documented by a number of Greek, Roman, Persian, and Arab travelers and scientists through the centuries dating back to 500 BCE. Modern geological observations and mapping in the region began with European explorers in the nineteenth century and continued throughout the twentieth century for scientific, geopolitical, and commercial objectives. The works of Jacques de Morgan (French mining geologist and archaeologist, 1857–1924), Alexander Friedrich von Stahl (German geographer, 1850–1952), and G.E. Pilgrim (British geologist working at the Geological Survey of India, 1875–1943) in Iran spanning the 1890s–1920s epitomize these multinational, multi-purpose studies and exploration. The Caucasus region (Azerbaijan, Armenia, and Georgia), which was forcefully annexed to the former Russian/USSR empire, was explored and mapped mainly under Russian administration, and even today much of the literature about the geology of the Caucasus is in Russian. After World War II, native government efforts in Turkey, Iran, and the Arabian Peninsula in collaboration with Western scientists and scientific institutions founded survey departments, professional associations, research institutes, and universities, which were engaged in geological mapping and investigations, often closely related to oil and gas exploration, groundwater, and earthquake hazard studies. Much of the geoscience literature produced in the region is in the native languages, notably Arabic, Armenian, Hebrew, Persian, Russian, and Turkish. Nevertheless, there are several scientific journals in these countries published in English (Table 1), which mostly present works conducted by native researchers. BIG PICTURE The various views of the “big picture” geology of southwest Asia that have been used in the past century may be categorized into four perspectives briefly described below. (1) Tethyan perspective. The idea that a marine Tethyan realm bounded by Gondwanaland to the south and Laurasia to the north was later deformed to give rise to the long orogenic belt from the Alps through Anatolia, Caucasus, and Iran to the

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084703/spe525-01.pdf by University of Wisconsin Madison user

5

Himalayas and Tibet was developed by the Austrian geologist Eduard Suess (1831–1914) in his masterpiece, Das Antliz der Erde (Suess, 1885–1909; The Face of the Earth, Suess, 1904– 1924). This view, which was considerably based on findings in India and the Himalayas (Sorkhabi, 1995, 1997), was advanced by Alfred Wegener (1880–1930) in Germany in support of his idea of continental drift (Wegener, 1915, 1966), Francis B. Taylor (1860–1938) in America for the origin of “Tertiary mountain belt” (Taylor, 1910), and Emile Argand (1879–1940) in Switzerland for interpreting the origin of Eurasian mountains (Argand, 1924, 1977). Alexander L. Du Toit (1878–1948) in South Africa (Du Toit, 1937) and Arthur Holmes (1890–1965) in England (Holmes, 1944) were also champions of integrating Suess’s paleogeography with Wegener’s idea of continental drift. The Tethyan views of Suess and Argand in conjunction with the Alpine thrust and nappe (lateral compression) theory were widely applied to Asian mountains even before the advent of plate tectonic theory (e.g., Gregory, 1929; Wadia, 1931; Heim and Gansser, 1939; Huang, 1945; Norin, 1946). Indeed, this Tethyan perspective was such a popular scheme for interpretation of Asian orogenies that even the geographer L. Dudley Stamp used it in the various editions of his celebrated textbook on Asia (Stamp, 1962). (2) Geosynclinal perspective. The idea of geosynclines as the origin of mountain belts was developed for the Appalachians by James Hall (1811–1898) and James Dwight Dana (1813–1895, who actually coined the term “geosyncline”). The idea was elaborated by European geologists Gustave Emile Haug (1861–1927), Leopold Kober (1883–1970), and Hans Stille (1876–1966), as well as American geologists Charles Schuchert (1858–1942) and Marshall Kay (1904–1975), in the first half of the twentieth century. Despite its popularity in North America and parts of Europe, the geosynclinal idea was seldom used for any serious interpretation of mountain belts in Asia. Of course, some geologists casually referred to the Asian Tethys as a geosyncline, but the complex and at times misleading scheme and terminology of geosynclinal evolution was not a popular interpretation in Asia. Indeed, some geologists in Asia (e.g., Kashfi, 1976) ascribed to the geosynclinal view as a critique of the plate tectonic interpretation, which was not a successful attempt. (3) Morphotectonic perspective. Some physical geographers (Fisher, 1950, 1961, 1978; Held, 1989), in their attempt to integrate tectonics with landforms in southwest Asia, suggested a three- to fivefold classification of “folded zone,” “thrust zone,” “median zone,” “foreland zone,” and “undeformed basement.” While some elements of this perspective are useful when combined with the Tethyan perspective, the changing position of the so-called “median zone” in various publications has been a confusing, misleading (and currently outdated) view. For instance, Fisher (1950) and Held (1989) correlated the median zone (or trough) with the Zagros foreland basin; Fisher (1961) defined the median zone as a vast area between the highly deformed Zagros and the undeformed Arabian Peninsula (thus including both the foreland basins and the “slightly folded region”). Fisher (1978)

6

Country

R. Sorkhabi TABLE 1. ENGLISH JOURNALS RELATED TO GEOSCIENCE PUBLISHED IN OR ABOUT SOUTHWEST ASIA Journal Publisher Year Electronic Journal of Natural Sciences

National Academy of Sciences of the Republic of Armenia

2003

Stratigraphy and Sedimentology of Oil-Gas Basins: International Scientific Journal (formerly Proceedings Earth Sciences)

Azerbaijan National Academy of Science, Institute of Geology and Geophysics

2006 (1959 USSR)

Georgia

Georgia Journal of Science

Georgia Academy of Science

1942 (USSR)

Jordan

Jordan Journal of Applied Sciences–Natural Sciences

Applied Science Publishers

2004

Iran

Journal of Sciences, Islamic Republic of Iran

Tehran University

1989

Iran

Iranian Journal of Science and Technology, Transactions A: Science

Shiraz University & Springer

1976

Iran

Iranian Journal of Earth Sciences

Islamic Azad University, Mashhad

2009

Iran/UAE

Journal of Middle East Applied Science and Technology

Amadgaran-e Andishe Ofogh Research Institute, Iran, based in Dubai

2012

Iraq

Iraq National Journal of Earth Sciences

Mosul University, Ministry of Higher Education and Scientific Research of Iraq

2001

Iraq

Iraq Journal of Science

Baghdad University

1959

Israel

Israel Journal of Earth Sciences

National Council for Research and Development, Jerusalem; Israel Geological Society

1963

Kuwait

Kuwait Journal of Science and Engineering

Kuwait University

1973

Oman

Sultan Qaboos University Journal for Science

Sultan Qaboos University

1996

Oman

Oman Journal of Applied Sciences

Oman Ministry of Higher Education

2009

Russia

Russian Journal of Earth Sciences

1933

Russia

Russian Journal of Geology and Geophysics

Geophysical Center, Russian Academy of Science V.S. Sobolev Institute of Geology and Mineralogy of the Siberian Branch of the Russian Academy of Sciences & Elsevier

Armenia Azerbaijan

1960

Saudi Arabia

Arabian Journal of Geosciences

Saudi Society for Geosciences and Springer

2008

Saudi Arabia

Saudi Arabian Directorate General of Mineral Resources Bulletin

Ministry of Petroleum and Mineral Resources

1965

Saudi Arabia

Arabian Journal of Science and Engineering

King Fahd University of Petroleum & Minerals and Springer

1975

Turkey

Turkish Journal of Earth Sciences

Scientific and Technical Research Council of Turkey

1976

Turkey

Bulletin of the Technical University of Istanbul

Istanbul Technical University

1948

Turkey

Turkish Association of Petroleum Geologists Bulletin

Turkey

Bulletin of the Mineral Research and Exploration (Turkish and English)

Turkish Association of Petroleum Geologists, Ankara General Directorate of Mineral Research and Exploration (Maden Tetkik ve Arama Dergisi, MTA)

Yemen

Yemen Journal of Sciences

Yemeni Scientific Research Foundations

1996

Yemen

Bulletin of the Faculty of Science, Sana’a University

Faculty of Science, Sana’a University

1987

International

GeoArabia

Gulf PetroLink, Bahrain

1993–2015

International

Journal of Asian Earth Sciences

Elsevier

1997 (1986 Journal of Southeast Asian Geology)

International

Arabian Journal of Earth Sciences

Arabian Scientific Research Organization, Norway and Austria

2004

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084703/spe525-01.pdf by University of Wisconsin Madison user

1989 1935

Introduction defined the median mass as the areas within Iran and Turkey, which were deformed by “Hercynian or earlier” tectonic events. The idea of “median mass” (Zwischenmassiv) in Iran was first proposed by Alexander Friedrich von Stahl (1911), who thought that Central Iran was a stable Precambrian block (“median mass”) fringed by the fold belt of Alborz to the north and Zagros to the south. This concept too is outdated. (4) Plate tectonic perspective. The plate tectonic theory emerged in the 1960s following the discoveries of ocean-floor spreading at mid-oceanic ridges and subduction along trenches. In the early 1970s, the theory was extensively applied to continental tectonics (e.g., Dewey and Bird, 1970; Dewey and Horsfield, 1970; McKenzie, 1970; McKenzie et al., 1970; McKenzie, 1972; Dickinson, 1971; Dewey et al., 1973; Bird et al., 1975; Molnar and Tapponnier, 1975). In the context of southwest Asia, the earliest plate tectonic interpretations based on seafloor spreading (Le Pichon, 1968) or seismic activity (Isacks et al., 1968) depicted a two-plate (Arabia-Eurasia) collision; this view still dominates geology textbooks today. Some geologists have considered the Iranian Plateau as a microplate (Morgan, 1968, 1971), while others have included Iran as part of Eurasia and have instead considered the Anatolian Plateau as a separate microplate (e.g., Bird, 2003; Stern and Johnson, 2010). Dewey et al. (1973) considered both Iran and Anatolia as separate microplates. Still other geologists, following Dewey and Bird (1970), have identified several mobile blocks in the region (such as Afghan block, Lut block, Central Iran, South Caspian block, and Caucasian block) in view of the widespread active deformation and the poly-phased tectonic assembly of southwest Asia (Nowroozi, 1971; McKenzie, 1972; Dewey et al., 1973; Edgell, 1992; Reilinger et al., 2006). Oceanfloor spreading and subduction were keystones in the discovery of plate tectonics. However, oceanfloor rocks do not predate the Jurassic. One particular aspect of Asian geology that has shed much light on the Tethyan tectonic history is the presence of ophiolite belts of various ages on this continent. Studies of these ophiolite belts have led to the idea that two or three Tethys oceanic basins opened and closed during the Phanerozoic Era (Dewey and Birke, 1973; Jenkyns, 1980; Sengör, 1985). Field mapping from Afghanistan (De Lapparent et al., 1970), Tibet (Cheng-Fa and Hsi-Lan, 1973), northern Iran (Stöcklin, 1974, 1977), and northern Turkey and Black Sea margin (Sengör, 1979; Sengör et al., 1980) demonstrate the existence of a Paleo-Tethyan suture zone of Late Triassic–Early Jurassic age. In this view, the Anatolian-Iranian-Tibetan plateaus sandwiched between the Paleo-Tethyan suture to the north and the Neo-Tethyan suture to the south were parts of the Cimmerian continent (or blocks) that drifted away from the northern margin of Gondwana. The Paleo-Tethys probably opened in the Ordovician as evidenced from rift basalts of this age atop the Cambrian platform carbonates in the Alborz mountains of northern Iran (Alavi, 1996; Lasemi 2001); the Neo-Tethys opened in the Permian (Şengör and Kidd, 1979). Cimmerian blocks collided with the paleo-Asian continent during the Late Triassic–Early Jurassic (Sengör, 1984, and references therein). The closure of

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084703/spe525-01.pdf by University of Wisconsin Madison user

7

the Paleo-Tethys was followed by the fragmentation of Gondwana in the Middle Jurassic as well as intense subduction of the Neo-Tethys along the Cimmerian blocks. The Neo-Tethys eventually closed in the Paleogene, in an oblique and diachronous manner: in the Tibetan and Anatolian margins in the middle Eocene, and along the Iranian margin (from northwest to southeast) by the end of the Oligocene (Berberian and King, 1981; Berberian et al., 1982; Searle et al., 1987; Le Pichon et al., 1988; Yin, 2010; Agard et al., 2011; McQuarrie and van Hinsbergen, 2013; Berra and Angiolini, 2014). The Neo-Tethyan closure was followed by the continental collision of the Africa-Arabian and Indian plates with Eurasia and intense structural deformation and orogenic uplift in the entire region. In summary, the plate tectonic theory that was applied to continental collision and deformation in Asia incorporated many of the Tethyan views of Suess and Argand; this integration actually enriched the plate tectonic theory, which had been formulated on the basis of oceanfloor tectonics. The combined Tethyan-plate tectonics view explains the Asian continent in terms of a tectonic assembly that spans the past 550 m.y. of Earth’s history. ACTIVE TECTONICS Geodetic measurements indicate that the Arabian plate currently moves toward Eurasia at a rate of 2–3 mm/yr (Reilinger et al., 2006), although slip rates for the Arabian plate boundaries vary. Cenozoic tectonics has not only produced new deformational structures and orogenic uplift, but has also reactivated the older structures in the uplands. Indeed, diverse structural and magmatic processes in southwest Asia reveal the complex nature of convergent continental tectonics. The Makran subduction zone in southeast Iran represents a still active volcanic arc-trench system (Farhoudi and King, 1977; McCall and Kidd, 1982), reminiscent of Neo-Tethyan subduction. The Pliocene–Quaternary volcanism in northwest Iran, eastern Turkey, Azerbaijan, Armenia, and Georgia, which took place long after the Neo-Tethyan subduction and continental collision, adds another layer of complexity to continental collision tectonics; this volcanism (still active in places like Sabalān) is probably a magmatic imprint of slab breakoff (delamination), asthenospheric flow, and crustal melting subsequent to crustal thickening (Dilek et al., 2010; Chiu et al., 2013). The active volcanism in the region offers opportunities for geothermal resources and geotourism. Major thrust faults and nappes have deformed and uplifted the passive northern margin of the Arabian platform along Neo-Tethyan sutures resulting in the Zagros-Bitlis ranges and associated foreland basins, which are critical for petroleum resources as well as human settlements and agriculture. Escape (extrusion) tectonics of continental blocks is also taking place along major strike-slip faults in central Iran, Turkey, and the Caucasus. The ongoing convergence of Africa-Arabian plate with respect to Eurasia has resulted in a broad zone of structural deformation, mountain uplift, and depositional environments in southwest Asia. The entire region hosts a large number of active

8

R. Sorkhabi

faults, which cause, time and again, tragic earthquakes. The tectonic physiography has, in turn, shaped the sedimentary, ecological, and climatic patterns in the region. For instance, it has been recognized that following the Paleogene Neo-Tethyan closure and the collision of the Africa-Arabian plate with Eurasia, a vast marine realm, of which the Caspian and Black Seas are the remnants, was separated by orogenic uplifts of northern Iran and Turkey. This marine realm of Oligocene–Miocene age, first called Para-Tethys by Vladimir Laskarev in 1924, largely experienced anoxic, cold water conditions in contrast to the southern shallow warm marine waters (e.g., Rogl, 1999). The opening of the Red Sea and Gulf of Aden since the Miocene has added another dimension to southwest Asian tectonics as it has separated the Arabian plate from Africa, and has caused basement uplifts along the western and southern margin of the Arabian Peninsula. These igneous and metamorphic (PanAfrican) basement rocks reveal significant information about the Precambrian evolution of the Arabian Peninsula in the context of Gondwana assembly. Precambrian basement rocks in the orogenic belts of Turkey and Iran are sporadically outcropped and cluster at 600–500 Ma, indicating a significant event of juvenile crustal formation in Iran and Turkey. The tectonic setting of this magmatism, whether passive–margin extensional or active–margin subductional, has been debated; some have argued for an oceancontinent convergence along the Cimmerian margin of Gondwana (Hassanzadeh et al., 2008; Shakerardekani et al., 2015, and references therein). In a larger framework, the Ediacaran–Early Cambrian magmatism was a widespread phenomenon affecting Gondwana—the so-called Pan-African event. Our knowledge of older basement rocks in Iran, Caucasus, and Anatolia is a blank sheet as no rocks older than Edicaran have been found; nevertheless, older U-Pb ages of 2–1 Ga in inherited zircon crystals indicate records of an older continental curst (Hassanzadeh et al., 2008; Nutman et al., 2014; Shakerardekani et al., 2015). ABOUT THIS VOLUME Nearly two-thirds of this volume pertains to the geology of Iran not only because it is the country where Dr. Manuel Berberian, dedicatee of this volume, has worked, but also due to the focus of the papers received for this volume. In Chapter 2, a brief biography of Dr. Berberian, as well as a bibliography of his publications are given by the editor (Sorkhabi). The volume progresses to a significant paper by Dr. Berberian himself on the history of geological ideas and works in Iran dating from Zoroastrian mythology (ca. 1200 BCE) to the era of the plate tectonic theory in the late twentieth century. This informative and insightful paper as well as Berberian (1997, 2014) should form benchmarks for further research on the history of geology in Asia, which is relatively underrated in our academic work and the public perception of how science has evolved. Seismic studies including archaeological and historical investigations are urgently needed in the earthquake-prone and populous countries of southwest Asia. There is a long liter-

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084703/spe525-01.pdf by University of Wisconsin Madison user

ary record of earthquakes in the region, but mostly available in native languages (Arabic, Armenian, Persian, Turkish, etc.) and sometimes in manuscripts written centuries ago; these need to be archived, translated, and examined by scholars. Moreover, there is much room for seismic studies and age determinations by the modern techniques, as have been done in Italy and other parts of Europe. Margottini and Serva (1987) provide examples of such studies from the central and eastern Mediterranean region. Two papers in this volume deal with similar issues. Berberian and Yeats (Chapter 4) detail a case study of Iran’s capital city, Tehran, which sits on active faults at the southern foot of the Alborz Mountains and has a dense population of 8.6 million (http:// worldpopulationreview.com/world-cities/tehran-population/). Berberian, Moqaddas, and Kabiri (Chapter 5) offer a comprehensive (and somewhat speculative) analysis of the 1316 earthquake at Borujerd; the authors use this event as a point of departure to compile all the information at their disposal to examine the seismicity of the Zāgros Main Recent fault in western Iran. Structural mapping and kinematic analysis of faults is not only extremely valuable for seismic studies, but also for deciphering the tectonic evolution of the region. In this context, Bagheri, Madhanifard, and Zahabi (Chapter 6) utilize field observations coupled with satellite imagery to describe the Pees Kuh Complex, a well-preserved Cenozoic structure developed upon the Great Kavir fault near Jandaq in central Iran. This case study demonstrates the poly-phased deformation history of most faults in the region. Two papers deal with the ophiolites of Iran. Khalatbari Jafari, Babaie, and Moslempour (Chapter 7) discuss geochemical, chronological, and field data from the Upper Cretaceous Khoy, Kermanshah, Fannuj, Nosratabad, south Fariman, northwest Fariman, Dehshir, and Sabzevar ophiolite massifs of Iran. This paper also provides a bibliography of previous studies on the ophiolites of Iran. A second paper by Khalatbari Jafari and Babaie (Chapter 8) correlates the Khoy ophiolites in northwest Iran with the Elazığ volcano-sedimentary rocks in southeast Turkey. The tectonostratigraphy of the Sanandaj-Sirjan zone between central Iran and the Zagros fold-and-thrust belt has been relatively poorly understood, owing to its complex structural and metamorphic history. Sheikholeslami (Chapter 9) describes a field study and mapping of the southern part of this zone (the Neyriz-Sirjan area). He details a stratigraphic record comprised by 11 rock units spanning the Paleozoic–Cenozoic times that can be compared and contrasted with the tectonostratigraphy of both central Iran and the Zagros orogen. The author then interprets the evolution of the Sanandaj-Sirjan as a part of the Cimmerian terrane that was accreted to central Asia (Eurasia) in the Late Triassic. The zone was later incorporated into the Zagros orogen, sensu lato, after the continental collision and suturing between the Afro-Arabian plate and the Eurasian (Iranian subplate). Moving to north Iran and back in time, Lasemi and AminRasouli (Chapter 10) present paleontological and field data from the Lalun and Mila Formations of the Alborz Mountains and offer a new interpretation for the facies and sequence stratigraphy

Introduction of the Lower–Middle Cambrian transition in northern Iran. They argue that an unconformity boundary on the quartzite marker unit was related to the late early Cambrian global Hawke Bay (Toyonian) regression (Sauk I-II unconformity). Also from the Alborz Mountains in northern Iran, Ghavidel-Syooki (Chapter 11) reports new biostratigraphic data on acritarchs and chitinozoans from the Lower Paleozoic rock units (the Abastu and Abarsaj Formations) in the Fazel Abad area. He then compares these Ordovician data with coeval acritarch and chitinozoan assemblages reported from Gondwana, Laurentia, and Baltica. The line of research presented in these two chapters, that is, placing local stratigraphic and paleontological studies in a global framework of geologic record, is an endeavor that requires more attention from the native geologists in southwest Asia. Such efforts will not only help the geologists better examine (or re-examine) their study areas but also contribute to global tectonostratigraphy. Moving to the Cenozoic collision zone in southwest Iran, Pirouz et al. (Chapter 12) present a sequence stratigraphic architecture for the Neogene foreland basin in Zagros based on field and paleontological data as well as strontium isotope signatures. These authors examine both tectonic and seal-level controls and find distinct variations in the sequence stratigraphy between the western and eastern sectors of the orogen-parallel foreland basin. Ghazipour and Simpson (Chapter 13) analyze the size distribution of 335 landslides in the Zagros mountain belt using digital maps and field data; they use principal component analysis to understand the effects of geological and geomorphological factors on their data, and note that lithology, elevation, and slope mainly control the distribution of large landslides in Zagros. The last six chapters of the book pertain to the Caucasus (Armenia and Georgia, two chapters each) and Anatolia (Turkey, the last two chapters). We regret that we did not receive contributions on the geology of Azerbaijan for this volume to complete the discussion on the Caucasus. Karakhanyan and colleagues from the Institute of Geological Studies, National Academy of Sciences of Armenia have contributed two papers to this volume. In Chapter 14 Karakhanyan et al. present new data and analyses on the seismotectonics of Armenia. The authors review the geometry, kinematics, and slip rates of selected active faults in Armenia (the Pambak-Sevan-Syunik fault system, and Garni, Akhouryan, Javakhq, Sizavet, and Akera faults). Two case studies are highlighted in this chapter. First, the authors re-examine data for the Dvin earthquakes of the ninth century CE related to the Garni fault located 17 km north of the ancient city; second, they re-examine the active faulting in the Lake Sevan basin and identify pull-apart features on segments of the Pambak-Sevan-Syunik fault system. In Chapter 15, Karakhanyan et al. detail the seismic history of the southeastern segment of the Pambak-Sevan-Syunik fault system (PSSF-3) based on archaeological and paleoseismological evidence studied at six localities by the authors during an eight-year project. They find earthquake signatures on the fault dating back to twenty-fourth–sixteenth centuries BCE as well as twelfth-ninth centuries BCE; they also report radiocarbon dating and historical record for a 368 CE earthquake. The authors

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084703/spe525-01.pdf by University of Wisconsin Madison user

9

estimate earthquake-related horizontal displacements as much as 7.5–8.0 m for the strike-slip fault. Georgia consists of the fold-and-thrust belts of the Greater and Lesser Caucasus and the foreland basins bounded by these two mountain belts. Two papers by Shota Adamia and his colleagues at Tbilisi State University deal with Georgia. The first of these two papers (Adamia et al., Chapter 16) reviews the tectonic evolution and continental collision in Georgia constrained by new field observations and geochemical data from several key mafic rock units. The authors then present a tectonic evolution model for Georgia for the past 500 m.y. involving the birth and demise of the Paleo- and Neo-Tethys oceans. The second paper by Adamia et al. (Chapter 17) discusses the postcollisional Neogene–Quaternary tectonics of Georgia with emphasis on the trends of active faults and associated seismicity. The Racha earthquake of 29 April 1991 (Mw 6.9) rejuvenated modern seismic studies in Georgia, and Chapter 17 is actually a result of these recent studies. The Caucasus is actually a physiographic “knot” between the Iranian Plateau on the southeast and the Anatolian Plateau to the west. The last two papers in this volume deal with the Anatolian Plateau; both papers were contributed by researchers from Turkey and deal with suture zones. Akbayram et al. (Chapter 18) discuss the evolution of the Intra-Pontide suture (between the İstanbul Zone and the Sakarya Continent in northwest Turkey) in the light of late Cretaceous–early Tertiary mélanges, which they studied using paleontological and stratigraphic data in addition to reported radiometric ages. The authors argue for various metamorphic-accretionary prisms along the suture zone. The paleontological data indicate an early Ypresian collision between Sakarya and İstanbul blocks. Sarıfakıoğlu et al. (Chapter 19) offer a new synthesis of the Izmir-Ankara-Erzincan (Neo-Tethyan) suture zone and the Ankara mélange (Late Cretaceous) in northern Turkey based on new field observations, geochemical data, and radiometric ages. The authors then offer a tectonic evolution model involving the subduction of Northern Neo-Tethys beneath the Sakarya continental terrane from the Late Permian to the middle Eocene associated with suprasubduction zone and seamount volcanoes. The closure of the Neo-Tethys resulted in the collision of the Central Anatolian crystalline complex with Sakarya. FRONTIER RESEARCH AND REGIONAL COOPERATION The Tethyan record of southwest Asia has important implications for understanding the tectonic evolution of Asia as well as the ongoing processes of structural deformation and the development of petroliferous sedimentary basins in the region. In addition, Cenozoic tectonics of the region has direct relevance for hazardous seismic activity, landscape development, and climatic changes that have shaped the region’s ecosphere. This GSA Special Paper cherishes the heritage of previous books and volumes on southwest Asian geology, geophysics, tectonics, and

10

R. Sorkhabi

seismology. The Appendix gives a select bibliography for the benefit of researchers and students in this field. In the process of editing the papers for this volume (as well as two other volumes related to Tethyan Asia, i.e., Macfarlane et al., 1999; Sorkhabi and Heydari, 2008) and reading these and other papers on the geology, tectonic evolution, and sesimotectonics of southwest Asia, I decided to select the six fields of geoscience that are most essential either for elucidating the tectonic evolution of the region or for addressing environmental challenges. These are outlined below. (1) High-resolution geologic mapping. Cartographic and survey departments in southwest Asia have produced a large number of maps since the 1950s. These efforts need to be strengthened by large-scale mapping (1:100,000 or even larger for critical areas) combined with high-resolution paleontological analyses, facies mapping, and sequence stratigraphy. (2) Basement-basin tectonics. Our knowledge of the basement rocks and structures in Turkey, the Caucasus, and Iran is limited largely due to scarcity of exposures. Geophysical imaging of the subsurface and analyses of exposed igneous and metamorphic bodies can shed much light on this issue. It is crucial that geochronological studies are coupled with geochemical analyses on the same samples; dates alone do not clarify tectonic settings. (3) Ophiolites. There is a large number of ophiolite exposures of various sizes and ages in the region, some of which have been dated by modern radiometric methods. Nevertheless, there is still much work that needs to be done in this area, and it is important to integrate geochronological and geochemical data with detailed structural mapping and paleontological associations. Such data will help us better understand the opening and closure of Tethyan oceanic basins and constrain the regional tectonic models. (4) Paleobiology. The Tethyan and Cenozoic sedimentary and fossil records of southwest Asia offer significant data for understanding the evolution of life forms, and climatic and environmental changes of the geologic past, often in relation to tectonic events. However, research in traditional stratigraphy and paleontology needs to be combined with geochemical analyses and ecological frameworks. (5) Seismotectonics. Large settlements with weak building structures in the vicinity of active faults in the region have been, time and again, the scenes of tragic earthquakes. Integrated seismic studies should include structural mapping, paleoseismological trench excavations, Coulomb stress modeling and elastic stress transfer of active faults, archaeo- and historical seismology, InSAR (interferometric synthetic aperture radar) technology for mapping the active ground deformation, waveform modeling, and designing hazard risk maps and codes for earthquakeresistant buildings. A research institute dedicated to this field, such as the Tectonics and Seismotectonics Department (established in 1971 at the Geological Survey of Iran) or International Institute of Earthquake Engineering and Seismology (founded in 1989 in Tehran), can be very helpful. (6) Hydrogeology. Water is a critical resource and its management in the arid and semi-arid climates of southwest Asia

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084703/spe525-01.pdf by University of Wisconsin Madison user

poses particular challenges. History shows profound relationships between water and civilization in southwest Asia; indeed, the ingenious qanāt system of ancient Iran was designed to address the water needs for drinking and irrigation. In recent decades, with population growth and concentration in big cities coupled with mismanagement and climate change, the availability of fresh ground- and surface water has become an acute problem, and at times a cause of conflicts in southwest Asia. Water resources and management in the region thus require special attention and application of modern science and technology. In order to achieve the above-mentioned objectives, several plans of actions are imperative and are briefly discussed below: (1) Geoscience education. This should include both school and public education (books, magazines, documentary films, museums, geoparks, geotourism, etc.) and universities capable of conducing doctoral and postdoctoral research. Geoscience education should address outstanding scientific and societal problems. Close relationships among scientists, engineers, industrialists, planners, and policymakers can train the necessary expert workforce for these tasks. (2) Regional cooperation among southwest Asian universities, surveys, institutions, and academies on research projects. A regular conference on Southwest Asian Earth and Environmental Sciences (held every two years), a regional journal in the same field, and university scholarships for exchange researchers and students will facilitate such cooperation. (3) Cooperation of southwest Asian educational and research circles with their counterparts in Western and East Asian countries, as well as with international scientific bodies will be hugely beneficial for the transfer of science and technology to, and internationalization of, geoscience in southwest Asia. Some aspects of southwest Asia have particular significance to science and society, such as Tethyan evolution, ophiolites, continental deformation, seismotectonics, seismic hazard mitigation, salt tectonics, petroleum geology, tectonic geomorphology, and archaeogeology; these aspects will aptly be of more interest to international research programs. (4) Allocation of funds and a reasonable budget, as well as creating a constructive work environment for geoscientists by southwest Asian governments. Finally, cooperation in geoscience can contribute to peace and progress in southwest Asia, where many resource and environmental problems can be better addressed regionally and scientifically. ACKNOWLEDGMENTS This introductory chapter was read by Manuel Berberian and Ezat Heydari. The papers for this volume were reviewed by a large number of fellow scientists (listed alphabetically below). In addition, I would like to express my gratitude to Manuel Berberian as well as to GSA Science Editor Christian Koeberl and GSA staff for their collaboration and efforts on behalf of this volume.

Introduction Reviewers: Hassan Babaie, Manuel Berberian, Caroline Burberry, Ricardo Caputo, Nicola Casagli, William Cavazza, Ömer Elitok, Mohammad Fakhāri, Adam Forte, Eldon Gath, Mohammad Ghassemi, Mehmet Çemal Göncüoğlu, Jean Robert Grasso, Levent Gülen, Jamshid Hassanzadeh, Bruno Helly, Ezat Heydari, Stephen Homke, James Jackson, Arvind K. Jain, Sudeep Kanungo, Pierre Kruse, Pascal Lacroix, Juliette Lamarche, Angela Landgraf, Jeong-Hyun Lee, Ian Lindsay, Michele Marroni, Kathleen Nicoll, Daniela Pantosi, Delmar Keith Patton, Brian Pratt, JeanFrançois Ritz, Thomas Servais, Hadi Shafaii Moghaddam, Jack Shroder Jr., Hareshwar Sinha, Manuel Sintubin, Gerard Stampfli, Stathis C. Stiros, Lans Taylor, Vladimir G. Trifonov, Eutizio Vittori, John Wakabayashi, Xiao Wenjiao, Yücel Yimlaz, Mehdi Zare, Andrey Zhuravlev. POSTSCRIPT In November 2017, during the finalization of this volume for production, two sad events occurred. First, on 12 November a strong earthquake (M 7.3) ~30 km south of Halabja near the border of Iraq and Iran resulted in the death of hundreds of people and destruction of thousands of houses and buildings. The earthquake (34.905° N, 45.956° E), like many other earthquakes in the region, had a shallow depth (19 km) and occurred on an oblique-thrust fault. Second, one of the key authors of this volume, Dr. Arkadi Karakhanyan (senior author of Chapters 14 and 15) passed away. His untimely death at the age of sixty-six was a great loss to the seismologic community as Dr. Karakhanyan was an eminent scientist in Armenia, having served at the Institute of Geological Sciences, National Academy of Sciences of Armenia in Yerevan, for decades. REFERENCES CITED Agard, P., Omrani, J., Jolivet, L., Whitechurch, H., Vrielynck, B., Spakman, W., Monié, P., Meyer, B., and Wortel, R., 2011, Zagros orogeny: A subduction-dominated process: Geological Magazine, v. 148, no. 5–6, p. 692–725, doi:10.1017/S001675681100046X. Alavi, M., 1996, Tectonostratigraphic synthesis and structural style of the Alborz mountain system in northern Iran: Journal of Geodynamics, v. 21, p. 1–33, doi:10.1016/0264-3707(95)00009-7. Argand, E., 1924, La tectonique de l’Asie: Comptes Rendus de la 13e Congrès Géologique International, Belgique (1922), Vaillant-Carmanne, Liège, p. 171–372. Argand, E., 1977, The Tectonics of Asia, translated by A.V. Carozzi: New York, Hafner, 218 p. Berberian, M., 1997, An Investigation into the History of Cosmogony and Earth Science in Ancient Iran [Jostāri Dar Dānesh-e Kayhān va Zamin Dar Irānveij]: Tehran, Bonyād Neyshābur Publishers, 551 p. [in Persian]. Berberian, M., 2014, Earthquakes and Coseismic Surface Faulting on the Iranian Plateau; A Historical, Social, and Physical Approach: Amsterdam, Elsevier, Developments in Earth Surface Processes 17, 714 p. Berberian, M., and King, G.C., 1981, Towards a paleogeography and tectonic evolution of Iran: Canadian Journal of Earth Sciences, v. 18, p. 210–265, doi:10.1139/e81-019. Berberian, F., Muir, I.D., Pankhurst, R.J., and Berberian, M., 1982, Late Cretaceous and early Miocene Andean-type plutonic activity in northern Makran and central Iran: Journal of the Geological Society, London, v. 139, no. 5, p. 605–614, doi:10.1144/gsjgs.139.5.0605.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084703/spe525-01.pdf by University of Wisconsin Madison user

11

Berra, F., and Angiolini, L., 2014, The evolution of the Tethys region throughout the Phanerozoic: A brief tectonic reconstruction, in Marlow, L., Kendall, C., and Yose, L., eds., Petroleum Systems of the Tethyan Region: American Association of Petroleum Geologists Memoir 106, p. 1–27, doi:10.1306/13431840M1063606. Bird, P., 2003, An updated digital model of plate boundaries: Geochemistry Geophysics Geosystems, v. 4, no. 3, 1027, doi:10.1029/2001GC000252. Bird, P., Toksoz, M.N., and Sleep, N.H., 1975, Thermal and mechanical models of continent-continent convergence zone: Journal of Geophysical Research–Solid Earth, v. 80, p. 4405–4416, doi:10.1029/ JB080i032p04405. Bonine, M.E., 2012, Of maps and regions: Where is the geographer’s Middle East? in Bonine, M.E., Amanat, A., and Gasper, M.E., eds., Is There a Middle East? The Evolution of a Geopolitical Concept: Stanford, California, Stanford University Press, p. 56–99. Bonine, M.E., Amanat, A., and Gasper, M.E., eds., 2012, Is There a Middle East? The Evolution of a Geopolitical Concept: Stanford, California, Stanford University Press, 341 p. Brice, W.C., 1966, South-West Asia, A Systematic Regional Geography, Volume VIII: London, University of London Press, 448 p. Cheng-Fa, C., and Hsi-Lan, C., 1973, Some tectonic features of the Mt. Jolmo Lungma area, southern Tibet, China: Scientia Sinica, v. 16, no. 2, p. 257–265. Chiu, H.-Y., Chung, S.-L., Zarrinkoub, M.H., Mohammadi, S.S., Khatib, M.M., and Lizuka, Y., 2013, Zircon U-Pb age constraints from Iran on the magmatic evolution related to Neotethyan subduction and Zagros orogeny: Lithos, v. 162–163, p. 70–87, doi:10.1016/j.lithos.2013.01.006. Cressey, G.B., 1960, Crossroads: Land and Life in Southwest Asia: Chicago, J.B. Lippincott Co., xiv + 593 p. De Lapparent, A.F., Tremier, H., and Tremier, G., 1970, Sur la stratigraphie et la paléobiologie de la série permocarbonifére du Dacht-e Nawar (province de Ghazni, Afghanistan): Bulletin de la Société Géologique de France, ser. 7, v. 12, p. 565–572. Dewey, J.F., and Bird, J.M., 1970, Mountain belts and the new global tectonics: Journal of Geophysical Research–Solid Earth, v. 75, p. 2625–2647, doi:10.1029/JB075i014p02625. Dewey, J.F., and Birke, K.C., 1973, Tibetan, Variscan, and Precambrian basement reactivation: Products of continental collision: The Journal of Geology, v. 81, p. 683–692, doi:10.1086/627920. Dewey, J.F., and Horsfield, B., 1970, Plate tectonics, orogeny, and continental growth: Nature, v. 224, p. 1031–1035. Dewey, J.F., Pitman, W.C., III, Ryan, W.B.F., and Bonnin, J., 1973, Plate tectonics and the evolution of the Alpine system: Geological Society of America Bulletin, v. 84, p. 3137–3180, doi:10.1130/0016-7606(1973)842.0.CO;2. Dickinson, W.R., 1971, Plate tectonic models for orogeny at continental margins: Nature, v. 232, p. 41–42, doi:10.1038/232041a0. Dilek, Y., Immamverdiyev, N., and Altunkaynak, Z., 2010, Geochemistry and tectonics of Cenozoic volcanism in the Lesser Caucasus (Azerbaijan) and the peri-Arabian region: Collison-induced mantle dynamics and its magmatic fingerprint: International Geology Review, v. 52, p. 536–578. Du Toit, A.L., 1937, Our Wandering Continents: Edinburgh, Oliver and Boyd, 346 p. Dudley, S.L., 1962, Asia: A Regional and Economic Geography (11th ed.): London, Methuen & Co., xvii + 730 p. East, W.G., Spate, O.H.K., and Fisher, C.A., eds., 1971, The Changing Map of Asia: A Political Geography (5th ed.): London, Methuen & Co., xvi + 678 p. Edgell, H.S., 1992, Basement tectonics of Saudi Arabia as related to oil field structures, in Rickard, M.J., et al., eds., Basement Tectonics 9: Dordrecht, Netherlands, Kluwer Academic Publishers, p. 169–193, doi:10.1007/978 -94-011-2654-0_10. Farhoudi, G., and King, D.E., 1977, Makran of Iran as an active arc system: Geology, v. 5, p. 664–667, doi:10.1130/0091-7613(1977)52.0.CO;2. Fisher, W.B., 1950, The Middle East (1st ed.): London, Methuen, xiii + 514 p. Fisher, W.B., 1961, The Middle East (4th ed.): London, Methuen, xiv + 527 p. Fisher, W.B., 1978, The Middle East: A Physical, Social, and Regional Geography (7th ed.): London, Methuen, 615 p. Gregory, J.W., ed., 1929, The Structure of Asia: London, Methuen, xii + 227 p. Hassanzadeh, J., Stockli, D.F., Horton, B.K., Axen, G.J., Stockli, L.D., Grove, M., Schmitte, A.K., and Walker, J.D., 2008, U-Pb zircon geochronology of late Neoproterozoic–Early Cambrian granitoids in Iran: Implications

12

R. Sorkhabi

for paleogeography, magmatism, and exhumation history of Iranian basement: Tectonophysics, v. 451, p. 71–96, doi:10.1016/j.tecto.2007.11.062. Heim, A., and Gansser, A., 1939, The Central Himalayas—Geological Observations of the Swiss Expedition of 1936: Mémoires de la Société Helvétique des Sciences Naturales, v. 73, 245 p. Held, C.C., 1989, Middle East Patterns: Places, Peoples, and Politics: Boulder, Colorado, Westview Press, xxii + 442 p. Holmes, A., 1944, Principles of Physical Geology (1st ed.): London, Thomas Nelson & Sons, 540 p. Huang, T.K., 1945, On Major Tectonic Forms of China: Pehpei, Chungking, National Geological Survey of China, Geological Memoirs, series A, no. 20, 165 p. Isacks, B., Oliver, J., and Sykes, L.R., 1968, Seismology and the new global tectonics: Journal of Geophysical Research, ser. B, v. 73, p. 5855–5899, doi:10.1029/JB073i018p05855. Jenkyns, H.C., 1980, Tethys: Past and Present: Proceedings of the Geologists’ Association: Geologists’ Association, v. 91, p. 107–118, doi:10.1016/ S0016-7878(80)80018-6. Kashfi, M., 1976, Plate tectonics and structural evolution of the Zagros geosynclines, southwestern Iran: Geological Society of America Bulletin, v. 87, p. 1486–1490, doi:10.1130/0016-7606(1976)872.0.CO;2. Lasemi, Y., 2001, Facies Analysis, Depositional Environments and Sequence Stratigraphy of the Upper Pre-Cambrian and Paleozoic Rocks of Iran: Tehran, Geological Survey of Iran, 180 p. [in Persian]. Laskarev, V., 1924, Sur les equivalents du Sarmatien supérieur en Serbie, in Vujevic, R., ed., Receuil de traveaux offert à M. Jovan Cvijiic par ses amis et collaborateurs: Beograd, Drzhuavna Shtamparija, p. 73–85. Le Pichon, X., 1968, Sea-floor spreading and continental drift: Journal of Geophysical Research, ser. B, v. 73, p. 3661–3697, doi:10.1029/ JB073i012p03661. Le Pichon, X., Bergerat, F., and Roulet, M.J., 1988, Plate kinematics and tectonics leading to the Alpine belt formation; a new analysis, in Clark, S.P., Jr., Burchfiel, B.C., and Suppe, J., eds., Processes in Continental Lithospheric Deformation: Geological Society of America Special Paper 218, p. 111–132, doi:10.1130/SPE218-p111. Macfarlane, A., Sorkhabi, R.B., and Quade, J., eds., 1999, Himalaya and Tibet: Mountain Roots to Mountain Tops: Geological Society of America Special Paper 328, 330 p. Margottini, C., and Serva, L., eds., 1987, Historical Seismicity of CentralEastern Mediterranean Region: Italian Commission for Nuclear and Alternative Energy Sources and International Atomic Energy Agency, ENEA-IAEA Workshop, Rome, 27–29 October 1987, 337 p. McCall, G.J.H., and Kidd, G.W., 1982, The Makran, southeastern Iran: The anatomy of a convergent plate margin active from Cretaceous to Present, in Leggett, J.K., ed., Trench-Forearc Geology: Sedimentation and Tectonics on Modern and Ancient Active Plate Margins: Geological Society, London, Special Publication 10, p. 387–397, doi:10.1144/GSL .SP.1982.010.01.26. McKenzie, D.P., 1970, Plate tectonics of the Mediterranean region: Nature, v. 226, p. 239–243, doi:10.1038/226239a0. McKenzie, D.P., 1972, Active tectonics of the Mediterranean region: Geophysical Journal of the Royal Astronomical Society, v. 30, p. 109–185, doi:10.1111/j.1365-246X.1972.tb02351.x. McKenzie, D.P., Davies, D., and Molnar, P., 1970, Plate tectonics of the Red Sea and East Africa: Nature, v. 226, p. 243–248, doi:10.1038/226243a0. McQuarrie, N., and van Hinsbergen, D.J.J., 2013, Retrodeforming the ArabiaEurasia collision zone: Age of collision versus magnitude of continental subduction: Geology, v. 41, p. 315–318, doi:10.1130/G33591.1. Molnar, P., and Tapponnier, P., 1975, Cenozoic tectonics of Asia: Effects of a continental collision: Science, v. 189, p. 419–426, doi:10.1126/ science.189.4201.419. Morgan, W.J., 1968, Rises, trenches, great faults, and crustal blocks: Journal of Geophysical Research, ser. B, v. 73, p. 1959–1982, doi:10.1029/ JB073i006p01959. Morgan, W.J., 1971, Convection plumes in the lower mantle: Nature, v. 230, p. 42, doi:10.1038/230042a0. Norin, E., 1946, Geological explorations in western Tibet: Stockholm, SinoSwedish Expedition Publication 29, 205 p. Nowroozi, A.A., 1971, Seismo-tectonics of the Persian Plateau, eastern Turkey, Caucasus and Hindu-Kush region: Bulletin of the Seismological Society of America, v. 61, p. 317–341.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084703/spe525-01.pdf by University of Wisconsin Madison user

Nutman, A.P., Mohajjel, M., Bennett, V.C., and Fergusson, C.L., 2014, Gondwanan Eoarchean-Neoproterozoic ancient crustal material in Iran and Turkey: Zircon U-Pb-Hf isotopic evidence: Canadian Journal of Earth Sciences, v. 51, p. 272–285, doi:10.1139/cjes-2013-0138. Reilinger, R., McClusky, S., Vernant, P., Lawrence, S., Ergintav, S., Cakmak, R., Ozener, H., Kadirov, F., Guliev, I., Stepanyan, R., and Nadariya, M., 2006, GPS constraints on continental deformation in the Africa-ArabiaEurasia continental collision zone and implications for the dynamics of plate interactions: Journal of Geophysical Research–Solid Earth, v. 111, no. B5, B05411, doi:10.10292005JB004051. Rogl, F., 1999, Mediterranean and Paratethys: Facts and hypotheses of an Oligocene to Miocene paleogeography (short overview): Ecologica Carpathica, v. 50, p. 339–349. Searle, M.P., Windley, B.F., Coward, M.P., Cooper, D.J.W., Rex, A.J., Rex, D., Tingdong, L., Xuchang, X., Jan, M.Q., Thakur, V.C., and Kumar, S., 1987, The closing of Tethys and the tectonics of the Himalaya: Geological Society of America Bulletin, v. 98, p. 678–701, doi:10.1130/0016 -7606(1987)982.0.CO;2. Sengör, A.M.Ç., 1979, Mid-Mesozoic closure of Permo-Triassic Tethys and its implications: Nature, v. 279, p. 590–593, doi:10.1038/279590a0. Sengör, A.M.Ç., 1984, The Cimmeride Orogenic System and the Tectonics of Eurasia: Geological Society of America Special Paper 195, 82 p. Sengör, A.M.Ç., 1985, The story of Tethys: How many wives did Okeanos have?: Episodes, v. 8, p. 3–12. Sengör, A.M.Ç., 1997, Asia, in Moores, E.M., and Fairbridge, F.W., eds., Encyclopedia of European and Asian Regional Geology: New York, Chapman & Hall, p. 34–51, doi:10.1007/1-4020-4495-X_6. Şengör, A.M.Ç., and Kidd, W.S.F., 1979, Post-collision tectonics of the Turkish-Iranian plateau and a comparison with Tibet: Tectonophysics, v. 55, p. 361–376, doi:10.1016/0040-1951(79)90184-7. Şengör, A.M.Ç., Yilmaz, Y., and Ketin, I., 1980, Remnants of a pre-late Jurassic ocean in northern Turkey: Fragments of a Permian-Triassic PalaeoTethys?: Geological Society of America Bulletin, v. 91, part I, no. 10, p. 599–609, doi:10.1130/0016-7606(1980)912.0.CO;2. Shakerardakani, F., Neubauer, F., Masoudi, F., Mehrabi, B., Liu, X., Dong, Y., Mohajjel, M., Monfaredi, B., and Friedl, G., 2015, Pan-African basement and Mesozoic gabbro in the Zagros orogenic belt in the Dorud–Azna region (NW Iran): Laser-ablation ICP–MS zircon ages and geochemistry: Tectonophysics, v. 647–648, no. 8, p. 146–171, doi:10.1016/j.tecto .2015.02.020. Sorkhabi, R., 1995, From Eduard Suess to Emile Argand: Reflections on Tethys and India-Asia tectonics: Current Science, v. 68, p. 853–858. Sorkhabi, R., 1997, Emile Argand and Tethyan tectonics of Asia: Himalayan Geology, v. 18, p. 17–32. Sorkhabi, R., and Heydari, E., eds., 2008, Asia out of Tethys: Geochronologic, Tectonic, and Sedimentary Records: Tectonophysics, v. 451 (special issue), 392 p. Stahl, A.F. von, 1911, Persien, in Handbuch der regionalen Geologie V, Abt. 6, Heft 8, Berlin, p. 1–46. Stamp, L.D., 1962, Asia: A Regional and Economic Geography (11th ed.): London, Methuen & Co., xvii + 370 p. Stern, R.J., and Johnson, P., 2010, Continental lithosphere of the Arabian plate: A geologic, petrologic, and geophysical synthesis: Earth-Science Reviews, v. 101, no. 1–2, p. 29–67, doi:10.1016/j .earscirev.2010.01.002. Stöcklin, J., 1974, Possible ancient continental margins in Iran, in Burke, C.A., and Drake, C.L., eds., The Geology of Continental Margins: New York, Springer-Verlag, p. 873–887, doi:10.1007/978-3-662-01141-6_64. Stöcklin, J., 1977, Structural correlation of the Alpine ranges between Iran and central Asia: Paris, Mémoire Hors Série Société Géologique de France, v. 1, no. 8, p. 333–353. Suess, E., 1885–1909, Das Antlitz der Erde: Prague, F. Tempsky, vol. 1 (1885, 778 p.), vol. 2 (1888, 703 p.), vol. 3, part I (1901, 508 p.), vol. 3, part II (1909, 789 p.) Suess, E., 1904–1924, The Face of the Earth, 5 vols., translated by H.B.C. Sollas: Oxford, UK, Clarendon Press, vol. 1 (1904, 604 p.), vol. 2 (1906, 556 p.), vol. 3 (1908, 400 p.), vol. 4 (1909, 673 p.), vol. 5 (1925, index, 170 p.) Takin, M., 1972, Iranian geology and continental drift in the Middle East: Nature, v. 235, p. 147–150, doi:10.1038/235147a0. Taylor, F.B., 1910, Bearing of the Tertiary mountain belts and the origin of the Earth’s plan: Geological Society of America Bulletin, v. 21, p. 179–226, doi:10.1130/GSAB-21-179.

Introduction Wadia, D.N., 1931, The syntaxis of north-west Himalaya: Its rocks, tectonics and orogeny: Records of the Geological Survey of India, v. 65, p. 189–220. Wegener, A., 1915, Die Entstehung der Kontinente und Ozeane: Braunschweig, Fried. Vieweg und Sohn, 481 p.

13

Wegener, A., 1966, The Origin of the Continents and Oceans, translated by J. Biram from the 4th German edition (1929): London, Methuen; New York, Dover, 272 p. Yin, A., 2010, Cenozoic tectonic evolution of Asia: A preliminary synthesis: Tectonophysics, v. 488, p. 293–325, doi:10.1016/j.tecto.2009.06.002.

APPENDIX. SELECT BIBLIOGRAPHY OF BOOKS AND SPECIAL VOLUMES (IN ENGLISH) ON TECTONIC EVOLUTION AND SEISMICITY OF SOUTHWEST ASIA

Al Hosani, Kh., Roure, F., Ellision, R., and Lokier, S., eds., 2013, Lithosphere Dynamics and Sedimentary Basins: The Arabian Plate and Analogues: Springer, Berlin, Frontiers in Earth Science, 496 p., doi:10.1007/978-3 -642-30609-9. Al-Husseini, M., ed., 2004, Carboniferous, Permian and Early Triassic Arabian Stratigraphy: Manama, Bahrain, Gulf Petrolink, GeoArabia Special Publication 3, 221 p. Alizadeh, A., Guliyev, I.S., Kadirov, F.A., and Eppelbaum, L.V., 2016, Geosciences of Azerbaijan, Vol. 1, Geology: New York, Springer, 274 p. Alizadeh, A., Guliyev, I.S., Kadirov, F.A., and Eppelbaum, L.V., 2017, Geosciences of Azerbaijan, Vol. 2, Economic Geology and Applied Geophysics: New York, Springer, 340 p. Al-Naqib, K.M., 1967, Geology of the Arabian Peninsula—Southwestern Iraq: U.S. Geological Survey Professional Paper 560-G, p. 1–54. Alsharhan, A.S., and Nairn, A.E.M., 2003, Sedimentary Basins and Petroleum Geology of the Middle East: Amsterdam, Elsevier Science, 977 p. Alsharhan, A.S., and Scott, R.W., eds., 2000, Middle East Models of Jurassic/ Cretaceous Carbonate Systems: Society of Economic Paleontologists and Mineralogists (SEPM) Special Publication 69, 364 p. Ambraseys, N.N., and Melville, C.P., 1982, A History of Persian Earthquakes: Cambridge, UK, Cambridge University Press, 236 p. Anonymous, ed., 1970, On the Structure and Evolution of the Red Sea and the Nature of the Red Sea, Gulf of Aden and Ethiopia Rift Junction: Philosophical Transactions of the Royal Society, London, ser. A, v. 267 (special issue), no. 1181, p. 1–417. Anonymous, ed., 1980, Geodynamic Evolution of the Afro-Arabian Rift System (International Meeting, Rome, 18–20 April 1979): Rome, Accademia Nazionale Dei Lincei, Rome, 705 p. Argand, E., 1924, La tectonique de l’Asie: Comptes Rendus de la 13e Congrès Géologique International, Belgique (1922), Vaillant-Carmanne, Liège, p. 171–372. Argand, E., 1977, The Tectonics of Asia, translated by A.V. Carozzi: New York, Hafner, 218 p. Aubion, J., Le Pichon, Z., and Monin, A.S., eds., 1986, Evolution of the Tethys: Tectonophysics, v. 123, no. 1–4 (special issue), p. 1–315. Audley-Charles, M.G., and Hallam, A., eds., 1988, Gondwana and Tethys: Oxford, UK, Oxford University Press, 317 p. Ayele, A., Abdelsalam, M.G., and Yirgu, G., eds., 2007, The East African Rift System: Dynamics, Evolution, and Environment: Journal of African Earth Science, v. 48, no. 203 (special issue), p. 59–246. Barrier, E., and Vrielynck, B., eds., 2008, Paleotectonic Maps of the Middle East. Tectono-sedimentary Palinspastic Maps from Late Norian to Piacenzia: Commission for the Geological Map of the World (CGMW/CCGM)/ UNESCO, atlas of 14 maps, scale 1:18,500,000. Barrier, E., Bergerat, F., Angelier, J., and Granath, J.W., eds., 2002, Paleostress and Tectonics in the Peri-Tethyan Margins: Tectonophysics, v. 357, no. 1–4 (special issue), p. 1–294. Ben-Avraham, Z., ed., 1987, Sedimentary Basins within the Dead Sea and Other Rift Zones: Tectonophysics, v. 141, no. 1–3 (special issue), p. 1–275. Bender, F., 1974, Geology of Jordan (Beiträge zur regionalen Geologie der Erde, Bd. 7): Berlin, Gebrüder Borntraeger, xi + 196 p. Berberian, M., 1976, Contribution to the Seismotectonics of Iran, part II: Tehran, Geological Survey of Iran, no. 39, 518 p. Berberian, M., 1977, Contribution to the Seismotectonics of Iran, part III: Tehran, Geological Survey of Iran, Tehran, 300 p. Berberian, M., 1983, Continental Deformation in the Iranian Plateau (Contribution to the Seismotectonics of Iran, part IV; Berberian, M., ed.): Tehran, Geological Survey of Iran, no. 52, 625 p. [in English] + 74 p. [in Persian].

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084703/spe525-01.pdf by University of Wisconsin Madison user

Berberian, M., 2014, Earthquakes and Coseismic Surface Faulting on the Iranian Plateau; A Historical, Social, and Physical Approach: Amsterdam, Elsevier, Developments in Earth Surface Processes 17, 714 p. Berberian, M., Qorashi, M., Arzhangravesh, B., and Mohajer-Ashjai, A., 1983, Recent Tectonics, Seismotectonics, and Earthquake-Fault Hazard Study of the Greater Qazvin Region (Contribution to the Seismotectonics of Iran, Part VI; Berberian, M., ed.): Geological Survey of Iran, no. 61, 197 p. [in Persian]. Reprinted in 1993. Berberian, M., Qorashi, M., Arzhangravesh, B., and Mohajer-Ashjai, A., 1985, Recent Tectonics, Seismotectonics, and Earthquake-Fault Hazard Study of the Greater Tehran Region, Tehran Quadrangle Area (Contribution to the Seismotectonics of Iran, Part V; Berberian, M., ed.): Tehran, Geological Survey of Iran, no. 56, 316 p. [in Persian]. Reprinted in 1993. Berberian, M., Ghorashi, M., Talebian, M., and Shoja-Taheri, J., 1996, Seismotectonic and Earthquake-Fault Hazard Investigations in the Semnan Quadrangle Area (Contribution to the Seismotectonics of Iran, Part VII; Berberian, M., ed.): Tehran, Geological Survey of Iran, no. 63, 266 p. [in Persian]. Berckhemer, H., and Hsü, K.J., eds., 1982, Alpine-Mediterranean Geodynamics (Inter-Union Commission on Geodynamics Working Group 3): American Geophysical Union Geodynamics Series 7, 216 p. Beydoun, Z.R., 1964, The stratigraphy and structure of the Eastern Aden Protectorate: Overseas Geology and Mineral Resources: Supplement Series, London, Bulletin Supplement No. 5, 107 p. Beydoun, Z.R., 1966, Geology of the Arabian Peninsula, Eastern Aden Protectorate and Part of Dhufar: U.S. Geological Survey Professional Paper 560-H, 49 p. Beydoun, Z.R., 1988, The Middle East: Regional Geology and Petroleum Resources: Beaconsfield, Bucks, UK, Scientific Press, 292 p. Beydoun, Z.R., 1991, Arabian Plate Hydrocarbon Geology and Potential: A Plate Tectonic Approach: American Association of Petroleum Geologists Studies in Geology 55, 77 p. Beydoun, Z.R., Al-Saruri, M., El-Nakhal, H., Al-Ganad, I.N., Baraba, R.S., Nani, A.S.O., and Al-Aawah, M.H., 1988, International Lexicon of Stratigraphy (Volume III, Asia), Republic of Yemen. No. 34: IUGS Publication No. 34 and Ministry of Oil and Mineral Resources, Republic of Yemen, Sana’a, xvi + 245 p. Bhat, G.M., Craig, J., Thurow, J.W., Thusu, B., and Cozzi, A., eds., 2012, Geology and Hydrocarbon Potential of Neoproterozoic-Cambrian Basins in Asia: Geological Society, London, Special Publication 366, viii + 304 p. Biju-Duval, B., and Montadert, L., eds., 1977, Structural History of the Mediterranean Basins: International Symposium (Split, Yugoslavia, 25–29 October 1976): Paris, Technip, xi + 448 p. Bonatti, E., ed., 1988, Zabargad Island and the Red Sea Rift: Tectonophysics, v. 150, no. 1–2 (special issue), p. 1–260. Boschi, E., Mantovani, E., and Morelli, A., eds., 1993, Recent Evolution and Seismicity of the Mediterranean Region (NATO Advanced Research Workshop on Recent Evolution and Seismicity of the Mediterranean Region, Erice, Italy, 1992): Dordrecht, Netherlands, Kluwer Academic, xix + 422 p. Bosence, D.W.J., ed., 1997, Mesozoic Rift Basins of Yemen: Marine and Petroleum Geology, v. 14, no. 6 (special issue), p. 611–730. Boudier, F., and Nicolas, A., eds., 1988, The Ophiolites of Oman: Tectonophysics, v. 151, no. 1–4 (special issue), p. 1–401. Bowen, R., and Jux, U., 1987, Afro-Arabian Geology: A Kinematic View: London, Chapman and Hall, xiv + 296 p. Bozkut, E., Winchester, J.A., and Piper, J.D.A., eds., 2000, Tectonics and Magmatism in Turkey and the Surrounding Area: Geological Society, London, Special Paper 173, 540 p.

14

R. Sorkhabi

Breckhemer, H., and Hsü, K., eds., 1991, Alpine-Mediterranean Geodynamics: Washington, D.C., American Geophysical Union Geodynamics Series 7, 216 p. Brinkmann, R., 1976, Geology of Turkey: Amsterdam, Elsevier, ix + 156 p. Brunet, M.-F., Wilmsen, M., and Granath, J.W., eds., 2009, South Caspian to Central Iran Basins: Geological Society, London, Special Publication 312, 352 p. Buday, T., 1980, The Regional Geology of Iraq, Vol. 1: Stratigraphy and Paleogeography: Baghdad, GEOSURV Publications, 445 p. Buday, T., and Jassim, Z.A., 1987, The Regional Geology of Iraq, Vol. 2: Tectonism, Magmatism and Metamorphism: Baghdad, GEOSURV Publications, 352 p. Campbell, A.S., 1971, Geology and History of Turkey: Tripoli, Petroleum Exploration Society of Libya, xiv + 511 p. Cavazza, W., Roure, F.M., Spakman, W., Stampfli, G.M., and Ziegler, P.A., eds., 2004, The TRANSMED Atlas—The Mediterranean Region from Crust to Mantle: Berlin, Springer (CD-ROM). Çemen, I., Helvaci, C., and Ersoy, E.Y., eds., 2014, Cenozoic Extensional Tectonics in Western and Central Anatolia, Turkey: Tectonophysics, v. 635 (special issue), p. 1–164. Clift, P., Kroon, D., Gaedicke, C., and Craig, J., eds., 2003, The Tectonic and Climatic Evolution of the Arabian Sea Region: Geological Society, London, Special Publication 195, 512 p. Coleman, R.G., 1993, Geologic Evolution of the Red Sea: New York, Oxford University Press, x + 186 p. Coward, M.P., and Ries, A.C., eds., 1986, Collision Tectonics: Geological Society, London, Special Publication 19, 432 p. Coward, M.P., Dietrich, D., and Park, R.G., eds., 1989, Alpine Tectonics: Geological Society, London, Special Publication 45, 456 p. Crasquin-Soleau, S., and Barrier, E., eds., 1998, Epicratonic Basins of PeriTethyan Platforms: Peri-Tethys Memoir 4: Paris, Mémoires du Muséum National d’Histoire Naturelle 179, 294 p. Crasquin-Soleau, S., and Barrier, E., eds., 1998, Stratigraphy and Evolution of Peri-Tethyan Platforms: Peri-Tethys Memoir 3: Paris, Mémoires du Muséum National d’Histoire Naturelle 177, 262 p. Crasquin-Soleau, S., and Barrier, E., eds., 2000, New Data on Peri-Tethyan Sedimentary Basins: Peri-Tethys Memoir 5: Paris, Mémoires du Muséum National d’Histoire Naturelle 182, 266 p. Dercourt, J., Ricou, L.E., and Vrielynck, B., eds., 1993, Atlas Tethys Paleoenviromental Maps: Paris, Gauthier-Villars, 307 p.+ 14 maps, 1 plate. Dercourt, J., Gaetani, M., Vrielynck, B., Barrier, E., Biju-Duval, B., Brunet, M.-F., Cadet, J.-P., Crasquin, S., and Sandulescu, M., eds., 2000, Atlas Peri-Tethys Palaeogeographical Maps: Paris, CCGM/CGMW, 268 p. + 24 maps. Dewey, J.F., ed., 1991, Special Issue on Tectonics: Bulletin of the Technical University of Istanbul (Istanbul Teknik Universitesi Bulletin), v. 44, no. 1–2, 342 p. Dilek, Y., and Pavlides, S., eds., 2006, Postcollisional Tectonics and Magmatism in the Mediterranean Region and Asia: Geological Society of America Special Paper 409, 644 p. Dixon, J.E., and Robertson, A.H.F., eds., 1984, The Geological Evolution of the Eastern Mediterranean: Geological Society, London, Special Publication 17, 824 p. Durand, B., Jolivet, L., Horvath, F., and Serrane, M., eds., 1999, The Mediterranean Basins: Tertiary Extension within the Alpine Orogen: Geological Society, London, Special Publication 156, 570 p. Edgell, H.S., 2006, Arabian Deserts: Nature, Origin and Evolution: Dordrecht, Netherlands, Springer, 592 p., doi:10.1007/1-4020-3970-0. Eppelbaum, L.V., and Khesin, B., 2012, Geophysical Studies in the Caucasus: Berlin, Springer, 404 p., doi:10.1007/978-3-540-76619-3. Erler, A., ed., 1995, Geology of the Black Sea Region (Proceedings of the International Symposium on the Geology of the Black Sea Region, September 7–11, 1992, Ankara, Turkey): Ankara, Turkey, General Directorate of Mineral Research and Exploration and Chamber of Geological Engineers (Maden Tetkik ve Arama Genel Müdürlüğü; Jeoloji Mühendisleri Odası), 334 p. Flower, M.F.J., Chi, C.T., Yem, N.T., and Mocanu, V.I., eds., 2004, Asthenosphere-Lithosphere Dynamic Responses to Tethyan Plate Collisions: Tectonophysics, v. 393, no. 1–4 (special issue), p. 1–328. Follmi, K.B., ed., 1989, Evolution of the Mid-Cretaceous Triad: Platform Carbonates, Phosphatic Sediments, and Pelagic Carbonates along the Northern Tethys Margin: Berlin, Springer, 153 p., doi:10.1007/BFb0019667.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084703/spe525-01.pdf by University of Wisconsin Madison user

Garfunkel, Z., ed., 1981, The Dead Sea Rift: Tectonophysics, v. 80, no. 1–4 (special issue), p. 1–334. Garfunkel, Z., Ben-Avraham, Z., and Kagan, E., eds., 2014, Dead Sea Transform Fault System: Reviews: New York, Springer, 359 p. Gee, D., and Stephenson, R., eds., The European Lithosphere Dynamics: Geological Society, London, Memoir 32, 668 p. Geukens, F., 1966, Geology of the Arabian Peninsula, Yemen: U.S. Geological Survey Professional Paper 560-B, 23 p. Ghorbani, M., 2013, The Economic Geology of Iran: Mineral Deposits and Natural Resources: Dordrecht, Netherlands, Springer, 572 p., doi:10.1007/978-94-007-5625-0. Glennie, K.W., 2007, Oman’s Geological Heritage (2nd ed.): London, Stacey International, 250 p. Glennie, K.W., Boeuf, M.G.A., Hughes Clarke, M.W., Moody Stuart, M., Pilaar, W.F.H., and Reinhardt, B.M., 1974, Geology of Oman Mountains. Verhandelingen van het Koninklijk Nederlands geologisch minjbouwkundig Genootschap (Transactions of the Royal Dutch Geological and Mining Society) 31, part I (text): 423 p., part II (tables and illustrations), part III (enclosures). Göncüoğlu, M.Ç., and Derman, A.S., eds., 1997, Lower Paleozoic Evolution in Northwest Gondwana: Turkish Association Petroleum Geology Special Publication 3, 190 p. Görür, N., Papadopoulos, G.A., and Okay, N., eds., 2002, Integration of Earth Science Research on the Turkish and Greek 1999 Earthquakes (2nd ed.): Berlin, Springer, NATO Science Series IV, v. 9, 207 p. Grainger, D., ed., 2007, The Geologic Evolution of Saudi Arabia: A Voyage through Space and Time, vol. 1, 264 p. + The Geologic Evolution of Saudi Arabia: Geologic Excursion and Geonotes, vol. 2, 6 geological excursions + 16 Geonotes: Saudi Geological Survey, Jeddah. Gupta, H.K., and Delany, F.M., eds., 1981, Zagros, Hindu Kush, Himalaya: Geodynamic Evolution: American Geophysical Union Geodynamics Series 3, 323 p. Holland, C.H., ed., 1981, Lower Palaeozoic of the Middle East, Eastern and Southern Africa, and Antarctica: Chichester, UK, John Wiley & Sons, Lower Palaeozoic Rocks of the World 3, xii + 332 p. Homberg, C., and Bachmann, M., eds., 2010, Evolution of the Levant Margin and Western Arabia Platform since the Mesozoic: Geological Society, London, Special Publication 341, 352 p. International Geological Congress, ed., 1984, Tectonics of Asia: Colloquium 5, 27th International Geological Congress, Moscow, Nauka, 240 p. Jackson, M.P., 1990, Salt Diapirs of the Great Kavir, Central Iran: Geological Society of America Memoir 177, 140 p., doi:10.1130/MEM177-p1. Jassim, S.Z., and Goff, J.C., eds., 2006, Geology of Iraq: Brno, Czech Republic, Dolin, Prague and Moravian Museum, 341 p. Khain, V.E., 1985, Geology of the USSR, Part I. Old Cratons and Palaeozoic Fold Belts: Berlin, Gebrüder Borntraeger, 260 p. Khain, V.E., 1994, Geology of the USSR, Part II. Geology of Northern Eurasia (ex-USSR): Berlin, Gebrüder Borntraeger, 408 p. Kinsley, D.B., 1970, Geomorphological and Paleontological Study of the Playas of Iran: Washington, D.C., Geological Survey, U.S. Department of Interior, 2 vols. Kovach, R.L., and Ben-Avraham, Z., eds., 1990, Geologic and Tectonic Processes of the Dead Sea Rift Zone: Tectonophysics, v. 180, no. 1 (special issue), p. 1–137. Krasheninnikov, V.A., and Hall, J.K., 2005, Geological Framework of the Levant, Vol. 1. Cyprus and Syria; Vol. 2. The Levantine Basin and Israel; Vol. 3. Plates: Universiṭat Ben-Guryon ba-Negev, Department of Geological & Environmental Sciences; Universiṭat Tel-Aviv, Department of Geophysics and Planetary Sciences; Historical Productions-Hall, Jerusalem. Le Fort, P., Colchen, M., and Montenat, C., eds., 1986, Orogenic Evolution of Southern Asia (from Turkey to Indonesia): Mémoires Sciences de la Terre: Nancy, France, Fondation Scientifique de la Géologie 47, 429 p. Le Pichon, X., and Cochran, J.R., eds., 1988, The Gulf of Suez and Red Sea Rifting: Tectonophysics, v. 153, no. 1–4 (special issue), p. 1–320. Leturmy, P., and Robin, C., eds., 2010, Tectonic and Stratigraphic Evolution of Zagros and Makran during the Mesozoic-Cenozoic: Geological Society, London, Special Publication 330, 376 p. Lippard, S.J., Shelton, A.W., and Gass, I.G., 1986, The Ophiolite of Northern Oman: Geological Society, London, Memoir 11, 186 p. Margottini, C., and Serva, L., eds., 1987, Historical Seismicity of CentralEastern Mediterranean Region: Italian Commission for Nuclear and Alternative Energy Sources and International Atomic Energy Agency, ENEA-IAEA Workshop, Rome, 27–29 October 1987, 337 p.

Introduction Markis, J., Mohr, P., and Rihm, R., eds., 1991, Red Sea: Birth and Early History of a New Ocean Basin: Tectonophysics, v. 198 (special issue), no. 2–4, p. 129–468. Marlow, L., Kendall, C.C.G., and Yose, L.A., eds., 2014, Petroleum Systems of the Tethyan Region: American Association of Petroleum Geologists Memoir 106, 780 p., doi:10.1306/M1061343. Mohammed, A.R., Hussey, J.A., and Reindl, E., 1997, Catalogue of Oman Lithostratigraphy: Muscat, Petroleum Development Oman, 373 p. + 43 plates (CD-ROM). Monin, A.S., and Zonenshain, L.P., eds., 1987, Geological History of the Tethys Ocean: Moscow, Academy of Sciences of the USSR, P.P. Shirshov Institute of Oceanology, 155 p. Moores, E.M., and Fairbridge, F.W., eds., Encyclopedia of European and Asian Regional Geology: New York, Chapman & Hall, 804 p. Morris, A., and Tarling, D.H., eds., 1996, Paleomagnetism and Tectonics of the Mediterranean Region: Geological Society, London, Special Publication 105, 422 p. Nairn, A.E.M., and Stehli, F.G., eds., 1977, The Eastern Mediterranean (Vol. 4A. The Ocean Basins and Margins): New York, Plenum Press, 503 p., doi:10.1007/978-1-4684-3036-3. Nairn, A.E.M., and Stehli, F.G., eds., 1978, The Western Mediterranean (Vol. 4B. The Ocean Basins and Margins): New York, Plenum Press, 447 p., doi:10.1007/978-1-4684-3039-4. Nairn, A.E.M., Ricou, L.-E., Vrielynck, B., and Dercourt, J., eds., 1996, The Tethys Ocean (Vol. 8. The Ocean Basins and Margins): New York, Plenum Press, 530 p. Nakazawa, K., and Dickins, J.M., eds., 1985, The Tethys: Her Paleogeography and Paleobiogeography from Paleozoic to Mesozoic: Tokyo, Tokai University Press, 317 p. Oberlander, T., 1965, The Zagros Streams: A New Interpretation of Transverse Drainage in an Orogenic Zone: Syracuse, New York, Syracuse University Press, Syracuse Geographical Series 1, xiv + 168 p. Okada, H., and Mateer, N.J., eds., 2000, Cretaceous Environments of Asia: Amsterdam, Elsevier Science, 264 p. Oren, O.H., and Mart, Y., eds., 1978, Structure and Tectonics of the Eastern Mediterranean: Tectonophysics, v. 46, no. 3–4 (special issue), p. 219–408. Peron-Pinvidic, G., and Osmundsen, P.T., eds., 2013, The Gulf of Aden Rifted Margin System: Special issue dedicated to the YOCMAL project (Young Conjugate Margins Laboratory in the Gulf of Aden): Tectonophysics, v. 607 (special issue), p. 1–114. Persits, F.M., Ulmishek, G.F., and Steinshouer, D.W., 1998, Map Showing Geology, Oil and Gas Fields and Geologic Provinces of the Former Soviet Union: U.S. Geological Survey Open-File Report 97-470E (CD-ROM). Peterson, J., 1993, Regional Geology and Hydrocarbon Resource Potential, the Mediterranean Sea Region: Denver, Colorado, U.S. Geological Survey, 118 p. + CD-ROM. Pollastro, R.M., Karshbaum, A.S., and Viger, R.J., 1998, Map Showing Geology, Oil and Gas Fields, and Geological Provinces of the Arabian Peninsula: U.S. Geological Survey Open-File Report 97-470B (CD-ROM). Pollastro, R.M., Persits, F.M., and Steinshouer, D.W., 1999, Map Showing Geology, Oil and Gas Fields, and Geological Provinces of Iran: U.S. Geological Survey Open-File Report 97-470G (CD-ROM). Powers, R.-W., 1968, Saudi Arabia (excluding Arabian Shield): Lexique stratigraphique international–Asie, Paris, vol. III, fasc. 10, b. 1, 177 p. Powers, R.-W., Ramirez, L.F., Redmond, C.D., and Elberg, E.L., Jr., 1966, Geology of the Arabian Peninsula, Sedimentary Geology of Saudi Arabia: U.S. Geological Survey Professional Paper 560-D, 147 p. Prodehl, C., Fuchs, K., Altherr, R., and Müller, B., eds., 1997, Structure and Dynamic Processes in the Lithosphere of the Afro-Arabian Rift System: Tectonophysics, v. 278, no. 1–4 (special issue), p. 1–352. Purser, B.H., and Bosence, D.W.J., eds., 1998, Sedimentation and Tectonics for Rift Basins: Red Sea–Gulf of Aden: London, Chapman & Hall, 832 p., doi:10.1007/978-94-011-4930-3. Raküs, M., Dercourt, J., and Nairn, A.E.M., eds., 1989, Evolution of the Northern Margin of Tethys: The Results of IGCP Project 198, Vol. 2: Mémoires de la Société Géologique de France, Nouvelle Série, 154, and Earth Sciences and Resources Institute, University of South Carolina, Occasional Publication 3. Ratschbacher, L., and Ben-Avraham, Z., eds., 1995, Kinematics of Distributed Deformation in Plate Boundary Zones with Emphasis on the Mediterranean, Anatolia and Eastern Asia: Tectonophysics, v. 243, no. 1–3 (special issue), p. 1–207.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084703/spe525-01.pdf by University of Wisconsin Madison user

15

Robertson, A.H.F., Searle, M.P., and Ries, A.C., eds., 1990, The Geology and Tectonics of the Oman Region: Geological Society, London, Special Publication 49, 845 p. Robertson, A.H.F., Parlak, O., and Koller, F., eds., 2009, Tethyan Tectonics of the Mediterranean Region: Some Recent Advances: Tectonophysics, v. 473, no. 1–2 (special issue), p. 1–282. Robertson, A.H.F., Parlak, O., and Ünlügeç, U.C., eds., 2013, Geological Development of Anatolia and the Eastern Mediterranean Region: Geological Society, London, Special Publication 372, 694 p. Robinson, A.G., ed., 1997, Regional and Petroleum Geology of the Black Sea and Surrounding Region: Tulsa, Oklahoma, American Association of Petroleum Geologists Memoir 68, 385 p. Roure, F., ed., 1994, Peri-Tethyan Platforms: Paris, Technip, 275 p. Roure, F., Amin, A.A., Khomsi, S., and Garni, M.A.M., eds., 2017, Lithosphere Dynamics and Sedimentary Basins of the Arabian Plate and Surrounding Areas: Berlin, Springer, Frontiers in Earth Sciences, 202 p., doi:10.1007/978-3-319-44726-1. Scheidegger, A.E., ed., 1980, Tectonic Stresses in the Alpine-Mediterranean Region, Proceedings of the symposium held in Vienna, Austria, September 12–14, 1979: Wien, Springer-Verlag, vi + 255 p. Sengör, A.M.Ç., 1984, The Cimmeride Orogenic System and the Tectonics of Eurasia: Geological Society of America Special Paper 195, 82 p. Sengör, A.M.Ç., ed., 1989, Tectonic Evolution of the Tethyan Region: Dordrecht, Netherlands, Kluwer Academic Publishers, 698 p., doi:10.1007/978-94 -009-2253-2. Şengör, A.M.Ç., 1990, Plate tectonic and orogenic research after 25 years: A Tethyan perspective: Earth-Science Reviews, v. 27, p. 1–201, doi:10.1016/0012-8252(90)90002-D. Shakerardakani, F., Neubauer, F., Masoudi, F., Mehrabi, B., Liu, X., Dong, Y., Mohajjel, M., Monfaredi, B., and Friedl, G., 2015, Pan-African basement and Mesozoic gabbro in the Zagros orogenic belt in the Dorud– Azna region (NW Iran): Laser-ablation ICP–MS zircon ages and geochemistry: Tectonophysics, v. 647–648, no. 8, p. 146–171, doi:10.1016/j .tecto.2015.02.020. Sharland, P.R., Archer, R., Casey, D.M., Davies, R.B., Hall, S.H., Heward, A.P., Horbury, A.D., and Simmons, M.D., 2001, Arabian Plate Sequence Stratigraphy: Manama, Bahrain, Gulf Petrolink, GeoArabia Special Publication 2, 372 p. Siadat, H., and Farshad, A., 2017, The Soil of Iran: Berlin, Springer, 255 p. Simmons, M.D., 1994, Micropaleontology and Hydrocarbon Exploration in the Middle East: New York, Springer, British Micropaleontological Society Publications Series, 418 p. Sobolev, V.S., Lepezin, G.G., and Dobretsov, N.I., eds., 1982, Metamorphic Complexes of Asia: Moscow, Nauka, 350 p. Sorkhabi, R., and Heydari, E., eds., 2008, Asia out of Tethys: Geochronologic, Tectonic, and Sedimentary Records: Tectonophysics, v. 451, no. 1–4 (special issue), 392 p. Sosson, M., Kaymakci, N., Stephenson, R.A., Bergerat, F., and Starostenko, V., eds., 2010, Sedimentary Basin Tectonics from the Black Sea and Caucasus to the Arabian Platform: Geological Society, London, Special Publication 340, xiii + 509 p. Stampfli, G.M., Borel, G.D., Cavazza, W., Mosar, J., and Ziegler, P.A., 2001, The Paleotectonic Atlas of the Peritethyan Domain: Strasbourg, Germany, European Geophysical Society, 330 p. (CD-ROM). Stanley, D.J., ed., 1973, The Mediterranean Sea: A Natural Sedimentation Laboratory (Symposium on the Mediterranean Sea, August 30–September 1, 1971, Heidelberg, 8th International Sedimentological Congress): Stroudsburg, Pennsylvania, Dowden, Hutchinson & Ross, xvi + 765 p. Stanley, D.J., and Wezel, F.-C., eds., 1985, Geological Evolution of the Mediterranean Basin: New York, Springer, 589 p., doi:10.1007/978-1-4613-8572-1. Steininger, F.F., ed., 1995, Neogene of the Mediterranean Tethys and Paratethys: Stratigraphic Correlation Tables and Sediment Distribution Maps, 2 vols.: Vienna, Institute of Paleontology, University of Vienna, International Geological Correlation Programme Project No. 25. Stöckalin, J., 1971, Stratigraphic Lexicon of Iran: Tehran, Geological Survey of Iran Report No. 18, 410 p. Suess, E., 1885–1909, Das Antlitz der Erde. F. Tempsky, Prague; vol. 1 (1885, 778 p.), vol. 2 (1888, 703 p.), vol. 3, part I (1901, 508 p.), vol. 3, part II (1909, 789 p.) Suess, E., 1904–1924, The Face of the Earth, 5 vols., translated by H.B.C. Sollas: Oxford, UK, Clarendon Press; vol. 1 (1904, 604 p.), vol. 2 (1906, 556 p.), vol. 3 (1908, 400 p.), vol. 4 (1909, 673 p.), vol. 5 (1925, index, 170 p.).

16

R. Sorkhabi

Sweet, W.C., Dickins, J.M., Yin, H., and Yang, Z., eds., 2003, Permo-Triassic Events in the Eastern Tethys: Stratigraphy, Classification and Relations with the Western Tethys: Cambridge, UK, Cambridge University Press, 195 p. Taymaz, T., Westaway, R., and Reilinger, R., eds., 2004, Active Faulting and Crustal Deformation in the Eastern Mediterranean Region: Tectonophysics, v. 391, no. 1–4 (special issue), p. 1–374. Tchalenko, J.S., and Berberian, M., 1974, Contribution to the Seismotectonics of Iran, Part I: Tehran, Geological Survey of Iran No. 29, 155 p. Tekeli, O., and Göncüoğlu, M.Ç., eds., 1984, Geology of the Taurus Belt (Proceedings of International Symposium on the Geology of the Taurus Belt, 26–29 September 1983, Ankara-Turkey): Maden Tetkik ve Arama Derĝisi, Ankara, Geological Survey of Turkey, 342 p. Toksoiz, M.N., 1983, Seismicity and Earthquake Prediction Studies in Turkey: Final Report to the U.S. Geological Survey: Boston, MIT Press. Turgut, S., ed., 1991, Tectonic and Hydrocarbon Potential of Anatolia and Surrounding Regions, Ozan Sungurlu Symposium Proceedings: Ankara, Turkey, Ozan Sungurlu Foundation for Science, Education and Aid, November 1991, 410 p. van Buchem, F.S.P., Gerdes, K.D., and Esteban, M., eds., 2010, Mesozoic and Cenozoic Carbonate Systems of the Mediterranean and the Middle East: Stratigraphic and Diagenetic Reference Models: Geological Society, London, Special Publication 329, 424 p.

van Hinsbergen, D.J.J., Edwards, M.A., and Govers, R., eds., 2009, Collision and Collapse at the Africa-Arabia-Eurasia Subduction Zone: Geological Society, London, Special Publication 311, 368 p. Vita-Finzi, C., 1969, The Mediterranean Valleys: Geological Changes in Historical Times: London, Cambridge University Press, ix + 140 p. Yilmaz, P.O., and Isaksen, G.H., eds., 2007, Oil and Gas of the Greater Caspian Area (AAPG Regional Conference, July 9–12, 2000, Istanbul): American Association of Petroleum Geologists Studies in Geology 55, 169 p. Yin, A., and Harrison, T.M., eds., 1996, The Tectonic Evolution of Asia: New York, Cambridge University Press, 666 p. Ziegler, P.A., and Horvath, F., eds., 1996, Structure and Prospects of Alpine Basins and Forelands: Peri-Tethys Memoir 2: Paris, Mémoires du Muséum National d’Histoire Naturelle 170, 552 p. Ziegler, P., Cavazza, W., Robertson, A.H.F., and Crasquin-Soleau, S., eds., 2001, Peri-Tethyan Rift/Wrench Basins and Passive Margins: Peri-Tethys Memoir 6: International Geological Correlation Programme 369: Paris, Mémoires du Muséum National d’Histoire Naturelle 186, 762 p. Zonenshain, L.P., Kuzmin, M.I., Natapov, L.M., and Page, B.M., eds., 1990, Geology of the USSR: A Plate-Tectonic Synthesis: Washington, D.C., American Geophysical Union Geodynamics Series 21, xviii + 242 p. MANUSCRIPT ACCEPTED BY THE SOCIETY 20 NOVEMBER 2017 MANUSCRIPT PUBLISHED ONLINE DECEMBER 2017

Printed in the USA

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084703/spe525-01.pdf by University of Wisconsin Madison user

The Geological Society of America Special Paper 525

Manuel Berberian: An appreciation and bibliography of his lifelong contribution to geoscience Rasoul Sorkhabi* University of Utah, Energy & Geoscience Institute, Salt Lake City, Utah 84108, USA

This Geological Society of America (GSA) Special Paper is dedicated to the lifelong geologic research and works of Dr. Manuel Berberian (Fig. 1), so I would like to say a few words about his life and to provide a bibliography of his works for the geologists active in his areas of research.

Manuel Berberian (Mānvel Mānugiān in Armenian) was born on 27 October 1945 in Tehran, Iran. He comes from an Armenian Christian family, one of oldest ethnicities living in Iran for millennia. His father immigrated to Iran at the age of 10, having lost his parents and siblings during the tragic massacre of Armenians in 1915 in the then–Ottoman Empire (now Turkey). Manuel’s mother, an Assyrian Christian, was born in the city of Urumieh (Urmia, later Rezāiyeh), Āzarbāijān Province, northwest Iran. Manuel and his two siblings lost their father when Manuel was 12 years old; their hardworking mother raised the family. Manuel graduated from the Ferdowsi primary school in 1958 and from the Khārazmi high school in 1964 in Tehran. After passing the university entrance examination (concours), Manuel then entered the Department of Geology at the University of Tehran and graduated with a B.S. in geology with highest academic achievements in the class of 1968. He was granted the Allied Corporation, USA, scholarship for postgraduate studies in the United States; unfortunately, the compulsory two-year military service laws prevented his departure from Iran or enrolling in the postgraduate courses in Iran. He joined the Geological Survey of Iran in Tehran in 1971, where he spent most of his professional career and where he established the Tectonics and Seismotectonics Research Department. His earliest papers were with John Tchalenko in 1973–1974. Over the years, he has collaborated with many scientists both in Iran and overseas (Figs. 2–5). In 1976, Manuel married his wife and lifelong partner Rose in Grenoble, France, where she had just completed her master’s degree in mineralogy from Institut Dolomieu, Université de Grenoble, France. After two years in Tehran, Manuel and his wife

Figure 1. Dr. Manuel Berberian in 2009 (courtesy of Manuel Berberian).

*[email protected] Sorkhabi, R., 2016, Manuel Berberian: An appreciation and bibliography of his lifelong contribution to geoscience, in Sorkhabi, R., ed., Tectonic Evolution, Collision, and Seismicity of Southwest Asia: In Honor of Manuel Berberian’s Forty-Five Years of Research Contributions: Geological Society of America Special Paper 525, p. 17–23, doi:10.1130/2016.2525(02). © 2016 The Geological Society of America. All rights reserved. For permission to copy, contact editing@ geosociety.org.

17

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084707/spe525-02.pdf by University of Wisconsin Madison user

18

R. Sorkhabi

Figure 2. Manuel Berberian (far right) during a fieldwork expedition in the extremely remote and uninhabitable mountains of Dehsard, eastern Lut Desert, eastern Iran, 1971 (courtesy of Manuel Berberian).

went to England in 1978, where he obtained a Ph.D. degree in earthquake seismology and active tectonics from the University of Cambridge in 1981 for his thesis on “Continental Deformation in the Iranian Plateau.” His doctorate advisors were Geoffrey King and Dan McKenzie. (Dr. Berberian was the first Armenian and the second Iranian graduate from the Earth Science Depart-

Figure 3. Manuel Berberian (right) and Arsalan Mohajer-Ashja’i (left) studying the Dasht-e Bayāz aftershocks and coseismic surface rupture in 1972 (courtesy of John Tchalenko).

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084707/spe525-02.pdf by University of Wisconsin Madison user

ment of the University of Cambridge. His wife also completed a Ph.D. program in petrology at the University of Cambridge, becoming the first Iranian female graduate from that university.) In 1981, they returned to Iran when the country was going through the disastrous war with Iraq (1980–1988). Dr. Berberian continued his work at the Geological Survey of Iran while also teaching B.S. and M.S. courses on structural geology, tectonics, and seismotectonics at the University of Tehran and University of Tarbiat Moddaress (“University Educator’s Training College”), where he supervised postgraduate dissertations. In 1990, the Berberians migrated to the United States (based on Manuel’s research publications, U.S. immigration and naturalization were granted to him, his wife, and their young son), where he worked as a consulting geologist in New Jersey (1990–2007) and as an instructor at the Ocean County College, Toms River, New Jersey (until 2015). Dr. Berberian is a certified professional geologist (CPG; American Institute of Professional Geologists [AIPG]) and a licensed professional geologist (PG, New Hampshire). Since 1971, Dr. Berberian has published (in English or Persian) more than 110 research papers and over 200 geological reports on site selection and earthquake-fault hazard estimation for major industrial plants. He has authored or edited 11 books in the fields of tectonic evolution, historical and modern earthquakes, seismotectonics, active faulting and folding, archaeoseismology, and the structural evolution of the Iranian Plateau. He has also prepared more than 32 geological, seismotectonic, and earthquake-fault hazard maps. Among his works is the eightvolume Contributions to Seismotectonics of Iran (Berberian, 1974–2000), which is an encyclopedic work in its field. Dr. Berberian is an internationally known figure in Iranian earthquake studies, seismotectonics, and tectonics. He is one of the pioneers who applied the plate-tectonic theory to interpreting

Manuel Berberian: An appreciation and bibliography of his lifelong contribution to geoscience

19

Figure 4. Manuel Berberian on a camel during fieldwork in the remote, uninhabitable, and impassable mountains north of the Jāz Muriān Desert, Baluchestān, SW Iran, 1973 (courtesy of Manuel Berberian).

Figure 5. Dr. Manuel Berberian at the Sixth World Conference on Earthquake Engineering in the presidential palace reception banquet, New Delhi, India, 1977 (courtesy of Manuel Berberian). Fakhruddin Ali Ahmed, India’s president from 1974–1977, greets the participants.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084707/spe525-02.pdf by University of Wisconsin Madison user

20

R. Sorkhabi

Figure 6. Some of the speakers and participants at the topical session “Tethyan evolution and seismotectonics of southwest Asia: In honor of the 40 years of Manuel Berberian’s research contributions,” 27–30 October 2013, Geological Society of America 125th Anniversary Annual Meeting in Denver (photo courtesy of Rasoul Sorkhabi). Seated in front chairs are (from left to right): Dr. Robert Yeats, Dr. Manuel Berberian, and Dr. Rose Berberian (Manuel’s spouse).

the geologic history of the Iranian Plateau; he has also played a pioneering role in archaeoseismological research work in Iran. His selfless devotion to geologic research in one of the tectonically complex and earthquake-hazardous parts of the world is admirable, and his legacy of research and publications is and will remain a valuable asset for generations of geologists working on the Iranian Plateau. His most recent publication, Earthquakes and Coseismic Surface Faulting on the Iranian Plateau: A Historical, Social, and Physical Approach (Berberian, 2014), of 776 pages, is a testimony to his lifelong, tireless devotion to geology. He is currently working on a comprehensive descriptive earthquake catalogue of the Iranian Plateau, which will span several volumes. In 1983, his book on Continental Deformation in the Iranian Plateau won the Book of the Year Award in Earth Sciences in Iran (Berberian, 1983). In 1998, the Geological Society of Iran presented him with the “Outstanding Earth Scientist Honor” for his three decades of research on earthquake science and tectonics

Figure 7. Dr. Manuel Berberian attending the 2013 Geological Society of American 125th Anniversary Annual Meeting in Denver (photo by Rasoul Sorkhabi).

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084707/spe525-02.pdf by University of Wisconsin Madison user

Manuel Berberian: An appreciation and bibliography of his lifelong contribution to geoscience of the Iranian Plateau. In 2008, a new fossil species of Chondrichthyes (jawed fish), Manberodus fortis gen. nov. (genus: Manberodus, family: Aztecondotidae, order: Omalodontiformes; age range: early Frasnian–Late Devonian), was named in honor of Manuel Berberian. Dr. Berberian has been a member of various professional societies, including the Geological Society of America (Figs. 6 and 7), American Geophysical Union, Seismological Society of America, and the Geological Society of London, and he is a member of the New York Academy of Sciences. He speaks English, Persian, Armenian, Assyrian, and Āzari Turkish. Dr. Manuel Berberian and Dr. Rose Berberian have a son (Sam Berberian), who graduated from the Massachusetts Institute of Technology in mechanical engineering and management sciences/finance. The couple currently live in Toms River, New Jersey, and their son resides in New York City. REFERENCES In writing this biography and bibliography, I benefited from conversations with Dr. Manuel Berberian, as well as the following interviews published in Persian professional journals: [1] Manuel Berberian’s interview with Earth Science and Mining Monthly Journal: Geological Survey of Iran, Tehran, 2009, v. 4, no. 39, p. 10–15. [2] Manuel Berberian’s interview with Tethys Geological Journal: Geological Society of the Beheshti University, Tehran, 2010, v. 1, p. 13–19. [3] Manuel Berberian’s interview with Safir Geological Journal: Eslāmshahr Azad University, 2011, v. 10, p. 8–15.

SELECTED BIBLIOGRAPHY OF WORKS The following bibliography is updated from Dr. Manuel Berberian’s website (http://manuelberberian.com) as well as academia.edu, researchgate.net, and other sources. The publications are in English unless otherwise specified.

Books Berberian, M., 1976, Contribution to the Seismotectonics of Iran, Part II: Tehran, Geological Survey of Iran, Report 39, 518 p. Berberian, M., 1977, Contribution to the Seismotectonics of Iran, Part III: Geological Survey of Iran, Tehran, Report 40, 300 p. Berberian, M., 1983, Continental Deformation in the Iranian Plateau (Contribution to the Seismotectonics of Iran, Part IV; Berberian, M., ed.): Tehran, Geological Survey of Iran, Report 52, 625 p. [in English] and 74 p. [in Persian]. Berberian, M., 1995, Natural Hazards and the First Earthquake Catalogue of Iran; Volume 1—Historical Hazards in Iran Prior to 1900: Tehran, International Institute of Earthquake Engineering and Seismology (UNESCO/ IIEES [United Nations Educational, Scientific and Cultural Organization/ Institute of Earthquake Engineering and Seismology] publication during UN/IDNDR [International Decade of Natural Disaster Reduction, 1900– 2000] period), 649 p. Berberian, M., 1997, An Investigation into the History of Cosmology and Earth Science in Ancient Iran: Tehran, Balkh/Neyshabur Foundation Publication, 573 p. [in Persian]. Berberian, M., 2014, Earthquakes and Coseismic Surface Faulting on the Iranian Plateau; A Historical, Social, and Physical Approach (Developments in Earth Surface Processes 17): Amsterdam, Netherlands, Elsevier, 776 p. Berberian, M., Qorashi, M., Arzhangravesh, B., and Mohajer-Ashjai, A., 1983, Recent Tectonics, Seismotectonics, and Earthquake-Fault Hazard Study of the Greater Qazvin Region (Contribution to the Seismotectonics of Iran, Part VI; Berberian, M., ed.): Tehran, Geological Survey of Iran, Report 61, 197 p. [in Persian]. Reprinted in 1993.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084707/spe525-02.pdf by University of Wisconsin Madison user

21

Berberian, M., Qorashi, M., Arzhangravesh, B., and Mohajer-Ashjai, A., 1985, Recent Tectonics, Seismotectonics, and Earthquake-Fault Hazard Study of the Greater Tehran Region, Tehran Quadrangle Area (Contribution to the Seismotectonics of Iran, Part V; Berberian, M., ed.): Tehran, Geological Survey of Iran, Report 56, 316 p. [in Persian]. Reprinted in 1993. Berberian, M., Ghorashi, M., Talebian, M., and Shoja-Taheri, J., 1996, Seismotectonic and Earthquake-Fault Hazard Investigations in the Semnan Quadrangle Area (Contribution to the Seismotectonics of Iran, Part VII; Berberian, M., ed.): Tehran, Geological Survey of Iran, Report 63, 266 p. [in Persian]. Berberian, M., Qorashi, M., Shoja-Taheri, J., and Talebian, M., 2000, Seismotectonics and Earthquake-Fault Hazard Investigations in the MashhadNeyshabur Quadrangle Area (Contribution to the Seismotectonics of Iran, Part VIII; Berberian, M., ed.): Tehran, Geological Survey of Iran, Tehran, Report 72, 233 p. [in Persian]. Tchalenko, J.S., and Berberian, M., 1974, Contribution to the Seismotectonics of Iran, Part I: Tehran, Geological Survey of Iran, Report 29, 155 p.

Peer-Reviewed Papers Berberian, F., and Berberian, M., 1981, Tectono-plutonic episodes in Iran, in Gupta, H.K., and Delany, F.M., eds., Zagros–Hindu Kush–Himalaya Geodynamic Evolution: American Geophysical Union Geodynamics Monograph 3, p. 5–32, doi:10.1029/GD003p0005. Berberian, M., 1979a, Tabas-e Golshan (Iran) catastrophic earthquake of September 16, 1978; a preliminary field report: Disaster, v. 2, no. 4, p. 207– 219, doi:10.1111/j.1467-7717.1978.tb00099.x. Berberian, M., 1979b, Evaluation of the instrumental and relocated epicentres of Iranian earthquakes: Geophysical Journal International, v. 58, p. 625– 630, doi:10.1111/j.1365-246X.1979.tb04798.x. Berberian, M., 1979c, Earthquake faulting and bedding thrust associated with the Tabas-e Golshan (Iran) earthquake of September 16, 1978: Bulletin of the Seismological Society of America, v. 69, no. 6, p. 1861–1887. Berberian, M., 1979d, Discussion of the paper A.A. Nowroozi, 1976 “Seismotectonic provinces of Iran”: Bulletin of the Seismological Society of America, v. 69, no. 1, p. 293–297. Berberian, M., 1981, Active faulting and tectonics of Iran, in Gupta, H.K., and Delany, F.M., eds., Zagros–Hindu Kush–Himalaya Geodynamic Evolution: American Geophysical Union Geodynamics Monograph 3, p. 33–69, doi:10.1029/GD003p0033. Berberian, M., 1982a, Aftershock tectonics of the 1978 Tabas-e-Golshan (Iran) earthquake sequence; a documented active ‘thin-and thick-skinned tectonic’ case: Geophysical Journal International, v. 68, p. 499–530, doi:10.1111/j.1365-246X.1982.tb04912.x. Berberian, M., 1982b, Discussion on the paper A. Mohajer-Ashjai and A.A. Nowroozi, “The Tabas Earthquake of September 16, 1978 in East-Central Iran”: Geophysical Research Letters, v. 9, no. 3, p. 193–194, doi:10.1029/ GL009i003p00193. Berberian, M., 1983, The Southern Caspian; a compressional depression floored by a trapped, modified oceanic crust: Canadian Journal of Earth Sciences, v. 20, no. 2, p. 163–183, doi:10.1139/e83-015. Berberian, M., 1984a, Active tectonics of Iran, in Proceedings of the 27th International Geological Congress, Volume IV, Section 08, Geophysics, Seismicity, Moscow [abs.]: Utrecht, the Netherlands, VNU Science Press BV, p. 38. Berberian, M., 1984b, Structural evolution and tectonics of the Iranian Plateau: A plate tectonic approach, in Proceedings of the 27th International Geological Congress, Volume IV, Section 07, Tectonics of Continental Fold Belts, Moscow [abs.]: Utrecht, the Netherlands, VNU Science Press BV, p. 140. Berberian, M., 1989, Tectonics evolution of Iranian mountain belts, in Proceedings of the 28th International Geological Congress, Washington, D.C. [abs.]: Utrecht, the Netherlands, VNU Science Press BV, v. 1 of 3, p. 129–130. Berberian, M., 1991, Is the theory of earthquake Greek or Iranian? The oldest theory on earthquakes and faulting: Iranshenasi [Bethesda, Maryland], v. II, no. 4, p. 835–845. Berberian, M., 1992, Biruni Kharazmi and the theory of continental drift: Iranshenasi, v. IV, no. 1, p. 139–147. Berberian, M., 1995, Master ‘blind’ thrust faults hidden under the Zagros folds; active basement tectonics and surface morphotectonics: Tectonophysics, v. 241, p. 193–224, doi:10.1016/0040-1951(94)00185-C.

22

R. Sorkhabi

Berberian, M., 1996, The historical record of earthquakes in Persia, in Yarshater, E., ed., Encyclopaedia Iranica, Volume VII, Fascicle 6, Drugs-Ebn al-Atir: Costa Mesa, California, Mazda Publishers, p. 635–640. Berberian, M., 1997, Seismic sources of the Transcaucasian historical earthquakes, in Giardini, D., and Balassanian, S., eds., Historical and Prehistorical Earthquakes in the Caucasus: Dordrecht, Netherlands, Kluwer Academic Press, North Atlantic Treaty Organization (NATO) Advanced Study Institute (ASI) Series 2: Environment, v. 28, p. 233–311. Berberian, M., 2003, Managing earthquakes in Iran: Iran Analysis Quarterly: Iranian Studies Group at MIT, v. 1, no. 2, p. 2–4. Berberian, M., 2005, The 2003 Bam urban earthquake; a predictable seismotectonic pattern along the western margin of the rigid Lut block, southeast Iran: Earthquake Spectra, v. 21, no. S1, p. S35–S99 (Earthquake Engineering Research Institute), doi:10.1193/1.2127909. Berberian, M., 2009, Bam Earthquake of December 26, 2003: Encyclopaedia Iranica, http://www.iranicaonline.org (accessed 20 July 2009). Berberian, M., 2013, Early earthquake detection and warning alarm system in Iran by a telegraph operator; a 116-year old disaster prevention attempt: Seismological Research Letters, v. 84, no. 5, p. 816–819, doi:10.1785/0220130068. Berberian, M., 2016, this volume, Chapter 3, Development of the geological perceptions and explorations on the Iranian Plateau: From Zoroastrian cosmogony to plate tectonics (ca. 1200 BCE to 1980 CE), in Sorkhabi, R., ed., Tectonic Evolution, Collision, and Seismicity of Southwest Asia: In Honor of Manuel Berberian’s Forty-Five Years of Research Contributions: Geological Society of America Special Paper 525, doi:10.1130/2016.2525(03). Berberian, M., and King, G.C.P., 1981a, Towards a paleogeography and tectonic evolution of Iran: Canadian Journal of Earth Sciences, v. 18, no. 2, p. 210–265, doi:10.1139/e81-019. Berberian, M., and King, G.C.P., 1981b, Towards a paleogeography and tectonic evolution of Iran [reply]: Canadian Journal of Earth Sciences, v. 18, no. 11, p. 1764–1766, doi:10.1139/e81-163. Berberian, M., and Navai, I., 1978, Naghan (Chahar Mahal Bakhtiari–High Zagros, Iran) earthquake of 6 April 1977: A preliminary field report and a seismotectonic discussion: Annali di Geofisica, v. 31, p. 5–27. Berberian, M., and Papastamatiou, D., 1978, Khurgu (north Bandar Abbas, Iran) earthquake of March 21, 1977; a preliminary field report and a seismotectonic discussion: Bulletin of the Seismological Society of America, v. 68, no. 2, p. 411–428. Berberian, M., and Qorashi, M., 1994, Coseismic fault-related folding during the South Golbaf earthquake of November 20, 1989, in southeast Iran: Geology, v. 22, p. 531–534, doi:10.1130/0091-7613(1994)0222.3.CO;2. Berberian, M., and Walker, R., 2010, The Rudbār Mw 7.3 earthquake of 1990 June 20; seismotectoniocs, coseismic and geomorphic displacements, and historic earthquakes of the western ‘High-Alborz’ of Iran: Geophysical Journal International, v. 182, no. 3, p. 1577–1602, doi:10.1111/j.1365 -246X.2010.04705.x. Berberian, M., and Yeats, R.S., 1999, Patterns of historical rupture in the Iranian Plateau: Bulletin of the Seismological Society of America, v. 89, no. 1, p. 120–139. Berberian, M., and Yeats, R.S., 2001, Contribution of archaeological data to studies of earthquake history in the Iranian Plateau: Paul Hancock Memorial Issue: Journal of Structural Geology, v. 23, p. 563–584, doi:10.1016/ S0191-8141(00)00115-2. Berberian, M., and Yeats, R.S., 2016, this volume, Chapter 4, Tehran: An earthquake time bomb, in Sorkhabi, R., ed., Tectonic Evolution, Collision, and Seismicity of Southwest Asia: In Honor of Manuel Berberian’s Forty-Five Years of Research Contributions: Geological Society of America Special Paper 525, doi:10.1130/2016.2525(04). Berberian, M., Asudeh, I., Bilham, R.G., Scholz, C.H., and Soufleris, C., 1979a, Mechanism of the main shock and the aftershock study of the Tabas-e-Golshan (Iran) earthquake of September 16, 1978; a preliminary report: Bulletin of the Seismological Society of America, v. 69, no. 6, p. 1851–1859. Berberian, M., Asudeh, I., and Arshadi, S., 1979b, Surface rupture and mechanism of the Bob-Tangol (southeastern Iran) earthquake of December 19, 1977: Earth and Planetary Science Letters, v. 42, no. 3, p. 456–462, doi:10.1016/0012-821X(79)90055-4. Berberian, F., Muir, I.D., Pankhurst, R.J., and Berberian, M., 1982, Late Cretaceous and early Miocene Andean-type plutonic activity in northern

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084707/spe525-02.pdf by University of Wisconsin Madison user

Makran and Central Iran: Journal of the Geological Society, London, v. 139, no. 5, p. 605–614, doi:10.1144/gsjgs.139.5.0605. Berberian, M., Jackson, J., Ghorashi, M., and Kadjar, M.H., 1984, Field and teleseismic observations of the 1981 Golbaf-Sirch earthquakes in SE Iran: Geophysical Journal International, v. 77, p. 809–838, doi:10.1111/j.1365 -246X.1984.tb02223.x. Berberian, M., Qorashi, M., Jackson, J.A., Priestley, K., and Wallace, T., 1992, The Rudbar-Tarom earthquake of June 20, 1990 in NW Persia; preliminary field and seismological observations, and its tectonic significance: Bulletin of the Seismological Society of America, v. 82, no. 4, p. 1726–1755. Berberian, M., Jackson, J.A., Qorashi, M., Khatib, M.M., Priestley, K., Talebian, M., and Ghafuri-Ashtiani, M., 1999, The 1997 May 10 Zirkuh (Qa’enat) earthquake (Mw 7.2); faulting along the Sistan suture zone of eastern Iran: Geophysical Journal International, v. 136, no. 3, p. 671–694, doi:10.1046/j.1365-246x.1999.00762.x. Berberian, M., Jackson, J.A., Qorashi, M., Talebian, M., Khatib, M., and Priestley, K., 2000, The 1994 Sefidabeh earthquakes in eastern Iran; blind thrusting and bedding-plane slip on a growing anticline, and active tectonics of the Sistan suture zone: Geophysical Journal International, v. 142, no. 2, p. 283–299, doi:10.1046/j.1365-246x.2000.00158.x. Berberian, M., Jackson, J.A., Fielding, E., Parsons, B.E., Priestly, K., Qorashi, M., Talebian, M., Walker, R., Wright, T.J., and Baker, E., 2001, The 1998 March 14 Fandoqa earthquake (Mw 6.6) in Kerman, southeast Iran; rerupture of the 1981 Sirch earthquake fault, triggering of slip on adjacent thrusts, and the active tectonics of the Gowk fault zone: Geophysical Journal International, v. 146, no. 2, p. 371–398, doi:10.1046/j.1365 -246x.2001.01459.x. Berberian, M., Malek Shahmirzadi, S., Nokandeh, J., and Djamali, M., 2012, Archeoseismicity and environmental crises at the Sialk mounds, Central Iranian Plateau, since Early Neolithic: Journal of Archaeological Science, v. 39, no. 9, p. 2845–2858, doi:10.1016/j.jas.2012.04.001. Berberian, M., Petrie, C.A., Potts, D.T., Asgari Chaverdi, A., Dusting, A., Sardari Zarchi, A., Weeks, L., Ghassemi, P., and Noruzi, R., 2014, Archaeoseismicity of the mounds and monuments along the Kāzerun fault (Western Zāgros, SW Iranian Plateau) since the Chalcolithic Period: Iranica Antiqua, v. 49, p. 1–81, doi:10.2143/IA.49.0.3009238. Berberian, M., Moqaddas, M., and Kabiri, A., 2016, this volume, Chapter 5, Archaeological and architectural evidence of historical seismic activity along the Zāgros Main Recent fault at Borujerd (western Iranian Plateau); the 1316 C.E. earthquake, in Sorkhabi, R., ed., Tectonic Evolution, Collision, and Seismicity of Southwest Asia: In Honor of Manuel Berberian’s Forty-Five Years of Research Contributions: Geological Society of America Special Paper 525, doi:10.1130/2016.2525(05). Bordet, P., Berberian, M., Alavi, Tehrani, N., and Lotfi, M., 1976, Sur la géologie du massif du Sahand (Azerbaidjian Iran occidental): Comptes Rendus de l’Académie des Sciences, Paris, ser. D, v. 283, no. 13, p. 1481–1484. Djamali, M., de Beaulieu, J.-L., Andrieu-Ponel, V., Berberian, M., Miller, N.F., Gandouin, E., Lahijani, H., Shah-Hosseini, M., Ponel, P., Salimian, M., and Guiter, F., 2009, A late Holocene pollen record from Lake Almalou in NW Iran: Evidence for changing land-use in relation to some historical events during the last 3700 years: Journal of Archaeological Science, v. 36, p. 1364–1375, doi:10.1016/j.jas.2009.01.022. Djamali, M., Miller, N.F., Ramezani, E., Andrieu-Ponel, V., de Beaulieu, J.-L., Berberian, M., Guibal, F., Lahijani, H., Lak, R., and Ponel, P., 2010, Notes on arboriculture and agricultural practices in ancient Iran based on new pollen evidence: Paléorient, v. 36, no. 2, p. 175–188. Hedayati, A., Brander, J.L., and Berberian, M., 1976, Microearthquake survey of Tehran region: Bulletin of the Seismological Society of America, v. 66, no. 5, p. 1713–1725. Jackson, J., Priestley, K., Allen, M., and Berberian, M., 2002, Active tectonics of the South Caspian Basin: Geophysical Journal International, v. 148, no. 2, p. 214–245. King, G., Soufleris, C., and Berberian, M., 1981, The source parameters, surface deformation and tectonic setting of three recent earthquakes, Thessaloniki (Greece), Tabas-e-Golshan (Iran), and Carlisle (U.K.): Disasters, v. 5, no. 1, p. 36–46, doi:10.1111/j.1467-7717.1981.tb01127.x. Mohajer-Ashjai, A., Behzadi, H., and Berberian, M., 1975, Reflections of the rigidity of the Lut block and recent crustal deformation in eastern Iran: Tectonophysics, v. 25, p. 281–301, doi:10.1016/0040-1951(75)90066-9. Moinfar, A.A., Berberian, M., Qorashi, M., Zohurian, A.A., and Naderzadeh, A., 1987, Preliminary Seismic Zoning and Earthquake Hazard in Iran;

Manuel Berberian: An appreciation and bibliography of his lifelong contribution to geoscience For Use in the Iranian Code for Seismic Design of Buildings: Tehran, Building and Housing Research Center, Ministry of Housing and Urban Development, Report 74, 35 p. [in Persian]. Naderi Beni, A., Lahijani, H., Mousavi Harami, R., Arpe, K., Leroy, S.A.G., Marriner, N., Berberian, M., Andrieu-Ponel, V., Djamali, M., Mahboubi, A., and Reimer, P.J., 2013, Caspian sea-level changes during the last millennium: Historical and geological evidence from the south Caspian Sea: Climate of the Past, v. 9, p. 1645–1665, doi:10.5194/cp-9-1645-2013. Petrie, C.A., Sardari, A., Ballantyne, R., Berberian, M., Lancelotti, C., Mashkour, M., McCall, B., Potts, D.T., and Weeks, L., 2013, Mamasani in the fourth millennium BC, in Petrie, C.A., ed., Ancient Iran and Its Neighbours; Local Developments and Long-Range Interactions in the Fourth Millennium BC: Oxford, UK, Oxbow Books, p. 171–194. Tchalenko, J.S., and Berberian, M., 1974, The Salmas (Iran) earthquake of May 6, 1930: Annali di Geofisica, v. 27, no. 1–2, p. 151–212. Tchalenko, J.S., and Berberian, M., 1975, Dasht-e-Bayaz fault, Iran, earthquake and earlier related structures in bedrock: Geological Society of America Bulletin, v. 86, p. 703–709, doi:10.1130/0016-7606(1975)862.0.CO;2. Tchalenko, J.S., Berberian, M., and Behzadi, H., 1973, Geomorphic and seismic evidence for recent activity on the Doruneh fault, Iran: Tectonophysics, v. 19, p. 333–341, doi:10.1016/0040-1951(73)90027-9. Tchalenko, J.S., Braud, J., and Berberian, M., 1974, Discovery of three earthquake faults in Iran: Nature, v. 248, p. 661–663, doi:10.1038/248661a0.

Maps Berberian, M., 1973, The Seismicity of Iran: Preliminary Map of Epicentres and Focal Depths: Tehran, Geological Survey of Iran, single color map, scale 1:2,500,000. Berberian, M., 1976a, First Seismotectonic Map of Iran (Contribution to the Seismotectonics of Iran, Part II): Tehran, Geological Survey of Iran, Report 39, color map, scale 1:2,500,000. Berberian, M., 1976b, Epicentre Map of Iran (1900–1976) (Contribution to the Seismotectonics of Iran, Part II): Tehran, Geological Survey of Iran, Report 39, color map, scale 1:5,000,000. Berberian, M., 1976c, Generalized Fault Map of Iran (Contribution to the Seismotectonics of Iran, Part II): Tehran, Geological Survey of Iran, Report 39, color map, scale 1:5,000,000. Berberian, M., 1976d, Areas of Destructive Earthquakes in Iran (4th Century B.C. to 1976 A.D.) (Contribution to the Seismotectonics of Iran, Part II): Tehran, Geological Survey of Iran, Report 39, color map, scale 1:5,000,000. Berberian, M., 1976e, Macroseismic Epicentres of Destructive and Damaging Earthquakes in Iran (1900–1976) (Contribution to the Seismotectonics of Iran, Part II): Tehran, Geological Survey of Iran, Report 39, color map, scale 1:5,000,000. Berberian, M., 1977a, Historical Seismicity (Pre-1900) Map of Iran (Contribution to the Seismotectonics of Iran, Part III): Tehran, Geological and Mining Survey of Iran, Report 40, color map, scale 1:5,000,000. Berberian, M., 1977b, Intensity Zone Map of Iran (4th Century B.C. to 1900 A.D.) (Contribution to the Seismotectonics of Iran, Part III): Tehran, Geological and Mining Survey of Iran, Report 40, color map, scale 1:5,000,000. Berberian, M., 1977c, Maximum Intensity of Earthquakes in Iran (1900–1977) (Contribution to the Seismotectonics of Iran, Part III): Tehran, Geological and Mining Survey of Iran, Report 40, color map, scale 1:5,000,000. Berberian, M., 1977d, Isoseismal Map of Iran (1900–1977) (Contribution to the Seismotectonics of Iran, Part III): Tehran, Geological and Mining Survey of Iran, Report 40, color map, scale 1:5,000,000.

23

Berberian, M., 1977e, Intensity Zone Map of Iran (1900–1977) (Contribution to the Seismotectonics of Iran, Part III): Tehran, Geological and Mining Survey of Iran, Report 40, color map, scale 1:5,000,000. Berberian, M., 1977f, Intensity Zone Map of Iran (4th Century B.C. to 1977 A.D.) (Contribution to the Seismotectonics of Iran, Part III): Tehran, Geological and Mining Survey of Iran, Report 40, color map, scale 1:5,000,000. Berberian, M., 1983a, Generalized Tectonic Map of Iran (Contribution to the Seismotectonics of Iran, Part IV): Tehran, Geological Survey of Iran, Report 52, scale 1:1,500,000. Berberian, M., compiler, 1983b, Geological Map of the Mirjaveh Area: Tehran, Geological Survey of Iran, Sheet 8248, color map, scale 1:100,000 Berberian, M., compiler, 1983c, Geological Map of the Zahedan Area: Tehran, Geological Survey of Iran, Sheet 8148, color map, scale 1:100,000 Berberian, M., 1983d, Active Fault Map and Seismicity of the Qazvin Quadrangle and Ipak Area (Contribution to the Seismotectonics of Iran, Part VI): Tehran, Geological Survey of Iran, Report 61, scale 1:250,000. Berberian, M., 1983e, Seismic Zoning Map of the Qazvin-Ipak Region (Contribution to the Seismotectonics of Iran, Part VI): Tehran, Geological Survey of Iran, Report 61, color map, scale 1:250,000. Berberian, M., 1985a, Seismicity and Active Fault Map of the Tehran Quadrangle (Contribution to the Seismotectonics of Iran, Part V): Tehran, Geological Survey of Iran, Report 56, scale 1:250,000. Berberian, M., 1985b, Active Fault Maps of the Tehran Quadrangle, and Tehran-Ray Region (Contribution to the Seismotectonics of Iran, Part V): Tehran, Geological Survey of Iran, Report 56, 3 maps, scales 1:50,000, 1:100,000, and 1:250,000. Berberian, M., 1990, Geological Map of Northeastern Jaz Murian Quadrangle: Tehran, Geological Survey of Iran, Report K13, color map, scale 1:250,000. Berberian, M., 2000, Active Fault and Historical Seismicity of the MashhadNeyshabur Quadrangle Area (Contribution to the Seismotectonics of Iran, Part VIII): Tehran, Geological Survey of Iran, Report 72, scale 1:250,000. Berberian, M., and Alavi-Tehrani, M., 1977, Structural Pattern of the Hamadan Metamorphic Tectonites Contribution to the Seismotectonics of Iran, Part III: Tehran, Geological and Mining Survey of Iran, Report 40, one map. Berberian, M., and Mohajer-Ashjai, A., 1977, Seismic Risk Map of Iran, A Proposal (Contribution to the Seismotectonics of Iran, Part III): Tehran, Geological and Mining Survey of Iran, Report 40, color map, scale 1:5,000,000. Berberian, M., and Soheili, M., 1991, Geological Map of Chardeh, East Dehsalm, Lut Block: Tehran, Geological Survey of Iran, Report K9, color map, scale 1:250,000. Mohajer-Ashjai, A., and Berberian, M., 1985, Seismic Zoning and Isoacceleration Contour Maps of the Tehran-Ray Region (Contribution to the Seismotectonics of Iran, Part V): Tehran, Geological Survey of Iran, Report 56, 2 maps, scale 1:100,000. Mohajer-Ashjai, A., and Berberian, M., 1983, Iso-Acceleration Contour Maps of the Qazvin-Ipak Region (Contribution to the Seismotectonics of Iran, Part VI): Tehran, Geological Survey of Iran, Report 61, scale 1:250,000. Moinfar, A.A., Berberian, M., Qorashi, M., Zohurian, A.A., and Naderzadeh, A., 1987, Seismic Zoning and Earthquake Hazard Map of Iran for Use in the Iranian Code for Seismic Design of Buildings: Tehran, Building and Housing Research Center, Ministry of Housing and Urban Development, Report 74, color map, scale 1:2,500,000. Tchalenko, J.S., Iranmanesh, H., and Berberian, M., 1974, Seismotectonic Map of North Central Iran (Central Alborz): Tehran, Geological Survey of Iran, Report 29, color map, scale 1:100,000. MANUSCRIPT ACCEPTED BY THE SOCIETY 15 JUNE 2016 MANUSCRIPT PUBLISHED ONLINE 8 NOVEMBER 2016

Printed in the USA

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084707/spe525-02.pdf by University of Wisconsin Madison user

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084707/spe525-02.pdf by University of Wisconsin Madison user

The Geological Society of America Special Paper 525

Development of geological perceptions and explorations on the Iranian Plateau: From Zoroastrian cosmogony to plate tectonics (ca. 1200 BCE to 1980 CE) Manuel Berberian* Fellow, School of Mathematics, Science, and Technology, Department of Science, Ocean County College, Toms River, New Jersey 08754-2001, USA, and Onduni Grung Scientific Enterprise, 1224 Fox Hollow Drive, Toms River, New Jersey 08755-2179, USA

ABSTRACT This is an in-depth review and analysis of the long and untold history of development of earth science, geological thinking, research, and exploration on the Iranian Plateau within its historical, political, and socioeconomic context. Widespread mineral resources and ancient civilization helped in exploration, excavation, smelting, and usage of different metals, precious stones, and minerals since the Neolithic Period. Extant ancient Avestan and Middle Iranian Pahlavi Zoroastrian texts, as well as the classic Greek and Roman scholars, clearly demonstrate the Iranian geological activity through the Median (ca. 615 BCE), Achaemenid (550–330 BCE), Parthian (250 BCE–224 CE), and Sassanid (224–642 CE) Dynasties, interrupted by disrupting periods of socioeconomic and political problems, followed by foreign invasions and devastation in 330 BCE–250 CE and 637–652 CE, when the Iranians could no longer make scientific advancements. Long after the invasion of Alexander III of Macedon (330 BCE), scientific activity culminated in the establishment of the academies of Gundishāpur, Ctesiphon, and Resaina, the three higher educational centers of the Sassanid Dynasty that focused on comprehensive observation, painstaking research, and advanced education during the sixth and seventh centuries CE. Careful observation, research, and experiment by brilliant and genius scholars such as Karaji, Biruni, and Avicenna took place during a period of great activity and growth in science, engineering, medicine, literature, art, architecture, and philosophy in the tenth and eleventh centuries CE in Iran. This Iranian two-century “intermezzo intellectual zenith,” with a stable state and economic prosperity, was nurtured by the vast heritage of the ancient Iranian, Mesopotamian, Indian, and Egyptian civilizations and elements of the ancient Avestan, Sanskrit, and Pahlavi writings since ca. 1200 BCE. Social, economic, and political conflicts followed by invasions by Central Asian nomadic tribe warlords and their accompanying hordes in 1000–1040 CE (Saljuqs), 1218–1231 CE, and 1256 CE (Mongols), and 1370 CE (Timurids), and their occupation caused the

*E-mails: [email protected], [email protected] (corresponding e-mail). Berberian, M., 2016, Development of geological perceptions and explorations on the Iranian Plateau: From Zoroastrian cosmogony to plate tectonics (ca. 1200 BCE to 1980 CE), in Sorkhabi, R., ed., Tectonic Evolution, Collision, and Seismicity of Southwest Asia: In Honor of Manuel Berberian’s Forty-Five Years of Research Contributions: Geological Society of America Special Paper 525, p. 25–85, doi:10.1130/2016.2525(03). © 2016 The Geological Society of America. All rights reserved. For permission to copy, contact [email protected].

25

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084715/spe525-03.pdf by University of Wisconsin Madison user

26

M. Berberian process of irreversible decay, retrogression, and general intellectual decadence until the Safavids (1491–1772 CE). During this relatively long dark period, there was a drastic decline in interest in geological research and writing, though some old mining efforts were active. Throughout the eighteenth to the mid-twentieth centuries, foreign travelers made some contributions to the geology and mineral resources of Iran. It was during the second half of the twentieth century when once again earth science research blossomed in Iran with the help of European geologists. This ushered in a new period of modern geologic studies of Iran by native geologists.

In memory of Emil Tietze (1845–1931), Alexander von Stahl (b. 1850), Setrāk Ābdāliān (1894–1963), Eugène Rieben (1899–1972), Heinrich Martin Huber (1917–1992), Jovan Stöcklin (1921–2008), Ricardo Assereto (1939–1976), and all pioneers in the past, who enthusiastically and rigorously intruded ever deeper into virtually unexplored territories in difficult and uncomfortable circumstances, extremely devoted to scientific pursuits, and shaped our understanding of the geology, tectonics, mineral resources, earthquakes, and seismotectonics of the Iranian Plateau.

1. INTRODUCTION1 The Iranian Plateau, with an average elevation of 1500 m above sea level (amsl; Damāvand: 5671 m, ‘Alamkuh: 4850 m, and Takht-e Solaymān: 4650 m in the Alborz; Denā: 4550 m in the Zāgros), is an extensive active area of crustal deformation, seismic activity, and mineral resources located within the convergence zone between the stable Arabian and Eurasian plates. The Iranian Plateau, with different metallogenic provinces, incorporates a mosaic of several continental blocks separated from Gondwana and accreted to Eurasia during the closure of the Paleo-Tethys Ocean in the north, and the Neo-Tethys Ocean branches in the south, north, and east, all within its presentday boundaries. Its present high elevation, active deformation, seismicity, and mineral resources, with complex interactions of active thrusts and strike-slip faults, are caused by the driving

1

Note that the Persian and Arabic names and words in this paper are written as correctly pronounced and written originally, with direct and simplified transliteration into English. Diacritical marks and special characters are used to differentiate vowel “a” (short; e.g., ant) from “ā” (long, e.g., Ārmenian), and Arabic “ain” (used also in Persian) as “‘a” (e.g., ‘Ābbās). Iran (pronounced “Irān”) and Tehran (“Tehrān”) are exempted from this rule due to familiarity of the readers with their correct pronunciation. The recognition of the Persian possessive (afzudeh; or ezāfé in Arabic), which inaccurately appears variously in English, especially as “-i” (thus: Shahr-i Kurd [sic]), is correctly shown as “-e” (cf. French “é”; thus: Shahr-e Kord), as conforms to the correct current usage in the Persian language (Fārsi). Locality and ethnic names, which have generally been misspelled in most foreign writings and maps, such as Isfahan, Kirman, Hormuz, Bushire, Masjid Soleiman, Baluchistan, Kurd and Kurdistan, Turk, Golistan, etc., are correctly transliterated in this work as: Esfahān, Kermān, Hormoz, Bushehr, Masjed Solaymān, Baluchestān, Kord and Kordestān, Tork, and Golestān. The name of the country since the dawn of the Iranian civilization has been “Irān” (“Airin”, “Irān-vaij”); therefore, the usage of “Persia” is restricted here, unless in quoting or in relation to the language, as in the “Persian language” (Pārsi, Fārsi). Because of their important historical value, some of the invaluable and irreplaceable unique photographs with lower resolution, scanned from old print photographs (especially Figs. 8, 10, 11, 12, 14, and 16), were used in this study. My hope is that this publication will initiate a systematic search for finding and cataloging the historic photographs of the geologists with better resolution.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084715/spe525-03.pdf by University of Wisconsin Madison user

convergence forces of the plates. The Iranian Plateau is characterized by different tectonic units and metallogenic provinces with inherited structures, some bordered by Paleozoic and Mesozoic ophiolitic sutures, organized in diverse directions that have undergone a long and complicated plate-tectonic evolution since the Late Neoproterozoic Era (Fig. 1). The Iranian Plateau consists of a composite system of collision-oblique, transpressive fold-and-thrust mountain belts with active reverse and strike-slip faulting, range-and-basin terrains, an active subduction zone, recent volcanic activities, variable crustal thickness and rigidity, and relatively stable aseismic blocks of different dimensions with low topographic relief, and nearly flat areas, generally covered by deserts (Stöcklin, 1968a, 1974a, 1974b, 1977; Takin, 1972; Berberian and King, 1981; Berberian and Berberian, 1981; Berberian et al., 1982; Şengör, 1984, 1990; Berberian, 1983a, 1983b, 1983c, 1984, 1989, 2014). The Iranian Plateau is also home to the long-lived Iranian civilization, which has contributed to the scientific and cultural heritage of humanity (Sarton, 1927; Durant, 1942; Gershevitch, 1985; Yarshater, 1983a, 1983b). I have always been interested in the ancient civilizations of Iran, Armenia, Mesopotamia, India, and Egypt, and the root causes of their rises and declines throughout the history. This has resulted in comprehensive long-term research in the ancient Avestan, Pahlavi, and Sanskrit texts concurrent with my earth science research activities (Berberian, 1997). In the exposition that follows, I try to explain, as completely as possible, the little-known, untold, and unacknowledged history of geological perceptions and explorations and their development in Iran from ca. 1200 BCE to 1990 CE within the socioeconomic and political settings of the country. In the current work, I greatly expand upon my previous attempt focused on earth science activities in Iran (Berberian, 1997), which has not yet received sufficient consideration even from Iranians scholars. Documenting the historical dimensions of the previous geological observations, thinking, and studies is one of the valuable

Development of geological perceptions and explorations on the Iranian Plateau

27

Figure 1. Simplified tectonic units of the Iranian Plateau with the main subdivisions and geosutures, modified from Berberian (2014, and reference cited therein). A—Ābādeh belt; AS—Arasbārān (Qaradāgh range) ophiolite belt, NW Iran; E—Middle Eocene Chāpedoni complex; EOF—Eocene–Oligocene flysch (Makrān); Hz—High Zāgros; JZB—Jāz Muriān Depression; K—Karkas belt; MM—Miocene molasses (Makrān); OMF—Oligocene–Miocene flysch (Makrān); NP—Neoproterozoic; P—Paleo-Tethys nappe; S—Sirjān belt; U—Lake Urumieh.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084715/spe525-03.pdf by University of Wisconsin Madison user

28

M. Berberian

tools for understanding evolution, prosperity, shrinkage, and stagnation of the Iranian civilization influenced by socioeconomic and political conflicts since at least 330 BCE. This indepth review also sheds light on the historical development of earth science in this region of the world. I have relied on first-hand original documents in different ancient and modern languages, and I have greatly benefited from correspondence and discussions with a number of foreign scientists who have worked in Iran (most of whom now have passed away). It is shown that, unlike the historical periods of ca. 1200 BCE, the Achaemenid (550–330 BCE) and Sassanid (224–642 CE) Dynasties, and the tenth- and eleventh-century “intermezzo intellectual zenith,” when the Iranians were pioneers in the earth sciences, the majority of geological investigations and ideas since the nineteenth century have been carried out mostly by people of European descent. After centuries of hiatus, Iranian geologists were gradually trained in the 1950s, and their activities have increased since the 1970s. Mention should also be made here of some works on the history of minerals and mining (Zavosh, 1969, 1976; Alipur, 1993) and petroleum (Sorkhabi, 2005) in Iran, though I have not utilized them in this study. The untold history of the pioneering, dedicated, and successful geologists who worked on the Iranian Plateau, and their contributions to the knowledge of Iranian geology and mineral resources, and earth science at large has not yet been systematically collected, reviewed, and published. The steps taken here are to: (1) appreciate the unacknowledged and widely forgotten, earlier pioneering foreign and Iranian scholars who with greatest fascination worked under harsh conditions in remote areas, and (2) to express the hope that further studies by young Iranian earth scientists will complete the elementary steps undertaken by the pioneers, who on foot, camel, donkey, and horseback, with no access roads in remote desert environments, lived on dates, camel milk, and boiled eggs, or potato and boiled water, and who searched for geologic data when the countryside was not safe and the risks were very high. 2. HISTORICAL ACCOUNTS ON THE GEOLOGICAL KNOWLEDGE OF IRAN The East, especially the Iranian Plateau, including Asia Minor (Anatolia and Greater Armenia), Mesopotamian lowlands, the Indian subcontinent, and Egypt, which was the cradle of the oldest civilizations, formed the background and basis for later Greek and Roman, as well as the early Western medicine, science, architecture, culture, art, and literature (Sarton, 1927; Durant, 1942, 1950). Despite this ancient glorious heritage, and unlike the industrialized developed countries, with the British Geological Survey (BGS) founded in 1835, the Geological Survey of Canada (GSC) in 1842, and the U.S. Geological Survey (USGS) in 1879, geological investigation on the Iranian Plateau is in its very early stages. Unexpectedly, the Geological Survey of Iran (GSI) was only established in 1962 by the United Nations

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084715/spe525-03.pdf by University of Wisconsin Madison user

Development Programme and largely ran by its experts during the 1960s and 1970s (Berberian, 1997). Early modern geologic investigations during the middle and late nineteenth century by foreign pioneers and travelers were mostly based on spot observation and reconnaissance of rocks, fossils, and mineral resources along the main access dirt roads under rather unsafe circumstances. These sporadic and nonsystematic inspections were followed in the early twentieth century by some trained European geologists. As G. Ladame (1945; quoted in Dayton, tr., 1971, p. 51), who spent three years in Iran from 1939 on and tried to set up a geological department, remarked: “I discovered soon after my arrival in Tehran that information on the Iranian geology was very fragmentary.” Despite this very late start, ancient Iranian Zoroastrian sacred books and chronicles reveal that earliest Iranian scholars, during the period that Sarton (1927, chapter II) coined as the “Dawn of Iranian Knowledge,” had developed interesting views about geology and the planet Earth that were much more advanced than those of the Europeans during the same time frame. In the following sections, I review the Iranian geological perceptions based on critical observations of landforms, surface deformations, and mountain belts focusing specifically on: (1) the ancient proto-Zoroastrian and Zoroastrian cosmogony of ca. 1200 BCE; (2) the tenth–eleventh-century CE “intermezzo intellectual zenith” perceptions with brilliant ideas; (3) the late nineteenth century to 1950; and (4) the second half of the twentieth century, which were separated by interludes of inactivity. In this process, I attempt to review the history of the development of opinions by Iranian scholars concerning the origin of mountains, the process of mountain building during earthquakes, sediments, mineral resources, and other physical geological phenomena. 3. ANCIENT IRANIAN PROTO-ZOROASTRIAN AND ZOROASTRIAN GEOLOGICAL PERCEPTIONS (CA. 1200–550 BCE) The following is a brief review of the ancient cosmogony and geological perceptions recorded in the extant sacred Zoroastrian chronicles since ca. 1200 BCE2 (for dating of the period in question, see Boyce, 1989). The sacred Zoroastrian Avestan and the later Pahlavi texts (which are not considered here for their

2

The exact lifetime of Zoroaster (Zarthosht), the Iranian teacher and educator of the highest moral values and the author of the Hymns of Gāthās, is not known, and the proposed dates range from ca. 1750 BCE to 258 yr before Alexander III of Macedon, i.e., ca. 588 BCE (Shaked, 2005; Kellens, 2006; Skjærvø, 1997, 2003; Yarshater, 2004; Malandra, 2009). On the basis of the Pahlavi text of Bundahishn [Bondahesh in Persian], some scholars (Henning, 1951; Gnoli, 1980, 1995; Gershevitch, 1995) hold to the unconstrained and doubtful traditional birth date of Zoroaster ca. 258 yr before the invasion of Alexander, whereas others, on the basis of the extant Avestā, especially the hymns of Zoroaster in the Gāthās in comparison with the oldest hymns of the Rigvedā (ca. 1700 BCE), place Zoroaster around the middle of the second millennium BCE, ca. 1200 BCE, and interpret that Zoroaster could not have lived later than 1000 BCE (Boyce, 1989).

Development of geological perceptions and explorations on the Iranian Plateau religious significance) are rich in the portrayal of human beings in relation to the forces of nature during the time that nature was sacred and considered the arena in which the divine became visible, and people lived in reasonable harmony with nature (Berberian, 1997, 2014). The ancient Iranians, like contemporary Indians on the Indian subcontinent, believed that planet Earth was split into seven large regions (keshvar) when rain first fell upon Earth. We read in the Iranian (Greater) Bundahishn, VIII (“As Regards the Whereabouts of the Lands” [ed. Anklesaria, 1956; ed. Bahar, 1990; Berberian, 1997; Peterson, 2002)] that:

29

Earth was also held by the Iranians and Indians ca. 1500 (Vedās)–1200 BCE (Avestā; Boyce, 1989), i.e., much earlier than the sixth century BCE of Greece (Berberian, 1994, 1997, 2014). According to the Iranian cosmogony described in the sacred books such as the Greater (Iranian) Bundahishn (ed. Anklesaria, 1956; ed. Bahar, 1990), “Earth” is thought of as “round” and “originally with no mountains and valleys,” lying on the cosmic waters as a yolk floats in an egg, encased by the sky as in a bag, originally thought to be made of stone or crystal, and later of shining metal (Kreyenbroek, 1993; Berberian, 1997; Hinnells, 1975). 3.2. Interior Structure of Earth (Ca. 1200 BCE)

One says in the Scripture, “[There are] thirty-three kinds of land.” [As] that which I [have written above in the matter of the land]: when Sirius produced [the] rain wherefrom the seas arose, [the land,] having seized the damp everywhere, [broke] into seven pieces,—[the lower premises] became [the upper premises, the crown became the bottom;]—one piece, as much as one half, is in the middle, and six pieces are around it; and these six pieces are as much as Xhwaniratha [Khwāniras]; the name “keshwar” was applied to them, that is, they had circumference. (ZandAkasih, Iranian or Greater Bundahishn, ch. VIII, §1, ed. Behramgur Tehmuras Anklesaria, 1956)

I would like to note that some of the proto-Zoroastrian and Zoroastrian issues addressed herein have a remarkably modern earth science attitude. 3.1. Exterior Shape and Structure of Planet Earth (Ca. 1200 BCE) The ancient world viewed by the Sumerians and Babylonians in the Mesopotamian region (the Zāgros foreland) is presumed to have consisted of a “dish-shaped flat Earth” surrounded by a moat of sea (ocean), beyond which the inverted bowl of the sky came down all around; Earth was also considered to be hollow, providing space for their underworld. The Egyptians thought Earth was a square with mountains at the edge supporting the vault of sky. Later, the dish-like Earth concept was found in the Jewish biblical accounts. It had similarities to the sort of world portrayed later in the Hellenistic Iliad of Homer and Hesiod’s poetry in classic Greece (Aaboe, 1958; Seely, 1991, 1997; Kreyenbroek, 1993; Berberian, 1997; Simanek, 2006). We know that Pythagoras (sixth century BCE) and Aristotle (fourth century BCE) argued for a spherical Earth and geocentrism. However, we read in the ancient Iranian Zoroastrian sacred books such as the Avestā (Yashts [Hymns of Praise]: Mehr Yasht: 22/85, 23/89, 24/95, 24/98; Farvardin Yasht: 1/2; Art/Ard/Ashi Yasht: 2/19; Poure Davoud, 1977a, 1977b; Kanga, 2001); Bundahishn (Bondéhéshn; West, 1880a; ed. Anklesaria, 1956; Bahar, 1990), Minavi Kherad (Mainog-i Khirad; West, 1885), as well as in the Indian Rigvedā (1/115, 1/164; Griffith, 1896), that Earth has a “spherical” form. This clearly shows that the view of the spherical shape of planet

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084715/spe525-03.pdf by University of Wisconsin Madison user

Unlike the Sumerian and Babylonian concept of a “dishshaped flat hollow Earth,” the ancient indigenous Iranian and Indian cosmogony of ca. 1200 BCE (Avestā—Yasnā: 11/7; Farvardin Yasht: 1/2; Poure Davoud, 1977a, 1977b; Kanga, 2001; Vedās—Griffith, 1896; and Bondéhéshn/Bundahishn— West, 1880a; Anklesaria, 1956; Bahar, 1990) interpreted the round Earth to be “layered” and composed of “three layers.” However, no specific description of these layers is given. We read in Mandala 10 of Rigvedā that “Agni [fire, fire Deva] is housed in the center of the Earth” (Griffith, 1896), and the Iranian Bundahishn (III/27) specifies that “Hell is in the Middle of the Earth” (ed. Anklesaria, 1956; ed. Bahar, 1990; Kreyenbroek, 1993; Berberian, 1994, 1997, 2014). These statements, reflecting that the fiery materials are stored beneath Earth’s solid surface, could have been based on observation of active volcanoes. They also show that the solid surface (crust) and the fiery hell in the middle of planet Earth constitute two distinct layers. Nonetheless, we have no idea about the nature of the third layer mentioned in the proto-Zoroastrian and Zoroastrian cosmogony. 3.3. Mountain-Building Processes during Earthquakes (Ca. 1200 BCE) According to the ancient Iranian proto-Zoroastrian myths described in the Bubdahishn (Bondéhéshn; ed. West, 1880a; ed. Anklesaria, 1956; ed. Bahar, 1990), Selections of Zād Spram (ed. West, 1897), and Zamyād Yasht (the Avestā Yashts, 19.1– 8; Darmesteter, 1898; Poure Davoud, 1977a, 1977b; Kanga, 2001), during the assault by the storm-god’s evil spirit on Earth, earthquakes and chaos materialized and ultimately led to the growth of mountain belts, which acted as ramparts and shelter against the “Evil Spirit” (Ahriman; Kreyenbroek, 1993; Berberian, 1991, 1994, 1997, 2014). Chapter VI, part C of the Iranian (Greater) Bondéhéshn (ed. Anklesaria, 1956; ed. Bahār, 1990) states that: As the Evil Spirit entered and the Earth trembled [quaked], the substance of the mountains was produced in the Earth; (on account of the shaking/ quaking of the Earth, the mountains were immediately in motion): first,

30

M. Berberian

Alburz [Alborz3] of Divine destiny [Harāiti Bārez in Zāmyād Yasht], then the other mountains within the Earth; for as Alburz [Alborz] grew up, all the mountains were in motion; for, they have all grown up from the roots of Alburz; at that time, they proceeded from the Earth, like trees, which cause the tendrils to run above and the roots underneath. Their roots were so arranged by connection, passing into one another. And thereafter, [it was not possible] for the Earth to shake from its place. [As] one, says [in the Scripture]: ‘The mountain is a great joint [anchor] of lands.” (ed. Anklesaria, 1956, §VI,C.1–4; tr. West, 1880a, §VIII.1–4, p. 29–30; Bahār, 1990, §VIII.66, p. 65)

Indeed, this is probably the earliest recorded Earth and environmental regulation in history. People were even forbidden to spit on the soil, water, fire, and air. The Zoroastrian funeral rite did not allow bodies to be buried; the corpse was laid in a special chamber on top of dedicated hills in an arena (dakhmeh), where the flesh was exposed to sun and consumed by carrion-eating animals and birds. The cleaned skeletal remains were then collected and stored in astodān (lit. bone-container; ossuary), which could not pollute or harm the soil and groundwater (The Vendidād of Avestā, 5.10–11, 8.4–9; tr. Darmesteter, 1880).

This shows that in the ca. 1200 BCE intermingled protoZoroastrian and Zoroastrian cosmogony, the indigenous Iranians believed mountains were the direct result of earthquakes initiated by the Evil Spirit assault of storm-demons on the otherwise good creation of Ahurā Mazdā (in Avestan; Ohrmazd in Pahlavi; lit. “the Wise Lord”; Berberian, 1994, 1997, 2014). This seems to be the earliest known concept of the mountain-building process in which: (1) earthquakes are incremental episodes in the mountainbuilding process (orogeny); (2) earthquakes accompany abrupt ground surface changes; and (3) earthquakes shape the large crustal structures on planet Earth. These expressions and visions of antiquity were conceived during the time when people lived a quasi-free and independent life in open nature. They indicated an early awareness of living in a seismically active region; such knowledge must have been achieved by the close surveillance of successive changes on the active Earth surface by people who observed a dynamic mechanism inside Earth. Other peoples living along seismically and volcanically active mountain belts, such as the Indians, Armenians, Greeks, Romans, Chinese, and Japanese, or the native Americans, might have held similar views. The significance of this view of mountain-building is amplified when compared to the now-outdated geosynclinals “theory of orogeny” introduced by the American geologists James Hall (1811–1895) and James Dwight Dana (1813–1895) in 1873.

3.5. Overlooked Gemstones from the Era of Mithra, Zoroaster, and Magi (Mogh), and the Ancient Iranian Knowledge of Mineral Exploration

3.4. The First Earth and Environmental Technical Regulation In the Zoroastrian doctrine, planet Earth and its four known primordial elements (soil, water, fire, and air), being good creation of Ahurā Mazdā, were all “sacred” and precious; homage was rendered to each and every angel to whose care these were entrusted, and thanks were offered to the Court of the Great Creator. Consequently, it was believed that Earth and its four primordial elements should not be contaminated by human action. 3

Alborz (Harāborzeiti, lit. “High Mountain”): It should be emphasized that ancient Iranians believed that the mighty Alborz Mountain belt made a ring around the ancient known world (Asia and Europe). Hence, they were referring to the Himalayan-Alpine mountain belt and not to the present Alborz Mountains of northern Iran and Caucasus (Berberian, 1994, 1997, 2014).

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084715/spe525-03.pdf by University of Wisconsin Madison user

The richness of the Iranian Plateau, with numerous mineral resources in different structural and metallogenic provinces, together with an ancient civilization brought some basic knowledge of geology, mineral exploration, excavation, metallurgy, and quarrying during ancient times. We know that the prehistoric ore mining and metallurgy started in the mountains of the Iranian Plateau and its western extension, the Armenian Highlands (Witter, 1938; Forbes, 1966, 1971; Wertime, 1964, 1968, 1973; Moorey, 1994; Berberian, 1997). Ghirshman (1938–1939, 1951) reported red ocher (hematite) spread on the bodies of Sialk I (6000 BCE) people; agate and turquoise found in Sialk II (5000–4300 BCE); sea shell, agate, turquoise, rock crystal, lapis lazuli, and jade in Sialk III (4100– 3400 BCE); and marble, lapis lazuli, sea shell, agate, and rock crystal in Sialk IV (3400–3100 BCE) periods in Central Iran. Turquoise is found from the fourth millennium BCE at Tapeh Hesār (Hisar [sic] in literature; Schmidt, 1937). During the third millennium BCE, lead was being smelted in the “Craftsmen’s Quarter” on the South Hill of Hesār (Dyson and Howard, 1989). In the Enmerkar (king of Uruk)–Lugalbanda Sumerian narrative poems (ca. twentyfirst century BCE), the people of Aratta (possibly Kermān region/ Shahdād [Hakemi, 1972; Hakemi and Sajjadi, 1989; Majidzadeh, 1976; Hansman, 1978] or northern Mesopotamia, Ārārāt (Māsis) area [Kramer, 1972; Kavoukjian, 1987]) are noted for quarrying and processing of stones and metals from their mountains with rich resources of gold, silver, copper, tin, and that they cut pure lapis lazuli from the lumps (Wilcke, 1969; Cohen, 1973). Although almost all the ancient Iranian texts were lost during numerous foreign invasions and occupations, Pliny the Elder (ca. 79 CE), who was contemporary with the Iranian Parthian Dynasty (312 BCE–224 CE), attributed some gemstones to Zoroaster, Magi (the Zoroastrian priests, Mogh, Moghān), and Mithra (the proto-Zoroaster god of dawn, later EOS in classic Greece, and the Zoroastrian angelic divinity, Yazata), and mentioned their locations and characteristics. The traces and legends associated with these gemstones are not preserved among the modern Zoroastrians and the Iranians at large. In classic Greece, the Magi (the Zoroastrian priests) were associated with magic, and Pliny (ca. 79 CE, XXX.2) called the Iranian Zoroaster “the inventor

Development of geological perceptions and explorations on the Iranian Plateau of magic” (we know that medicine was at first a function of the Zoroastrian priests [Magi]; Durant, 1942). Addressing Zoroaster (Zoroastres in Pliny), Magi, and Mithra probably indicates that Pliny was citing Zoroaster of the Avestā, the Teacher. Furthermore, there is no other famous Zoroaster in the chronicles, unless there are some pseudepigraphic names used by the classic Greek or Roman authors, which is not clear to me. Pliny (ca. 79 CE; Bostock and Riley, eds., 1855, XXXVII) recorded the following gemstones and the legends associated with them from the time of Mithra (proto-Zoroaster Iran), Zoroaster (ca. 1200 BCE), and the Zoroastrian priests, Magi in Greater Iran. The gemstones, and the attributes given to them as recounted in the writings of Pliny (ca. 79 CE), with some citations from the Greek scholars prior to him, occurred in Iran but have since been forgotten in modern Iran. Achates (agate): According to the Magi, there is an achates (agate) of one single color that makes athletes invincible. The method of testing such a stone is to throw it into a pot full of oil with various pigments; when it has been heated for no more than 2 h, it should have reduced all the pigments to a single shade of vermilion (Pliny, ca. 79 CE, XXXVII.54). This legend has been completely forgotten in modern Iran. Astriotes (star-like stone): They say Zoroaster sang its wondrous praises as an adjunct of the magic art (Pliny, ca. 79 CE, XXXVII.49). I was unable to find a phrase referring to this stone in the hymns of Zoroaster in the Avestā. The Avestan word for rock (as well as sky, heavens, firmament) is “asma” (asman, asmana, asana, asan4; Bahrami, 1990). There is an Avestan hard rock called “ishkata,” but its modern equivalent is unknown. Iron, lead, tin, copper, gold, silver, and brass have been mentioned in the Avestā. Atizoë (long-lived): On Mount Acidane in Persia [Iran], there is a stone found that is known as “atizoë” of a silver luster, three fingers in length, like a lentil in shape, possessed of a pleasant smell, and considered necessary by the Magi at the consecration of a king…. Other differences among achates (agate) are found in writings of Magi (Pliny, ca. 79 CE, XXXVII.54). I have not been able to trace “Mount Acidane” in the present-day Iranian orology. Bostrychitis (hair-stone, possibly iron alum, alum de plume, alunogen): Zoroaster says that it is a stone that is more like the hair of female than anything else (Pliny, ca. 79 CE, XXXVII.55). Daphnea (laurel stone): Mentioned by Zoroaster as curative of epilepsy (Pliny, ca. 79 CE, XXXVII.57). Eumithres (Eumitres; blessing of Mithra, a green tourmaline): Possibly the name is taken from Mithres (Mithra), the god 4

In proto-Zoroastrian and Zoroastrian cosmogony, traces of which are found in the Avestā, the heaven (sky) was hard and made of stone (Yasht 30:5; Poure Davoud, 1977a, 1977b; Kanga, 2001), and still in modern Persian, the sky is called “āsémān.” One of the spurs of the holy Alborz Mountain, Mount Osindām, was made of bloodstone. The conception of the heaven-stone found its perpetuation in the Pahlavi (Middle Persian) cosmogony (e.g., Dādestān-e Dinik; West, 1882), where the heaven was described as hard stone.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084715/spe525-03.pdf by University of Wisconsin Madison user

31

of the sun among the Persians (Pliny, ca. 79 CE, XXXVII.58). In the proto-Zoroaster era, Mithra was the god of dawn, later EOS in ancient Greece, and was never god of the sun as mentioned by Pliny. During Zoroastrian time, Mithra became an angelic divinity of covenant and oath, called Yazata. Exebenus: Zoroaster tells us, exebenus is a white, handsome stone, employed by goldsmiths for polishing gold (Pliny, ca. 79 CE, XXXVII.58). Mithrax (probably an opal): The stone comes from Persia (Pliny, ca. 79 CE, XXXVII.63). The name seems to be derived from Mithra (Mithres in Pliny). 4. THE ACHAEMENID PERIOD (550–330 BCE; PAX IRANICA) Our knowledge of science in general and earth science in particular during the prosperous socioeconomic period of the Achaemenid Dynasty is very limited. Medicine, which was at first a function of the Zoroastrian priests (Magi, Mogh), developed along with the growing power, wealth, culture, and peace in Iran. For example, during the reign of Artaxerxes II (404– 358 BCE) there was a well-organized guild of practicing physicians, surgeons, and pharmacists in Iran (Durant, 1942). The Foundation Charter of the Palace of Dārius I at Susā, DSf, DSz, and DSaa, gives some geological information during the early part of his reign (522–486 BCE).

DSf 10, 31–35: “And the gold, which was worked here, was brought from Sardis5 and Bactria6. And the precious stones [Kāsaka], such as the lapis lazuli [Kapautaka] and the carnelian [Sinkabrush], which were worked here, were brought from Sogdia7. And the turquoise [Akhshaina], which was worked here, was brought from Chorasmia.8” (Vallat, 2013)

Surprisingly, there is no mention of the famous ancient Neishābur turquoise mine of Khorāsān in the DSf. From DSf 11, 39–42: “And the stone columns which were carved here, were brought from a village called Hapiratush in Elam” (50 km north of Susā in Ilām province; Vallat, 2013).

5

Sardis: capital of the ancient kingdom of Lydia in western Asia Minor.

6 Bactria, Bakhtrish, Bākhtar: the plain and valley of the Āmu Daryā (Oxus) River. 7

Sogdia, Sugudā, Soghd: covering the ancient Iranian cities of Samarkand, Bukhārā, Khurjand, Panjikent, and Shahr-e Sabz in modern Tajikistan and Uzbekistan. 8

Chorasmia: Satrapy of Uvārazmiya, later Khārazm, in the Khiva (Khiveh) plain along the Āmu Daryā (Oxus) River, south of the Ārāl (Khārazm) Sea in ancient Iran.

32

M. Berberian

In DSaa 3, 11–17, Darius just mentioned that gold, silver, lapis lazuli, turquoise, carnelian, cedar wood, wood from Maka9, ebony, ivory, and relief decorations were used in the palace, and added that all the columns are of stone (Vallat, 2013). Discovery of intentional deposits of mercury in a liquid state in sun-dried mud-bricks of the lower parts of the walls (over the height of eight courses down to the foundation) at the Shāhur Palace (ca. 404–359 BCE) at Susā (Boucharlat, 2013) is very interesting. It indicates that during this period, pyrometallurgical extraction of mercury from its ore deposit of cinnabar (HgS), possibly from the Zaréhshurān mine in Takāb, northwest Iran, was well practiced. Strabo (63 BCE–24 CE), in The Geography, wrote that:

15.2.1: “Onesicritus [ca. 360–290 BCE, who accompanied Alexander’s invasions] speaks of a river in Carmania [Kermān province] that brings down gold-dust; and he says that there are also mines of silver10 and copper and ruddle, and also that there are two mountains, one consisting of arsenic and the other of salt.”

He also mentioned that:

15.2.10: “The Drangae [Drangiana, the ancient greater Sistān province of SE Iran], who otherwise are imitators of the Persians in their mode of life, have only scanty supplies of wine, but they have tin in their country.”

Pliny the Elder (ca. 79 CE; Bostock and Riley, eds., 1855, XXXVII) recovered some forgotten gemstones, minerals, and

9

Maka: The Achaemenid Empire satrapy along the northern coastal Gulf of Oman and Persian Gulf; the present-day Iranian and Pakistani Makrān. 10

The location of Arciotis (or Archeditis; a place associated with silver in ancient Greek; Potts, 1989), located ~180 km east of Sirjān on the road to Jiroft on the northern side of the Halilrud, seems to correspond with the Jebel al-Fiāa (al-faza; Kuh-e Noghreh; Silver Mountain) mentioned by ebn Huqal (978, ch. “Of the Province of Kerman”). Arciotis is located on Tabula Peutingeriana (the Poutinger Map), the illustrated road map in the Roman Empire based on a map prepared by Marcus Vipsanius Agrippa (63–12 BCE) during the reign of Emperor Augustus (27 BCE–14 CE; Dilke, 1985). 11

Carmania or Karmania was a satrapy/province of the Achaemenid and Sassanid Empires of southeast Iran (modern-day Kermān; Potts, 1989), mentioned by Pliny the Elder (ca. 79 CE) and Arrian of Nicomedia (second century CE, before 145 CE). Media (Māda, Mād), an Achaemenid satrapy/province in northwest Iran, was the political and cultural homeland of the ancient Iranian people, the Medes, ca. 615 BCE. It was located between Parthia (Pārt, in the E), Hyrcania (modern Gorgān, N), Elam (Ilām; S), and Babylon (Bābel, W) and covered modern-day Azarbāijan, Caucasus, Kurdistān ([sic], Kordestān), and southern Dāghestān. Bactria (Bākhtar, “east” in Persian) was the Achaemenid satrapy covering the plain and valley of the Āmu-Daryā (Oxus) River in modern-day northern Afghanistan and Tajikistan. Margiana (Margush) was the region in the Achaemenid satrapy of Bactria in the valley of the Murghāb River (Morghāb, lit., waterbird).

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084715/spe525-03.pdf by University of Wisconsin Madison user

mine names that were known from the Achaemenid (550–339 BCE) and Parthian (250 BCE–224 CE) Dynasties, Iran, especially from the ancient provinces of Carmania, Media, Bactria, Scythia, and Margiana.11 He also cited Democritus (ca. 460–370 BCE) regarding “Zathene,” which was purportedly native to Media, Iran: “Zathene—according to Democritus is a native of Media. It is like amber in color, and, if beaten up with palm-wine and saffron, it will become soft like wax, yielding a very fragrant smell” (Pliny, XXXVII.70). Pliny also cited the finest quality of the Murrhine vessels (Fluorine; Ganesa, Gansar, Kowari, in Persian and Arabic) imported from Carmania of the Parthian Empire, and the “Gassinade” gemstone, which was sent by the people of Media (Pliny, ca. 79 CE, XXXVII.8, ed. Bostock and Riley, 1855). Pliny also mentioned “smaragdus” of Bactriana (XXXVII.16, 17; the emerald of the Bactria Satrapy of the Achaemenid Empire), gassinade (XXXVII.59), and “iritis” (rainbow quartz), which was native to Persia (XXXVII.52). Pliny cited that Democritus (460–370 BCE) included in the “smaragdus” class the gemstones that were known as “herminei,” and as “Persian stones.” He also stated that the gemstone “tanos,” a variety of smaragdi, came from Persia (XXXVII.19). The modernday name of “emerald” is derived from an ancient Iranian word “zomorod,” translated into Latin as “smaragdus” mentioned by Pliny, which was later corrupted to “emerald.” Pliny likewise added that the smaragdi of Attica (XXXVII.18; possibly diallage, known also by the names of bronzite, schillerspath, schillerstein, and omphacite), gassinade (XXXVII.59), sapphiros (XXXVII.39; lapis lazuli, lāzhvard in Persian), and zathene (XXXVII.70; citing Democrites) came from “Media (Māda)” in Iran. Zmillampis, he wrote, was found in the Euphrates River (XXXVII.70), and the best kind of cyanos came from Scythia (XXXVII.38). According to Pliny esteemed kinds of alabastrites (XXXVI.12; calcareous alabaster, Rokhām in Arabic); asteria (XXXVII.47); astrion (XXXVII.48; star-stone; asteriated sapphire or corundum); ceraunia (XXXVII.51; aërolites or meteorites); minium (XXXVII.40 [citing Juba]; cinnabar, Shangarf in Persian); onyx (XXXVI.12 [citing Sudines]; Khalang in Persian); and callaina (XXXVII.33; turquoise or oriental peridotite) came from “Carmania” (ancient Kermān Province of southeast Iran). It should be noted that since no cinnabar has been discovered in the Kermān Province, it is probable that Juba’s reference to minium of Kermān is Soranj (Āzargun). In ancient Greece and Rome, minium (Greek) or minio (Latin) was used for both cinnabar (mercury sulfide) and minium (lead tetroxide). The modern-day Iranians have forgotten that their ancestors assigned weather-controlling power to some gemstones such as “agate” (“achates” in Pliny, XXXVII.54). According to Pliny (ca. 79 BCE), “In Persia [Persis in Pliny, Iran] they say, these stones are used, by way of fumigation, for arresting tempests and hurricanes, and for stopping the course of rivers, the proof of their efficacy being their turning the water cold, if thrown into a boiling cauldron” (XXXVII.54). Pliny wrote that “Nipparene” gemstone

Development of geological perceptions and explorations on the Iranian Plateau bears the name of a city and people of Persia, and resembles the teeth of the hippopotamus (XXXVII.64). I have not been able to find the location of the city in modern-day Iran and its neighbors on the Iranian Plateau. Pliny also wrote that they say the body of King Darius (I Achaemenid, d. 486 BCE) was buried in “chernites” (a variety of calcareous tufa?), a stone resembling ivory in appearance that preserves bodies without consuming them (XXXVI.28). As said by Pliny, gemstone choaspitis (XXXVII.56) is named after the “Kamasp” River (Gāmāsiāb, lit. “river with big fish”) in western Iran and added that a good quality of iron is the Parthian iron (XXXIV.40).12 The gray, black (the Albian–Turonian Sarvak Formation of the Zāgros), and white massive limestones used in the stone structural elements of the Achaemenid royal palaces were excavated from Mount Merci, Gondāshlu, Majdābād (in Marvdasht), Sivand (near Persepolis; Pārsé), Puzeh Palangi (Dashtestān), north of Susā (Shush), and other quarries13 (Schmidt, 1970; Stronach, 1978; Hunt, 2008), where evidence of Achaemenid quarrying is still preserved. Furthermore, the structural elements and statues required hand polishing or rubbing stones and powders that were discovered at Persepolis, some with adhering pulverized material or powder (Schmidt, 1953; Roaf, 1983; Hunt, 2008). Pliny the Elder (23–79 CE) in Naturalis Historia, Book XXXVI.10(7), wrote that: “For polishing marble statues, as also for cutting and giving a polish to precious stones, the preference was long given to the stone of Naxos [a city in Crete] such being the name of a kind of touchstone [Cotes] that is found in the Isle of Cyprus. More recently, however, the stones imported from Armenia for this purpose have displaced those of Naxos.” The extremely hard emery of 12 Amethystos: Oriental amethyst, violet sapphire, or violet corundum, and not the quartz amethyst; it is probable that Pliny includes them all, as well as violet fluorspar, and some other purple stones; inclusive, possibly, of garnet (Pliny, XXXVII.40). Amethystos, armenium (XXXV.28), sardonyx (XXXVII.58), and stone of Armenia (for polishing marble statues) come from Armenia (replacing naxos; XXXVI.10[7]). Eumeces: “of fair length”; possibly a variety of Pyromachic silex, or gun flint, nearly allied to chalcedony. Eumeces is a stone of Bactriana (XXXVII.58). Iaspis (Yashb, in Persian): jasper of Persia is sky-blue, and a variety of it is the Caspian iaspis (XXXVII.37). Iritis: the name of another stone, similar to iris (“rainbow”; Hyalin quartz iridized internally, or prismatic crystals of limpid quartz, which decompose the rays of the sun), native to Persia (XXXVII.52). Mithrax: comes from Persia (XXXVII.63). Sarda: Carnelian, a variety of chalcedony. This mineral, however, is said to be now extinct in Persia (XXXVII.31). Smaragdus (Zomorod in Persian): emerald. The Cythian smaragdus is superior to the other varieties. Next in esteem to this, as also in locality, is the smaragdus of Bactriana. Among the smaragdi, the precious stone known as tanos (euclase, or a green epidote) is also included. It comes from Persia (XXXVII.16, 17). Thelycardios: “Female heart” (the reading is doubtful) is like a heart in color, and it is held in high esteem by the people of Persia, in which country it is found. The name given to it by them is “mule” (XXXVII.68). Zmilampis: found in the Euphrates River (Pliny the Elder, ca. 79 CE, XXXVII.70 [Bostock and Riley, eds., 1855]). 13 Sadly, the Achaemenid quarries of Gondāshlu (Persepolis) and the Puzeh Palangi (Bardak Siāh Palace, Dashtestān) were destroyed in 2006 and 2009 (CAIS-SOAS, 2006, 2009).

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084715/spe525-03.pdf by University of Wisconsin Madison user

33

Armenia, which was under the Achaemenid hegemony, seems to be the closest natural source for polishing the Persepolis, Pasargadae, and Susā royal palace stones (Hunt, 2008). Dating of two wood fragments found in the ancient Nakhlak lead mine in Central Iran yielded 1790 ± 10 yr B.P. (Parthian Dynasty) and 1190 yr B.P. (Taherid Dynasty; Hallier, 1972). The use of iron clamps set into lead for bonding stone blocks in the Achaemenid royal structures (Nylander, 1970; Schmidt, 1953, 1970; Stronach, 1978) may indicate that the Nakhlak lead mine was active during the early Achaemenid period. The Achaemenid architects paid great attention to the geotechnical aspects of the foundation subsystem and substratum of the heavy and majestic royal structures. In the DSf royal Foundation Tablet (ca. 521 BCE) from the Achaemenid winter capital city of Susā, king Dārius I the Great (r. 522–486 BCE) stated how he constructed his royal residential palace (3.8 acres) and monumental throne hall (109 × 109 m) on the Āpādānā Mound at Susā and imported products from all over his empire. The palace complex was built on an elevated man-made platform bordered by a mud-brick retaining wall and filled with packed gravel (Vallat, 2013; Perrot, 2013). Although all the archives and books were destroyed during the invasion by Alexander III of Macedon in 330 BCE, it is probable that within the two centuries of the Achaemenid dynasty, Iran did not have much time and chance to contribute to the development of science in the vast empire. In particular, the Achaemenid kings were engaged in massive wars in 549–539 BCE (Lydia, Babylon), 529 (Saka/Scythia), 525 (Egypt), 514 (Scythia, Caucasus, the Black Sea), 494 (Greece), 490 (Marathon), 484 (Babylon), 480 (Greece), 405 and 532 (Egypt), and the 330 BCE invasion by Alexander, who put an end to the dynasty (Yarshater, 2004, 2012). 5. SOCIOECONOMIC AND POLITICAL PROBLEMS FOLLOWED BY DEVASTATING FOREIGN INVASIONS AND LONG HIATUSES IN SCIENTIFIC ADVANCEMENT (330 BCE–250 CE, 643–892 CE) Iranian mythology, such as that recorded in Chapter XXXIII of the Zoroastrian Pahlavi (Middle Persian language) text of Bundahishn (West, 1880a; Anklesaria, 1956; Bahar, 1990) and Dinkard (West, 1897), refers to two dark episodes that ruthlessly plundered the ancient Iranian civilizations by invasion by (1) mythical Afrāsiāb of Turān; and (2) Azhi Dahāk (Zahāk; see also Zand-e Vahuman/Bahmqn Yasht; West, 1880b). We do not have dates for these events. However, two early historical events of foreign invasion of ancient Iran are well documented (for example, History of Tabari, 850; Frye, 1975a; Yarshater, 1983a, 1983b; Jackson and Lockhart, 1986): those of (1) Alexander III of Macedon (330 BCE); and (2) the Muslim Arabs (635–652 CE). Kings, royals, priests, and people were massacred and/ or enslaved; cities, palaces, fortresses, libraries, and the Zoroastrian sacred fire temples were destroyed; the infrastructure of the country was devastated several times; and the enormous

34

M. Berberian

treasures (from the royal cities of Susā, Persepolis [Pārsé], Pasargadae, Stakhra, Bishāpur, Rhagae/Ray, and many more) were transported to Greece or Arabia. The Zoroastrian sacred Avestan texts and the Iranian books and archives were transported to Greece by Alexander III of Macedon, whereas the Muslim Arabs burned almost all the books and archives (see also Diodorus, 60–30 BCE; Arrian of Nicomedia, second century CE; Tabari, 850; Ibn Khaldun, 1377; Omstead, 1948; Ghirshman, 1951; Zarrinkub, 1975; Badian, 1985; Boyce, 1982; Curtis, 1993). Many of the documents and archives of these dark periods thereby were destroyed and/or lost, and we have little awareness of the history of the country and events. After the collapse of the Achaemenid Empire in 330 BCE, the country was occupied by the Hellenic Seleucids (312– 174 BCE; Yarshater, 1983a), and there was no encouragement or progress in scientific, medical, astronomy, art, engineering, and cultural activities. The Arab bibliographer Ibn al-Nadim (988 CE), quoting Abu Sahl al-Faāl ibn Nowbakht (ca. 800 CE), the astrologer in the court of the ‘Abbāsid caliphs al-Mansur (754–775 CE) and Harun al-Rashid (786–809 CE) in Baghdād, in describing the plundering of Persepolis (Pārsé) by Alexander III of Macedon, wrote that the books were translated into Greek, and the Persian original copies were burned (Dodge, 1970; Tajaddod, 1987; Burnett, 2015):

Alexander ruined whatever there was in the different buildings of scientific material, whether inscribed on stone or wood, and with this demolition there were conflagrations, with scattering of the books. Such of these things, however, as were gathered in collections and libraries in the city of Iastakhr he had transcribed and translated into the Greek and Coptic tongues. Then, after he finished copying what he had need of, he burned the material written in Persian. But there was a book called Al-Kushtaj [Kashtaj] from which he took what he needed of science of the stars, as well as medicine and the natural sciences. This book and the scientific material, and treasures which he hit upon, together with scholars, he sent to the land of Egypt. (Ibn al-Nadim, 988, ed. Dodge, 1970)

Ibn Khaldun, the Arab historian of Tunis who died in 1405 CE in Cairo, Egypt, also wrote about the fate of the Iranian books during the invasion and occupation of Iran by Alexander III of Macedon:

Among the Persians, the intellectual sciences played a large and important role, since the Persian dynasties were powerful and ruled without interruption. The intellectual sciences are said to have come to the Greeks from the Persians, at the time when Alexander killed Darius [Darius III Codamanus, 330 BCE] and gained control of the Achaemenid Empire. At that time, he appropriated the books and sciences of the Persians. (Ibn Khaldun, 1377, tr. Rosenthal, F., 1958)

Our knowledge about the Iranian Parthian Dynasty (250 BCE–224 CE) is also very limited, since, regrettably,

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084715/spe525-03.pdf by University of Wisconsin Madison user

almost all their records and monuments were deliberately eradicated by the subsequent Iranian Sassanid Dynasty (Yarshater, 1983a, 1983b). During the Sassanian Empire (224–642 CE), the country reclaimed its power and prosperity. The Sassanid kings Ardashir Bābakān (224–242 CE) and Shāpur I (242– 270 CE) sent envoys to India, China, and Byzantium to collect copies of the surviving Zoroastrian and other Persian texts and brought them to Iran for translation from Avestan, Sanskrit, and Greek into the Middle Persian Pahlavi language (Dodge, 1970; Tajaddod, 1987; Burnett, 2015). The academies of Gondishāpur (founded ca. 271 CE), Ctesiphon, and Resaina were established and carried out important research in medicine, pharmacology, astronomy, mathematics, and literature. The academies were versed in the ancient Zoroastrian and Iranian culture and traditions (Elgood, 1951; Hau, 1979; Frye, 1963). There was another revival phase of learning, translation, scholarship, and knowledge under the reign of Khosrau Anushirvān Sassanid (531–579 CE). Iranian scholars from Egypt and other countries were brought back to Iran. After closure of the Athenian Academy by Emperor Justinian I (r. 527– 565), Greek scholars were welcomed by Khosrau Anushirvān (Kedar and Wiesner-Hanks, 2015). Although the Sassanian Dynasty lasted for four centuries, Iran, which unfortunately became a Zoroastrian theocracy, never reached the wealth, power, and glory levels of the Achaemenid Empire. The empire suffered from numerous wars with Rome and Byzantium (11), Armenia (2), Arabia, Yemen, Syria, Egypt, Mesopotamia, and Turkic tribes (2), and showed signs of exhaustion, decline, and aging with no initiative to govern and resolve social, religion, and political crises. The army became weak and vulnerable, and authorities and people became corrupt and passive and lost their ability to defend and fight against the external enemies. This was followed by a succession of ineffectual and passive kings who could not run and defend the country. The fatigued country, therefore, collapsed without effective resistance when the nomadic Arab invaders relentlessly devastated and occupied the Iranian Plateau (Frye, 1963, 1975a; Gershevitch, 1985; Yarshater, 1983a, 1983b, 2004, 2012; Daryaee, 2013). Ibn Khaldun (1377), who himself was an Arab Muslim historiographer and historian from north Africa (Tunis, Tunisia, and Cairo, Egypt), wrote about the burning of Iranian books during the invasion by the Muslim Arabs: However, when the Muslims conquered Persia and came upon an indescribably large number of books and scientific papers, S’ad b. Abi Waqqas [the Arab commander invading Iran in 636] wrote to ‘Umar b. al-Khattab [the second Muslim Arab Caliph, r. 634–644], asking him permission to take them [the Iranian books] and distribute them as booty among the Muslims. On that occasion, ‘Umar wrote him: “Throw them into the water. If what they contain is right guidance, Allāh has given us better guidance. If it is error, Allāh has protected us against it.” Thus the Muslims threw them into the water or into the fire, and the sciences of the Persians were lost and did not reach us. (Ibn Khaldun, 1377, Rosenthal, F., tr., 1958).

Development of geological perceptions and explorations on the Iranian Plateau During this turbulent century, when the Armenian capital city of Dvin (which became part of Iran during the Sassanid Empire) was sacked by the Muslim Arabs in 640, Ānāniā Shirākātsi (Ananias of Shirāk; 610–685 CE), the Armenian philosopher, mathematician, astronomer, geographer, and alchemist, spent eight years studying at Trebizond in the Byzantine Empire (ca. 643–651 CE). In chapter XI of his treatise On the Cause of Earthquakes, he described the theory of entrapped subterranean winds14 (for a complete review of this ancient theory, see Berberian, 2014). Ānāniā Shirākātsi also wrote a small text about precious stones (apparently the first known in Armenian), listing 33 jewels and describing the features of each (Shirākātsi, tr. Ter-Davtyan and Arevshatyan, 1962, p. 112–114; Ārkādi Kārākhānyān, 28 July 2015, personal commun.). 6. THE TENTH- AND ELEVENTH-CENTURY IRANIAN “INTERMEZZO INTELLECTUAL ZENITH” AND GEOLOGICAL PERCEPTIONS Following occupation of the country by the Muslim Arabs, pockets of resistance and fights were recorded for almost 200 yr. Later, local Iranian dynasties and leaders gradually started ruling various parts of the country (Frye, 1975a). After establishment of their regime, the Arab caliphs in the capital city of Baghdād (near Ctesiphon, the ancient imperial capital of the Parthian and Sassanian Empires in 250 BCE–642 CE), realized: (1) the backwardness of the Arabs in science, engineering, philosophy, literature, and art, on one hand, and the wealth of impressive vestiges of the Iranian, Indian, Greek, Roman, and Syriac civilizations, knowledge, and art, on the other hand; (2) the surviving treatises of these ancient civilizations; and (3) the remaining scholars at colleges working at the Gondishāpur, Alexandria, Beirut, Antioch, Harran, and Nisibis academies. Hence, they ordered the available treatises collected and translated from Pahlavi (Middle Persian), Sanskrit (Indian), Greek, Syriac, and Roman into Arabic (the official language of the time), mainly by the Iranian, Nestorian 14

Since this translation from the Old Armenian (Grabar) to the modern Armenian is not published in English, and was kindly provided by Ārkādi Kārākhānyān (29 July 2015), I insert it here as a footnote for further research in the ancient theory, which was overlooked in the historical description of the theory in Berberian (2014). Ānāniā Shirākātsi (ca. 660–681 CE): “Chapter XI. On the cause of earthquakes: Earthquake is caused by winds that come deeply into the ground, but cannot shake it and just produce roaring under the ground. And there are also (strong winds) that are also unable to shake it because of its density. But if such a wind forces its way out to the surface, it will bring death to many cities and destroy buildings. If the wind does not find its way out, it passes with terrible rumbling and roaring (underground) and shakes the ground, but not entirely: rather in some places and from time to time in different countries and in different ways—at times strongly, at times weakly.” (Ānāniā Shirākātsi, tr. TerDavtyan and Arevshatyan, 1962)

He also described the world “as being like an egg with a spherical yolk [the globe] surrounded by larger layer of white [the atmosphere] and covered with hard shell [the sky]” (Hewsen, 1968, 2001; Greenwood, 2011). The Armenian State Award of the Ānāniā Shirākātsi Medal is awarded to scientists, engineers, and inventors.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084715/spe525-03.pdf by University of Wisconsin Madison user

35

Christian, and Jewish scholars from ca. 750 to 900 CE (Sarton, 1927, 1952; Durant, 1950). In 830 CE, the ‘Abbasid Caliph al-Ma’mun (r. 813–833) established a scientific academy, an observatory, and a public library called the “House of Wisdom” (Bayt al-Hikmah), at a cost of ~200,000 dinars (US$950,000) in the capital city of Baghdād (Durant, 1950). The names of the original Persian, Indian, Greek, Roman, and Syriac books, their authors, and the translators were documented by Ibn al-Nadim (988). As Sarton (1952, p. 27) correctly stated, “The Arabic tradition was a continuation and revivification not only of Greek science but also of Iranian and Hindu ideas. This is still very imperfectly known and will require many more investigations than have hitherto been possible.” It was for this fact that Sarton (1927, p. x, xi) divided the world history of science during Europe’s Medieval period according to the names of the prominent Iranian scientists and scholars, such as: (1) “the time of Jabir ibn Hayyām (second half of the eighth century)” (Chapter XXVIII, p. 520–542); (2) “Khārazmi (first half of the ninth century)” (Chapter XXIX, p. 543–582); (3) “Rāzi (second half of ninth century)” (Chapter XXX, p. 583–618); (4) “Biruni (first half of the eleventh century)” (Chapter XXXIII, p. 693–737); and (5) “Omar Khayyām (second half of the eleventh century)” (Chapter XXXIV, p. 738–783). Sarton (1927, 1952), Adams (1938), and Durant (1950) were among few researchers who clearly stated that these scholars were Iranians who wrote in Arabic. During 821 and 1062 CE, Iran was governed by the first nonArab Iranian dynasties, including the Tāherids (821–873 CE), Sāmānids (819–999 CE), Saffārids (861–1003 CE), Ziyārids (928–1090 CE), Buyids/Dailamites (934–1062 CE), and Sālārids (Mozaffarids, 942–979 CE) in various parts of Iran, all of whom esteemed their Iranian pre-Islamic heritage. The period is known as the “Iranian Intermezzo” (Golden Ages), since it was a national Iranian interlude between the Arab ‘Abbasid caliphate rulings and the invasion of the Ghaznavid and Saljuq Turkic nomadic tribes (Boyle, 1965; Frye, 1963, 1965, 1975a; Yarshater, 1983a, 1983b). It was during this thriving period that national identity, prestige, culture, art, literature, science (including geology), engineering, and medicine rekindled and flourished in Iran. The tenth to eleventh century in Iran was a period of close studies, observations, and writings by brilliant Iranian scholars such as Abu Ma’shar Balkhi (787–886 CE), Ahmad Farghāni (Alfraganus, 798–965 CE), Musā Khārazmi (Algoritmi/ Algaurizin; al-Khwarizmi [sic]; 780–850 CE), Tabari (839–923 CE), Ahmad Sahl Balkhi (850–943 CE), Zakaryā Rāzi (Rhazes, 854–925/932 CE), Fārābi (Alpharabus, 870–950 CE), Abul Faraj Esfahāni (897–967 CE), Abdolrahmān Sufi Rāzi (Azophi, 904– 986 CE), Abulwafā Buzjāni (940–998 CE), Karaji (953–1029 CE), Avicenna (Pursinā, ebn Sinā; 980–1037 CE), Biruni (Aliboron 973– 1048 or 1051? CE), Esmā’il Esfezāri (1045–1150 CE), Ferdowsi Tusi (935–1025 CE), and many more. Nasr (1963, 1964, 1966, 1968, 1976, 1980, 1993) in his publications addressed the Iranian scientists merely as “Muslim scientists and scholars” and attributed the Iranian science and civilization as “Islamic science and civilization” influenced by Shi’ism. Furthermore, Sir Hamilton

36

M. Berberian

Gibb (Gibb, 1993) added that the progress was the result of the impact of Hellenism on the Islamic scholars. Similar attributes were stated by numerous scholars. However, an Iranian citizen, Nasr (1968, p. 111) wrote that: “Of the geological studies by Muslim authors, few are as exact and penetrating as those of al-Biruni [Biruni, an Iranian scholar and polymath], who made acute observation of land forms and mountain structures.” Nasr (1968) and Gibb (1993), as well as Giorgio De Santillana (1968), Falagas et al. (2006), Bill Scheppler (2006), and Turner (2008), among others, have labeled these Iranian scientists merely as “Muslim scientists,” thus totally ignoring their national identity and their experimental methodology and quantification contributions, as well as the ancient Iranian cultural, scientific, engineering, and art heritage and influence during this Iranian intermezzo. Many Iranian scholars were active in the newly founded capital city of the ruling Abbasid caliphates at Baghdād, which was itself built in the shadow of the devastated Parthian and Sassanid capital city of Ctesiphon. In honor of the Iranian scholars of the tenth–eleventh centuries, five craters of the Moon are named after them (discussed later). Of course, these Iranian scholars wrote the bulk of their works in Arabic, which had become the predominant official and scientific language of Muslim lands. However, unlike the Arab scholars, the Iranian scholars of the period could also speak, read, and write in the Iranian language (Pārsi/Fārsi, Persian). Avicenna’s books Dāneshnāmeh ‘Alā’i (The Book of Scientific Knowledge, written for ‘Alā’ al-Dauleh [Abu Ja’far Doshmanziyāri, known as ‘Alā’ al-Dauleh Kākuti], 1008–1041 CE), Dānesh-e Rag (On the Science of Pulse), and his poems were written in the Persian language. Biruni’s book al-Tafhim l’Sanā’āt al-Tanjim (Understanding Astrology; 1029 CE) was written in Persian. Both Avicenna and Biruni created and employed purely Persian scientific vocabulary and had extant Persian correspondences debating scientific issues. Persian writing was especially rekindled after the work of the Books of Kings (Shāhnāmeh; see Joneidi, 2008) by the renowned Iranian poet Ferdowsi Tusi (1010), who composed the pre-Islamic mythology and history of Iran in an epic poem, probably the longest of its kind in world literature. We know that the Buyids/Dailamites (934–1062 CE) traced their lineage back to the Sassanian King Bahram V Gur (r. 420–438) and were interested in the Iranian past history, art, and cultural heritage, especially their proud association with the Sassanid Dynasty. For example, ‘Ad·ud al-Dawla (Azad alDaula; Panāh Khosrau) Dailamite (‘Azad al-Dauleh Dailami; r. 967–983), a prominent scientific patron who rebuilt the Sassanid city of Ardashir-Khawrrah (the Glory of Ardashir; Gur, modern Firuzābād in Fārs, in south Iran), appeared as a Sassanid king on the coins minted in Fārs (Bosworth, 1973) and proudly recorded his visit accompanied by a Zoroastrian priest to Persepolis (Frye, 1963). Scholars such as Durant (1942, 1950), Frye (1963, 1965, 1975b), Bosworth (1973), Busse (1975), and many more have identified the Iranian heritage and Sassanid Iran as the source of the so-called “Arab or Islamic golden age.”

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084715/spe525-03.pdf by University of Wisconsin Madison user

Musā Khārazmi (780–850 CE) researched and wrote about the Hindu numerals, astronomical tables, trigonometric tables, analytical and geometrical solutions of quadratic equations, and geographical knowledge. Ibn Kathir al-Farghāni (Farghāni, Alfraganus, 798–965 CE) wrote an astronomical text ca. 833 CE. In ca. 891 CE, Ahmad Ya’qubi, who lived in Armenia and Khorāsān, wrote a Book of the Countries (Kitāb al-Buldān), giving a reliable account of states, provinces, and cities. Zakaryā Rāzi (854–925/932 CE), the Iranian polymath, physician, chemist, and philosopher, wrote some 131 books, half of which were on medicine. His two books Kitāb al-Hawi (Comprehensive Book, Liber in Latin), which in 20 volumes covered every branch of medicine then known, and Kitāb al-Mansuri (Book for alMansur; Nonus Almansor, in ten volumes) were used as medical textbooks in Europe until the sixteenth century (Sarton, 1927). Biruni (1050) mentioned that the Iranian scholar Abu Sa’d Nasr (b. Ya’qub Dinawari, [sic] Dinévari, ca. 989 CE) commenced his great mineralogical treatise in Persian with enumeration of the names and characteristics of the precious and semiprecious stones and metals. Unfortunately, the original treatise of Dinévari (ca. 989 CE) is lost, and little is known about him. Fortunately, Biruni procured the material and used it in his book by citing the forgotten Iranian scholar. Avicenna (Pursinā, ibn Sinā; 980–1035 CE), also known as “the Leader among Wisemen” (raiis al-hukamā, in Arabic), discussed geological phenomena in his philosophical and scientific encyclopedia (Avicenna, 1000, 1020). The proto-Zoroastrian and Zoroastrian cosmogony of the mountains with “connected roots to each other” and being “active and in motion,” can also be traced in the thoughts of the Iranian scholars of this period such as Biruni (1025), and the poet Maulānā Jalāl U’ddin Balkhi (Rumi [sic] in literature, 1263–1273, in Mathnavi-e Ma’navi, Book IV:9; ed. Nicholson, 1926), among many others. Amazingly, geological perceptions and scientific discoveries of Karaji (ca. 1010 CE), Avicenna (1020 CE) and Biruni (1000, 1025, 1030, 1050 CE) and others are close to the modern interpretations in earth science. These scholars utilized indepth observations and experimentations of their own research and journeys and also benefited from their heritage from ancient civilizations, including their own Iranian legacy. In their treatises, they described in detail their observations and research trying to address the probable sources and causes of natural phenomena active on Earth. Since the views of these scholars on earthquakes are addressed elsewhere (Berberian, 2014), they are not addressed in this essay. Instead, some other aspects of their works related to geology are highlighted in the following pages. 6.1. Estakhri’s Observation of Flora and Fauna Fossils (Ca. 951 CE) Estakhri (951, ed. Afshar, 1961) mentioned that, “About two farsang/leagues [~12 km] to the north of the Shur route [‘Shur-rud’, lit. salty river; between Tabas and Kuhbanān], there might be seen curious stones, in likeness of various fruits, to wit,

Development of geological perceptions and explorations on the Iranian Plateau almonds, apples, nuts, and pears, while the forms of men and trees were simulated by the rocks here, with likeliness of other created things.” Apparently, Estakhri (951) was describing fossils in the plant-bearing Jurassic shales of the Shemshak/Nāyband Formation in southeast Iran. The desert route started from the village of Birah/Bireh in the Shur district on the frontier of Kermān near Kuhbanān. From there, the passage was made in seven or eight stages (manzel), with each halt at a watering place, to Kuri, a village on the Lut Desert border of the Qohestān Province, situated a few miles to the southeast of the oasis town of Tabas-e Golshan (Le Strange, 1905). 6.2. Biruni’s Rationale for the Artesian Wells (1000 CE) Biruni (al-Biruni/Aliboron, 973–1051), the brilliant Iranian philosopher, historian, traveler, geographer, linguist, mathematician, astronomer, poet, and physicist authored numerous highly technical treatises on different subjects. His book Athār al-Bāqiya (Vestiges of the Past, ca. 1000 CE) described the calendars and religious festivals of the Iranians, Zoroastrians, Syrians, Greek, Jews, Christians, Sabeans, and Arabs. Although, he was a Moslem, he retained a degree of Persian patriotism and condemned the Muslim Arabs for destroying the high civilization of the Sasanian Empire (Sarton, 1927; Durant, 1942; www.iranicaonline.org). In his book History of India (Tārikh al-Hind), Biruni dedicated 42 chapters to Hindu astronomy and gave the best medieval account of Indian minerals. He translated several Sanskrit works of science into Arabic. He formulated astronomical tables and wrote about the astrolabe, the planisphere, and the armillary sphere of the celestial globe. Biruni noted “the attraction of all things toward the center of the Earth” and remarked that astronomic data can be explained as well by supposing that Earth turns daily on its axis and annually around the sun, as by the reverse hypothesis (Sarton, 1927; Durant, 1942; www.iranicaonline.org). Biruni (1000 CE, Dānāseresht, A., tr., 1941, p. 409) discussed the conditions of free water outpouring and flowing out of wells [artesian wells]. He wrote that:

Elevation of the “confined aquifer” containing groundwater should be higher than the elevation of the location of the well to let rising of water to the surface and its flowing out of the well; if the elevation of the aquifer is high enough the water can freely flow to the top of castles and minarets.

This clearly shows that Biruni (1000) had knowledge of geology and hydrogeology and was fully aware of the late sixteenth-century Stevin’s law in communicating vessels, the concept used in irrigation systems since ancient days, and used in ancient Rome for indoor plumbing systems (Spellman and Whiting, 2005). It should be noted that de Vries (2006) in his chapter 1.3, “Development in Subsurface Hydrology,” labeled Biruni as “the Arabian philosopher.”

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084715/spe525-03.pdf by University of Wisconsin Madison user

37

6.3. Karaji’s Geology and Hydrogeology (Ca. 1010 CE) Karaji (953–1029 CE; also al-Karaji; misrepresented as alKarkhi/al-Karaki, even in Nasr, 1968) was a genius mathematician, engineer, and hydrogeologist born in Karaj in Iran. His surviving books include: al-Badi’ fi’l-Hisāb (Wonders of Calculus), al-Fakhri fi’l-Jabr wa’l-Muqābala (Glorious Algebra, a book dedicated to Fakhr al-Din), al-Kāfi fi’l-Hisāb (Sufficient on Calculus), and his masterpiece in hydrogeology. Karaji (1010), in his treatise the Extraction of Hidden Water [groundwater] to the Surface, possibly the oldest extant book on hydrogeology (ed. Khadiv-Djam, 1966; Berberian, 1997; tr. Zaghi, 2007), showed his geological and hydrogeological knowledge obtained by closely following the geological and engineering aspects of construction of qanāts.15 In 30 chapters, his book shows that Karaji (1010) was knowledgeable of basic hydrologic, hydrogeologic, geology, and engineering principles of groundwater and its extraction systems. He wrote about classification of soil types, searching for groundwater in dry lands, different types and hydraulic characteristics of aquifers (unconfined, confined, and perched), the use of plant growth as an indicator of groundwater aquifers, and invention of ingenious surveying devices used in qanāts (ed. Khadiv-Djam, 1966; Nadji and Voight, 1972; Zaghi and Finnemore, 1973; Pazwash and Mavrigian, 1980; Zaghi, 2007). Throughout history, water in semiarid to arid Central Iran has always been a critical issue. The possible invention of the qanāt (Kāriz) system in ancient Iran provided a critical technique for water supply for irrigation of cultivable lands far from the mountains. Karaji (1010) spent his lifetime in developing the construction techniques of qanāts and the required instruments. By understanding the geological processes of mountain building and erosion, Karaji (1010) wrote that:

After mentioning this, it has to be stated that there exists continual motions in the Earth, and some of these motions cause the collapse and the destruction of buildings and the inclination of objects from their vertical position. This is why mountains and hills little by little become lower and in seeking the Earth’s center they break into parts. Also in soft soils of the Earth, the continual motion exists, pressing pieces against each other to result in hardness and rigidity. The great-

15 Qanāt (Kahriz, Kāriz): Ancient Iranian irrigation technology in semiarid and arid regions of central and east Iran (Wulff, 1966, p. 250–260; Wulff, 1968; English, 1968; Reza et al., 1971; Mansuri, 1989; Beaumont et al., 1989). It is composed of several well-like shafts (with intervals ranging from 15 to 100 m; with an average of 50–70 m in the Khorāsān Province), connected by gentle sloping tunnels several kilometers long, directing the shallow aquifer from slightly higher elevations in the alluvial deposits, especially at the footwall of active reverse faults. In some cases, the mother shaft taps the elevated water table of the tip of a reverse hanging-wall block, where the aquifer is sealed by the active fault gouge. The oldest discovered qanāt system in Iran is dated about 4000 yr B.P. in an archaeological site at Semnān in northern Central Iran (Mehryar and Kabiri, 1986). Optically stimulated luminescence (OSL) dating of the Miām qanāt system in northeast Iran yielded ca. 3.4–4.3 ka; the qanāt was maintained until at least ca. 1.6 ka (Fattahi, 2015). This makes it thus far the oldest dated qanāt system in Iran and abroad, and it disqualifies the claim by Magee (2005) who insisted that the qanāt technology came to Iran from southern Arabia.

38

M. Berberian

est of the motions mentioned, is the flow of great water bodies and rivers from one land to another over a long period of time. When the sources of such water bodies are located and accumulated in a region and the center of this accumulation does not balance the opposite side of the Earth. Continually the materials of the Earth will move in order to maintain the required equilibrium. As a result, the geographical latitude, sunset and noontime of cities are changed; seas are relocated and springs caused to appear and dry up. The occurrence of these different processes will not happen within an hour, but it will be gradually similar to the relocation of the habitation. (Karaji, ca. 1010, ed. KhadivDjam, 1966; Eng. tr. Zaghi, 2007, p. 38–39)

6.4. Avicenna on Formation of Minerals and Metals (1020 CE) Avicenna (Pursinā, Ibn Sinā; 980–1037 CE) translated Euclid’s Elements (ca. 300 BCE), made astronomical observations, and devised an instrument like a vernier. He conducted original studies of motion, force, vacuum, light, heat, and specific gravity. His two masterpieces were Kitāb al-Shifā (Shafā; Book of Remedy or Healing; Sufficientia’ [of the soul]), an 18 volume encyclopedia of mathematics, physics, metaphysics, theology, economics, politics, and music; and Qānun fi al-Tibb (Canon of Medicine), a gigantic survey of physiology, hygiene, therapy, and pharmacology, with sundry excursions into philosophy (Sarton, 1927; Durant, 1942; www.iranicaonline.org). Avicenna (ca. 1020 CE), in al-Shafā, probably the longest encyclopedia of knowledge written by a single scientist and physician, discussed the principles of geology and mountain building, mineralogy, chemistry, meteorology, astronomy, and physics based on his observations, experiments, and readings. In Part II, Section five, of this work, the “Article on Mineralogy and Meteorology,” Avicenna presented a complete coverage of knowledge on what happens on Earth in six chapters of: (I) “Formation of the Mountains”; (II) “The Advantage of Mountains in the Formation of Clouds”; (III) “Sources of Water”; (IV) “Origin of Earthquakes”; (V) “Formation of Minerals”; and (VI) “The Diversity of Earth’s Terrain.” Avicenna’s (1020) Section V, “Formation of Minerals,” was translated into Latin by Alfred of Sareshel, ca. 1200 CE, under the title of De Mineralibus (it was included with Aristotle’s work and later was printed with Aristotle’s Meteorologica and Secretum Secretorum). Avicenna classified minerals into stones (ahjār), fusible substances (zāyebāt), sulfurs (kebārit), and salts (amlāh) and discussed the subdivisions and characteristics of each (Holmyard and Mandeville, 1927; Adams, 1938; Nasr, 1964, 1968, 1993; Toulmin and Goodfield, 1965; Berberian, 1997; Otte, 1988; ed. Sa’id, 1989; al-Rawi, 2002). Avicenna’s (1020) observation that meteoric stones (meteorites) originate in the heavens and fall upon the planet Earth differs from Aristotle’s view that such stones originated on the surface of Earth and were blown up into the heavens by a violent wind, subsequently falling to Earth again (Adams, 1938). He also disagreed with Aristotle and the alchemists, in that the metals cannot be changed or transmuted into one another, especially into gold, but that they are each composed of a separate and distinct kind of earth (Adams, 1938).

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084715/spe525-03.pdf by University of Wisconsin Madison user

6.5. Avicenna on the Origin of Mountains and Mountains Rising from Ocean Beds during Earthquakes (1020 CE) Based on studies as well as field observations through his many journeys in different terrains, Avicenna (1020) described the formation of sedimentary rocks, rocks formed under severe heat, the hardening of stones, formation of mountains through erosion of softer layers of rocks, the change of landmasses into sea (ocean) and vice versa, and the origin of fossils as remnants of marine fauna of older epochs. Avicenna hypothesized two causes of mountain building (“formation of cliffs and height” in his writings), but insisted that in order to understand the process, we have to understand the formation of sediments, then rocks, and finally the process of mountain building by: (1) uplift during earthquakes, and (2) erosion (Holmyard and Mandeville, 1927; Adams, 1938; Nasr, 1964, 1968, 1993; Toulmin and Goodfield, 1965; Berberian, 1994, 1997; Otte, 1988; al-Rawi, 2002). Avicenna’s (1020) process of “mountain building during earthquakes” is rooted in the ancient proto-Zoroaster and Zoroaster perception that goes back to ca. 1200 BCE (discussed earlier and in Berberian, 2014). Avicenna first put forward his accounts regarding formation of sediments and stones/rocks (ed. Holmyard and Mandeville, 1927, tr. Adams, 1938, p. 333–334) as:

We shall begin by establishing the condition of the formation of mountains and the opinion that must be known upon this subject. (i) The first topic is the condition of the formation of stone [rock]; (ii) the second is the condition of the formation of stones great in bulk or in number; and (iii) the third is the condition of the formation of cliffs and heights [mountains]. We say that, for the most part, pure earth does not petrify, because the predominance of dryness over [i.e., in] the earth, endows it not with coherence but rather with crumbliness. In general, stone is formed in two ways only: (a) through the hardening of clay, and (b) by the congelation of waters. Stone has been formed from flowing water in two ways: (a) by congelation of water as it falls drop by drop or as a whole during its flow; and (b) by the deposition from it, in its course, of something which adheres to the surface of its bed and then petrifies. Running waters have been observed, part of which, dripping upon a certain spot, solidifies into stone or pebbles of various colors, and dripping water has been seen which, though not congealing normally, yet immediately petrifies when it falls upon stony ground near its channel. We know, therefore, that in that ground there must be a congealing petrifying virtue which converts the liquid to the solid… or it may be that the virtue is yet another, unknown to us. Stones are formed, then, either by the hardening of agglutinative clay in the sun, or by the coagulation of aquosity by a desiccative earthy quality, or by reason of a desiccation through heat. If what is said concerning the petrification of animals and plants is true, the cause of this [phenomenon] is a powerful mineralizing and petrifying virtue which arises in certain spots, or emanates suddenly from the Earth during earthquakes and subsidences, and petrifies whatever comes into contact with it. As a matter of fact, the petrification of the bodies of plants and animals is not more extraordinary than the transformation of waters. (Avicenna, 1020; tr. Adams, 1938, p. 333–334)

Development of geological perceptions and explorations on the Iranian Plateau After describing the process of sedimentation and formation of rocks and fossils from petrification of fauna and flora by mineralization, Avicenna (1020) returned to his concept of mountain building (ed. Holmyard and Mandeville, 1927, tr. Adams, 1938) and wrote:

The formation of heights [mountains] is brought about by (i) an essential cause, and (ii) an accidental cause. (i) The essential cause is concerned when, as in many violent earthquakes, the wind, which produces the earthquakes, raises a part of the ground, and a height is suddenly formed [uplift]. (ii) In the case of the accidental cause, certain parts of the ground become hollowed out while others do not, by the erosive action of winds and floods which carry away one part of the earth but not another. That part which suffers the action of the current becomes hollowed out, while that upon which the current does not flow is left as a height. The current continues to penetrate the first-formed hollow until at length it forms a deep valley, while the area from which it has turned aside is left as an eminence. This may be taken as what is definitely known about mountains and hollows and passes between them.

Avicenna (1020) associated mountain-building process with earthquakes, utilizing the ancient Iranian concept of air movement through subterranean fissures causing earthquakes (for development of the concept see Berberian, 2014, p. 77–98). Avicenna (1020) continues:

Mountains have been formed by one or other of the causes of the formation of stones, most probably from agglutinative clay which slowly dried and petrified during ages of which we have no record. It seems likely that this habitable world was in former days uninhabited and, indeed, submerged beneath the ocean. Then, becoming exposed little by little [uplifted], it petrified in the course of ages the limits of which history has not preserved; or it may have petrified beneath the waters by reason of intense heat confined under the sea [ocean]. The more probable of these two possibilities is that petrification occurred after the earth had been exposed, and that the condition of the clay, which would then be agglutinative, assisted the petrification. It is for this reason [i.e., that the Earth was once covered by the ocean/sea] that in many stones, when they are broken, are found parts of aquatic animals, such as shells, etc. It is not possible that the mineralizing virtue was generated there [i.e., in the petrifying clay] and aided the process, while the water also may have petrified. Most probably, mountains were formed by all these causes. (Avicenna, 1020; tr. Adams, 1938, p. 334)

Here it seems that Avicenna (1020) categorized rocks into two groups of sedimentary (with aquatic marine fauna) and metamorphic and or volcanic rocks. Although he mentions “intense heat,” he does not talk about pressure or force.

The abundance of stone in them is due to the abundance, in the sea, of clay which was afterwards exposed. Their elevation is due to the excavating action of floods and winds on the matter which lies between them, for if you examine the majority of mountains, you will see that the hollows between them have been caused by floods. This action, however, took place and was completed only in the course of

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084715/spe525-03.pdf by University of Wisconsin Madison user

39

many ages, so that the trace of each individual flood has not been left; only that of the most recent of them can be seen. (Avicenna, 1020; tr. Adams, 1938, p. 334)

After induration of sediments at the bottom of oceans and gradual uplift of mountains from the bottom of the ocean, Avicenna (1020) paid attention to the process of weathering and erosion acting on the uplifted mountains:

At the present time, most mountains are in the stage of decay and disintegration, for they grew and were formed only during their gradual exposure by the waters. Now, however, they are in the grip of disintegration, except those of them which God wills should increase through the petrifaction of waters upon them, or through floods which bring them a large quantity of clay that petrifies on them. It is also possible that the sea may have happened to flow little by little over the land consisting of both plain and mountain [transgression] and then have ebbed away from it [regression]. It is possible that each time the land was exposed by the ebbing of the sea a layer was left, since we see that some mountains appear to have been piled up layer by layer [stratified], and it is therefore likely that the clay from which they were formed was itself at one time arranged in layers. One layer was formed first, then, at a different period, a further layer was formed and piled (upon the first and so on). Over each layer there spread a substance of different material, which formed a partition between it and the next layer; but when petrification took place something occurred to the partition which caused it to break up and disintegrate from between the layers. As to the beginning of the sea, its clay is either sedimentary or primeval, the latter not being sedimentary. It is probable that the sedimentary clay was formed by the disintegration of the strata of mountains. Such is the formation of mountains. (Avicenna, 1020; tr. Adams, 1938, p. 334–335)

Avicenna’s astonishingly clear geological observations in 1020 CE regarding stratified sediments, transgression, regression, formation of rocks from sediments, and gradual mountain-building processes by earthquakes are remarkably close to the modern scientific views after an elapsed time of more than eight centuries. 6.6. Biruni on the Process of Mountain Rise from the Seabed (1025 CE) Biruni was one of the most learned men of his age and an outstanding intellectual figure. In his astronomical works, Biruni (1025) discussed with approval the theory of Earth’s rotation on its axis and made accurate calculations of latitude and longitudes for different cities (‘Ali, 1967; Kennedy, 1973; Sparavigna, 2013). In physics, he explained natural springs by the laws of hydrostatics and determined with a remarkable degree of accuracy the specific weight of 27 precious stones and minerals (Sai’d, 1989). In geology, he advanced the daring view that the valley of the Indus had once been at sea (‘Ali, 1967; Kennedy, 1973). Based on his observations, Biruni properly stated that the regions he investigated in India, Arabia, the desert between Jorjān

40

M. Berberian

(Arabicized Gorgān; modern Gonbad Kāvus SE of the Caspian Sea) and Khārazm, and Darband (lit. gate; modern Derbent [sic]; in the Dāghestān, NW of the Caspian Sea) were once sea (ocean) bottoms, since their surface strata are composed of geological marine deposits containing fossils (petrified sea fauna remains; he called them gush māhi in Persian, i.e., bivalve shells, or seashells) in the hard sediments now on Earth’s surface. Kennedy (1973, p. 5) literally translated the Persian word gush māhi to “fish-ears” and commented that “the ‘fish-ears’ can hardly be identified precisely.” Biruni wrote that (1025; ed. Sachau, 1910; ed. Krenkow, 1961; Nasr, 1968; ed. Ārām, 1973; ed. ‘Ali, 1967; Kennedy, 1973; Berberian, 1997):

This steppe of Arabia was at one time sea which was later unturned, and traces of that sea become evident on digging for wells and springs, because the desert then unfolds various strata of earth, sand, and soft pebbles, intermingled with pieces of pottery, glass, and bones, which could not have been buried there intentionally. Again, a variety of stones is excavated which reveals, on breaking up, definite sea products: shells, cowrie shells, and what are called “fish ears” [gush māhi in Persian; bivalve shell fossils]. These products will be found, either fully preserved, or in a state of complete decay, and in the latter case they will have left their figure completely imprinted in cavities in the stones [fossil imprints in rocks]. Such remains can also be found in [the city of] Bāb al-Abwāb [Arabic translation of Darband/Derbent in modern Dāghestān] on the coast of the Caspian Sea. The duration and dates of such transformation are completely unknown… We find the like of these stones, with “fish ears” [gush māhi in Persian; bivalve shell fossils] in their middle, in the sandy desert between Jorjān [Arabicized Persian Gorgān; the modern Gonbad Kāvus] and Khwārizm.16 (Biruni, 1025, tr. ‘Ali, 1967, §44.1–15, p. 18; Ārām, 1973, p. 20)

district of Kirmān [(sic) Kermān], the stems of date-palms which used to grow there, but the climate of the locality grew colder and the palms dried and withered, and that at his time there were no palms within a radius of twenty farsakhs [~120 km] from the castle. He also added explicitly that, when the level of that locality rose, many brooks and rivers, that had been flowing in the adjoining land, were sunk. (Biruni, 1025, tr. ‘Ali, 1967, §43.2–10, p. 17; Ārām, 1973, p. 19)

Further on, Biruni (1025) gave another example and interpretation of land sinking and sedimentation:

Abu al-‘Abbās al-Irānshahri [Neyshāburi/Neishāburi] related that a canal was dug out in the district on Busht [possibly Bāsht/Bāshtin, 29 km WSW of Sabzévār], from the borders of Neishābur city, and that the trunks of willow trees, which had been sawed with a saw, were found in the canal at a depth of over fifty Zara’ [cubits?]. It is obvious that, since the trees were cut on the surface of the Earth, the time during which that depth of earth had been compressed is too long to be assessed accurately. Also, one should not be surprised that the wood has been preserved for so long, because when it is kept away from a place which is exposed to extremes of heat and cold that alternate through the year, it is more likely to be preserved longer. (Biruni, 1025, tr. ‘Ali, 1967, §51.4–10, p. 22; Kennedy, 1973, p. 5; Ārām, 1973, p. 24)

The 50 Zara’ (Zar’) depth seems either exaggerated or usage of a different system by Neyshāburi. One Zar’ equals 16 gaz, ~1.04 m (used in Tehran and Fārs provinces); Zar’-e Shāhi (Imperial) Zar’ was ~1.12 m (used in Tabriz); and Zar’-e Neyshāburi was ~2× the length of Zar’-e Shāhi (Dehkgoda Persian Lexicon, icps .ut.ac.ir, icps.ut.ac.ir/f-index.html, Vajehyab.com, parsi.wiki).

6.7. Biruni’s Concept of Uplift and Change in River Base Level (1025 CE)

6.8. Biruni’s Concept of Continental Movement and Drift (1025 CE)

Biruni’s (1025) description of changes in the lowest point that rivers flow (base level) is also interesting because he related it to the uplift of the land (ed. Aram, 1352; ed. ‘Ali, 1967; Kennedy, 1973; Berberian, 1997):

In the introductory section to his masterpiece treatise on geodesy, Biruni (1025) conversed on related topics about causes of geological changes and the distribution of Earth masses on the terrestrial globe. Based on his examination of stratigraphy, fossils (seashells), and accurate surveying and geodesy, Biruni (1025) wrote about the possibility of shifting of large masses of Earth or sea (ocean) that may cause a shift of the Earth centroid, resulting in a change in the tilt of Earth’s axis (ed. Ārām, 1973; ‘Ali, 1967; Kennedy, 1973, p. 15; Berberian, 1992, 1997):

If the land adjoining a district had risen up, or had been sunk, then the waters in that district would have been diminished, the sources would have been sunk, the valleys would have been made deeper, and the district would have been rendered uninhabitable. Its inhabitants would have moved to another district; and people would have ascribed that destruction to old age in the former district, and the building up of the desolate land in the latter would have been ascribed to its youth. Abu al-‘Abbās al-Iranshahri [Neishāburi/Neyshāburi17] related that he had seen in a castle, known as the “White” [Qal’eh Baizā], at a distance of one farsakh [~6 km], from Shirjān [Sirjān], a town in the

We discussed previously the state of the Earth when a transfer of its parts takes place along its surface, and the consequent transfer of the intermediary parts lying in the direction of the Drift, and that the entire 17

16

Khārazm (Khwarazm or Khwārizm [sic]): The vast oasis region in the Āmu Daryā (modern Oxus) River delta in western Central Asia, south of the Khārazm Sea (modern Ārāl Lake), in Greater Iran.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084715/spe525-03.pdf by University of Wisconsin Madison user

Irānshahri (from the city of Irānshahr, modern-day Neishābur) was an Iranian polymath in the ninth century and taught Zakaryā Rāzi (Rhazes; 858–925 CE) and Biruni (973–1048 CE). None of Irānshahri’s writings is extant; however, the names of three of his books, Kitāb al-Jalil, and Kitāb al-Athir, and Masā’el U’ttabi’a have survived (Minorsky, 1942; Kennedy, 1973).

Development of geological perceptions and explorations on the Iranian Plateau Earth must necessarily Move in this direction, and that the natures of regions and their atmospheric conditions are subjected to change because of the changes of distances of those regions from the center of the Whole.

After his introductory narrative about the distribution of the surface masses on planet Earth in the direction of drift, and changes in the regional and climate environments, Biruni (1025) states that the shifting of large surface masses of the Earth (land and sea) may cause shifting of the Earth’s centroid, henceforth, a change in the tilt of the Earth’s axis (Kennedy, 1973):

Now I say that this movement—though it is haphazard, irregular, sensibly small over a sensible period of time, and takes place gradually along a diameter of the Whole—may be rotational about the center, or a resultant of the two motions whose direction may be one of the four cardinal directions or an intermediary direction, or may be an impulsive jerk because of a sudden transfer of the weights from one position to another. Such a movement would affect adversely a fundamental principle of astronomy, like the sun’s declination, though its amount in the celestial sphere remains the same. Its critical test, however, would be the altitudes of the two solstices; for if that movement happens to take place between two observed solstices, it may increase or decrease the maximum declination. But the frequent and successive observations conducted so far have not detected that accidental defect. Latitudes may be changed sensibly by that movement, and even the direction of latitude may be affected; or a dangerous displacement may be produced which can cause havoc and destruction. Therefore, latitudes should be continually observed and examined. Apart from that change, the movement may slightly affect the parallax. The adverse effect of that movement on longitude is insignificant, if the movement is to the east or to the west, but if it is to the north, or to the south, its adverse effect would be considerable, because when similar arcs are exchanged, their difference becomes apparent and their difference in magnitude becomes obvious. (Biruni, 1025, tr. ‘Ali, 1967, §61.3–21, p. 31–32; tr. Āām, 1973, p. 33–34)

Prior to the above statement, Biruni (1025) wrote that

All those changes are necessarily of long duration, and their causes are [of an] unknown nature. They have influenced man’s habitation and social development over different parts of the Earth; for when big masses of the Earth move from one side to another, their weights move with them, and the Earth cannot keep its instability, unless its center of gravity remains at the center of the universe. But its center of gravity varies in position with the variation of the distribution of mass over its surface, and so the Earth must adjust the distribution to keep its stability. Now, the distances of different regions from its center of gravity are not invariable over long periods of time. (Biruni, 1025, tr. ‘Ali, 1967, §42.15–18, §43.1, p. 17; Ārām, 1973, p. 19)

Biruni’s (1025) conclusions were based on detailed surveying in different cities and numerous calculations. His concept of motion of continents, a practical application of the rational

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084715/spe525-03.pdf by University of Wisconsin Madison user

41

viewpoint, is made in phrases such as “displacement of Earth’s crust due to Earth’s movement,” “changes in the coordinates and, therefore, distances of two fixed points,” “changes of distances from Earth’s center to its surface,” and “the consequent weather changes.” Biruni’s view is amazingly very close to the modern concept of continental drift, but it was written nine centuries earlier in 1020 CE. 6.9. Biruni’s Concept of Weathering, Erosion, and Roundness (1025 CE) Biruni (1025) also describes the process of weathering and erosion by different natural agents such as mechanical action, water, and wind acting on rocks and mountains (ed. Ārām, 1973; ed. ‘Ali, 1967; Kennedy, 1973; Berberian, 1997):

We do not know the conditions of creation, except what is observed in its colossal and minute mountains, which were formed over long periods of time, for example, the high mountains that are composed of soft fragments of rocks, of different colors, combined with clay and sand which have solidified over their surfaces. A thoughtful study of this matter will reveal that the fragments and pebbles are stones [rocks] which were torn from the mountains by internal splitting and by external collision. The stones then wear off by the continuous friction [mechanical erosion] of enormous quantities of water that run over them [water erosion], and by the wind that blows over them [wind erosion]. This wearing off takes place, first, at the corners and edges, until they are rubbed off and the stones finally take an approximate spherical shape. As a contradistinction to the mountains, we have the minute particles of sand and earth (soil). When soft fragments and pebbles accumulated in the beds of valleys, they became a compressed mass; then sand and earth mixed thoroughly with it and formed a combined mass, and when torrents of water flowed over it, that mass became embedded in the deep bottom after it had been on the surface of the Earth. The mass was petrified by the cold, because the petrification of the interior of most mountains is caused by low temperature. This is why stones melt under the influence of heat, because what is formed by low temperature dissolves by heat, and what is done by heat is undone by low temperature. Whenever we find a mountain formed from such soft stones, and there are many such mountains, we know that it has been formed as we have described, and that had sunk once and has risen once more. (Biruni, 1025, tr. ‘Ali, 1967, §42.6–15, p. 16–17; Ārām, 1973, p. 18–19)

6.10. Biruni on the Sedimentary Nature of the Gangā’s (Ganges) Basin, India (Ca. 1030 CE) Biruni (973–1048) lived in India for 13 yr (1017–1030), and the result of his observations was composed in his Book of India (ca. 1030; ed. Sachau, 1910; Nasr, 1968; Soduqi-Sahā, 1983; Salam, 1984; Berberian, 1997), where he wrote that: One of these plains is India, limited in the south by the above-mentioned Indian Sea, and on all three other sides by the lofty mountains [the Himalaya], the waters of which flow down to it. But if you have seen the soil of India with your own eyes and meditate on its nature—if

42

M. Berberian

you consider the rounded stones found in the earth however deeply you dig, stones that are huge near the mountains and where the rivers have a violent current; stones that are of smaller size at greater distance from the mountains, and where the streams flow more slowly; stones that appear pulverized in the shape of sand where the streams begin to stagnate near their mouths and near the sea—if you consider all this, you could scarcely help thinking that India has once been a sea which by degrees has been filled up by the alluvium of the streams. (Sachau, 1910, v. 1, p. 198; Nasr, 1968, p. 114)

6.11. Biruni’s Mineralogy Book (1050) Biruni’s extensive lapidary on the sum of knowledge about precious stones, minerals, and metals was written when he was 80 yr old. The masterpiece, describing a great number of precious and semiprecious stones, minerals, and metals, was mainly based on his long lifetime observation, experience, and studies, and the mineralogical apparatus, the specific-gravity flask, constructed by him to measure the specific gravity of minerals and metals. Comparison of his calculations with those of the modern ones shows the remarkable accuracy of Biruni’s research work (Biruni, 1050; Krenkow, 1946; Anawati, 1976, 1979, 1989; Sa’id, 1989). Biruni also procured and cited previously written unique mineralogical treatises prepared by Abu Yusuf Ya’qub ibn Ishāq al-Kindi (Kandi, Alkindus, 801–873; al-Jawāhir wal-Ashbā’, 872), Hamzeh b. Hassan Esfahāni (893–971, al-Jawāhir), and Abu Sa’d Nasr b. Ya’qub Dinawari (sic, Dinévari, ca. 989, Iranian scholar, written in Persian). He definitely had access to the Sanskrit and Greek books on the subject of precious and semiprecious stones, as well as a book by Zakaryā Yuhannā (John) ebn Māsawaiyh (Mesue, 777–857; Assyrian Nestorian Christian and son of a pharmacist and physician working at the Academy of Gundishāpur in Iran) named Jewels and their Characters (al-Jawāhir wa Sefātehā; Sa’id, 1989). Biruni (1050) elaborately described physical characteristics of 27 minerals, eight metals, and shells, glassware, and ceramics found in Europe, Asia, and Africa. Furthermore, he discussed the origin of minerals and metals, growth of minerals, and included their physical and chemical properties and therapeutic effects (Biruni, 1050; Sa’id, 1989; Anawati, 1989; Berberian, 1997). 7. SOCIOECONOMIC AND POLITICAL PROBLEMS FOLLOWED BY FOREIGN INVASIONS AND LONG HIATUSES IN SCIENTIFIC THOUGHT (1030–1055, 1218–1334, AND 1370–1392 CE) The Iranian two-century “intermezzo intellectual zenith” was adversely affected during devastating invasions by the Central Asian nomadic Turkic and Mongolian tribes. Hence, once again, the infrastructure of the country collapsed and resulted in stagnation of science, engineering, philosophy, culture, art, literature, and life, when people turned away from the rationalistic traditions of the Iranian civilization.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084715/spe525-03.pdf by University of Wisconsin Madison user

After the invasion of the Saljuq Turks (1030–1055 CE), who ruled from 1055 to 1218 (Frye, 1975a; Boyle, 1965), the nomadic tribe rulers were influenced by the Iranian court, culture, heritage, and language, and Iranian art and architecture flourished with Iranian specialists and intellectuals (Pope, 1938, 1965; Hutt, 1984). The majority of the projects, especially the Nezāmiyeh schools, seminaries, mosques, and tomb towers of the rulers, were fulfilled under the Iranian scholar and prime minister Nezām al-Molk Tusi, ca. 1063–1092 (Boyle, 1965). Although less than the previous period, scholars such as Omar Khayyām Neyshāburi (1038 or 1048–1123 or 1132 CE), Ghazzāli (1058–1111 CE), Sohrevardi (1154–1191 CE; master of illuminationism and author of Theory of the Origin of Light), and others flourished during this era. In 1184 CE, the Armenian physician Mikhitār Herātsi (ca. 1120–1200 CE), who was born in the city of Her in Iranian Armenia (modern-day Khoi in NW modern Iran), in his fundamental work on Relief of (Consoling in) Fevers (Herātsi, 1184, ed. L. Ohanessyan, 1955, p. 209), wrote about medical properties of some minerals (Ārkādi Kārākhānyān, 28 July 2015, personal commun.). The Yerevan State Medical University is named after Herātsi. The Saljuq rule came to an end with the brutal invasions of Genghis (Chingis, Changiz) Khān and Hulāgu Mongols (1218–1334 CE) and later by Tamerlane (Taimur-e Lang, lit. Timur the Lame, 1370–1392 CE), who devastated the country and destroyed most of the books and libraries (Boyle, 1965; Jackson and Lockhart, 1986). After establishment of the state under the Mongols (1256–1334 CE) and Timurids (1405–1491 CE), and establishment of “Pax Mongolica [Mongol Peace!]” (Boyle, 1965; Surhone et al., 2011), once again Iranian art, architecture, astronomy, and literature thrived among Iranian scholars (Wilber, 1955; Hutt and Harrow, 1977, 1978; Hutt, 1984; O’Kane, 1987; Golombek and Wilber, 1988; Blair and Bloom, 1995; Kedar and Wiesner-Hanks, 2015). The Iranian scholar, statesman, historian, and physician Rashid al-Din Fazlollāh Tabib Hamédāni (1247–1318 CE) became the powerful prime minister of the Ilkhānid Mongol Mahmud Ghāzān Khān (1295–1304 CE) and Oljāitu (1304– 1316 CE). In 1304, he wrote Compendium of Chronicles (Jāme’ al-Tavārikh), covering the history of the Mongol dynasty in Iran (Rashid al-Din Fazlollāh Hamédāni, 1304; Melville, 2008). He wrote that after establishment of the Mongols in Iran, philosophers, astronomers, scholars, and historians of all religions and nations gathered in the glorious court possessing copies of the histories, stories, and beliefs of their own people (Rashid al-Din Fazlollāh Hamédāni, 1304). Later, Ulugh Beig (Mirzā Muhammad Tāraghay [Tāreq], b. Shāhrokh, 1394–1449), who was a famous Iranian mathematician and astronomer, became the Timurid ruler of Samarkand. Scholars such as Jowhari Neyshāburi (ca. 1196 CE), Nasir al-Din Tusi (1201–1274 CE), Sa’di Shirāzi (1210–1291 CE), Gutb al-Din Shirāzi (1236–1311 CE), ‘Abdollāh Kāshāni (ca. 1301 CE), Hāfez Shirāzi (1325–1389 CE), and Ulugh Beig (1394–1449 CE) appeared during the Saljuq and Mongol period.

Development of geological perceptions and explorations on the Iranian Plateau Abi ebn al-Barakāt Jowhari Neyshāburi wrote a book on the precious and semiprecious stones and metals, Javāher-Nāmeh Nezāmi (The Nezāmi Mineralogical Book), ca. 1196, using data from Biruni’s mineralogical masterpiece (Afshar, 2004). Later, Nasir al-Din Tusi compiled his mineralogical book, TansukhNāmeh Ilkhāni, ca. 1256–1259, where he described 71 minerals. He definitely used Biruni’s book of mineralogy (Modarres Razavi, 1984). Furthermore, Tusi established the Marāgheh astronomical observatory ca. 1259 in northwest Iran (Hamédāni, 1311). Based on Nasir al-Din Tusi’s book, Abolqāsem ‘Abdollāh Kāshāni (ca. 1301 CE) compiled a book of mineralogy (Afshar, 1996), and Sadr al-Din b. Mansur Dashtaki Shirāzi compiled his Javāher-Nāmeh Soltāni (ca. 1478?; University of Tehran Library, n.d.). All these mineralogical books largely profited from Biruni’s book of mineralogy (Biruni, 1050). In 1501, the Safavid dynasty of Turkic origin converted the official state, constitution, institution, court, and political religion from Sunnism to Shi’ism (Jackson and Lockhart, 1986; Perry, 2010). This drastic change in the history and politics of Iran gradually and progressively established a powerful independent hierarchical clergy, Shi’a theology and jurisprudence, and Shi’ite political regime (Nasr, 1974; Jackson and Lockhart, 1986; Avery et al., 1991; Perry, 2010). This historical process, which was initiated during the Safavid dynasty (1491–1722 CE), continued and strengthened by the Afshārids of Turkic origin (1736–1796 CE), Zands (1750–1779 CE), and Qājārs (1779–1925 CE) of Turkic origin (Jackson and Lockhart, 1986; Avery et al., 1991). During the Safavid dynasty, especially ca. 1587–1628, the majority of Iranian scholars focused on theology of Shi’a, and Iranian art and architecture culminated in gorgeous mosques and seminaries, particularly in the capital city of Esfahān (Pope, 1938, 1965; Hutt, 1984; Blair and Bloom, 1995). Based on the existing precious and semiprecious stone books, Hossain b. ‘Ali Baihaqi (Ma’dan al-Javāher fi Ma’refat al-Javāher, sixteenth century), and Najm al-Din Eskandar Āmoli (Sefāt al-Javāher, seventeenth century) compiled their mineralogical books (University of Tehran Library, n.d.). John-Baptiste Tavernier (who made nine journeys into Iran between 1632 and 1668) wrote about iron, copper, lead, gold, and silver in Iran and added that the mines were anciently wrought. He added that Shāh ‘Abbās Safavid (1587–1628 CE) tried to activate some gold and silver mines, but he found that the expenses would be more than its profit (Tavernier, 1681). John Chardin (1664–1670 and 1673– 1677 CE in Iran) mentioned numerous iron, copper, lead, and silver mines and naphtha springs in different parts of the country; he added that the Persians were too slothful to make any discoveries (Chardin, 1811). Since the mid-sixteenth century, influence of foreign powers in the Iranian court and the rivalry among the Ottoman Turks, Great Britain, Russian Empire, and France plunged Iran into a series of political conflicts, wars, hardships, and losses in the state and destiny of the country. Therefore, little scientific activity was conducted in Iran, and no progress in earth sciences was recorded.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084715/spe525-03.pdf by University of Wisconsin Madison user

43

Despite all the setbacks of the Turko-Mongol invasions, the new rulers gradually became Persianized, and some Iranian scholars were selected as prime ministers (grand viziers, vazir). Some of these prime ministers, such as ‘Amid al-Molk Kondori, Nezām al-Molk (during the Saljuq period), Shams al-Din Mohamad Joveini, Rashid al-Din Fazlullāh (Mongols), and Amir Kabir18 (Qājārs), founded the Iranian educational systems, supported Iranian scholars, and contributed to Iranian literature, art, architecture, science, and medicine. Unfortunately, most of these dedicated Iranian prime ministers and scholars were executed by the kings (for coverage of the scattered geological investigations in Iran since the eleventh century, see Berberian, 1997, 2014). 8. NINETEENTH- AND TWENTIETH-CENTURY GEOLOGICAL EXPLORATIONS After several centuries of stagnation in science and engineering, geological investigations of the Iranian Plateau were initiated by foreign explorers, mostly in the middle and late nineteenth century, followed by systematic studies by European geologists until the 1970s. Iranian geologists were gradually trained from 1950s, and their contributions have accelerated since 1970. 8.1. Nineteenth Century to 1950 During this period, the incompetent and corrupt rulers of the Qājār Dynasty of the Turkic tribes (1785–1925 CE) did not have proper control of the country, and the fate of Iran and its borders was decided and marked by the British and Russian Empires. The rivalry between the Russian and British Empires paralyzed the state for a long period, and no scientific activity was conducted by the Iranians. In 1804, Fath ‘Ali Shāh Qājār started the First Russo-Persian War (1804–1813 CE). This resulted in defeat, and under the Treaty of Golestān (Gulistan [sic] in literature, 1813), the country lost the northern Caucasus (Georgia, present-day republics of Āzarbāijān, Dāghestān, and eastern Caucasus) to the Russian Empire. In addition, Russia gained the exclusive right to maintain warships on the Caspian Sea (paradoxically, the treaty text was drafted by the British diplomat Sir Gore Ouseley, who wielded great influence at the Qājār Dynasty Court!). The treaty was the beginning of a long period of rivalry and conflict between the British and Russian Empires for supremacy in Central Asia and Iran (“The Great Game”), which lasted to the Anglo-Russian Entente Convention in 1907 (Avery et al., 1991; Majd, 1994). Under the Anglo-Russian agreement of 31 August 1907, the country was divided into the British and

18

In a decree dated August 1849, Premier Amir Kabir (1848–1851) asked Monsieur John Dāvud Khān Armani (Dāvid Dāvidiān; Tāvit Tāvitiān in Armenian) to travel to Austria and employ Austrian professors to teach in medicine, engineering, mining/geology, and military services in the newly established Polytechnic (Dar al-Fonun; officially started working on 29 December 1851) in Tehran. The professors arrived in Tehran on 24 November 1851; among them, Chrnota was the mining/geology professor (Adamiyat, 1975; Ashtiani, 1984).

44

M. Berberian

Russian influence spheres, with the passive Iranian king residing in his palace in Tehran. The third Ottoman-Iranian war (1821–1823 CE) was followed by the Second Russo-Iranian War, which ended in another disastrous defeat with the signing of the Treaty of Torkamanchāi (Turkmenchay [sic] in the literature, 1828) that annexed the southern Caucasus of Iran, north of the Araxes (Aras) River to Russia (Armenia, Nakhjavān, Tālesh). Later in 1857, during the Anglo-Persian War, Iran signed the Treaty of Paris and lost the sovereignty of Harāt (Herat [sic] in present-day Afghanistan), and the British Empire also gained control of the Persian Gulf. Finally, in 1881, under the Treaty of Akhal, Russia completed its conquest of the Iranian province of Khārazm, including Turkmenistan and Uzbekistan (Avery et al., 1991; Majd, 1994). Therefore, the Iranians went through a century of military defeats, imposed treaties, political corruption, and religious superstitions; meanwhile, modern science and engineering were not introduced into their educational and social life, and hence prominent scientists could not bloom. In the early twentieth century, the political and social situation in Iran was still not favorable for any scientific activity, although the Constitutional Revolution of 1906 was a positive trend: It had forced the Qājār king Muzaffar al-Din Shāh to sign a constitution limiting the king’s power and forming an elected parliament (Majels Shorā’ Melli). In 1908, the Russian soldiers shelled the newly established Iranian Parliament at Bahārestān Square in east Tehran, and in 1909, Russia supported the deposed Mohammad ‘Ali Shāh Qājār against the Constitutional Revolution (Avery et al., 1991). After 1919, the Soviet Union helped to create the first Soviet Republic of Guilān Province, and after World War II, the Soviet Union supported independence movements in Kordestān (Kurdistan [sic]) and Āzarbāijān (Azerbaijan [sic]). In 1925, the Qājār Dynasty gave way to the Pahlavi Dynasty with the rise of Rezā Shāh (1925–1941), a former Iranian officer and nationalist who overthrew the Qājār dynasty and began modernization of Iran while maintaining a dictatorial rule, a norm in the region. In 1941, as World War II spread far and wide, the British and the Soviets occupied Iran; Rezā Shāh Pahlavi was forced to live in exile, and his young unexperienced son Mohammad Rezā Shāh became his successor (1941–1979). At the end of World War II, the Soviet Union refused to withdraw its Red Army troops from northern Iran. Since 1941, the Soviet Union has supported the Iranian communist Tudeh Party, fueling the Cold War abroad and tension in the country. Under such circumstances, the country struggled for its survival (Avery et al., 1991). No proper knowledge of the geology of the country existed during the first half of the nineteenth century. M. Trezel (in Jaubert, 1821), during a trip to Armenia and Iran, was possibly the first person to recognize the granitic rocks in the area west of Rasht along the Tālesh-Boghrov Dāgh ranges, SW of the Caspian Sea (Fig. 1). In the summer of 1921, James Baillié Fraser, Esq., on his tedious voyage from Bombay (India) to Bushehr (Iran), with stops at Rausul-Heed and Muscat, wrote a scanty brief description of rocks in the Qeshm and other Persian Gulf

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084715/spe525-03.pdf by University of Wisconsin Madison user

islands, with no reference to the salt domes prevalent on those islands. He recorded “serpentine rocks” and asbestos at the Cove of Muscat (Fraser, 1824). Due to lack of major geological and mining exploration activity and data in the country, John Malcom, the British diplomat and administrator of India (Malcom, 1827 in Curzon, 1892, v. II, p. 510), wrote that: “Persia does not abound in valuable minerals; iron and lead, however, are found in many places.” Later, Polak (1865) stated that: “Scarcely any country of the Earth can vie with Persia as to riches in metals, especially copper. Its unbound wealth in coal, iron, and copper deposits only awaits exploration in order to set on foot a mighty industry” (in Curzon, 1892, v. II, p. 510). Dr. M.C. Bell (1840) published his geological observations on the Alborz Mountains (Fig. 1) along the road from Tehran via Damāvand, Firuzkuh, and Tālārrud Valley to the Caspian Sea shore, retuning to Tehran via Āmol and Harāz Valley. He reported volcanic rocks, limestone (at one location with two small ammonite fragments and bivalves), shale, sandstone, breccia, conglomerate (with “rolled” [rounded] pebbles of quartz, granite, and shell), claystone, and coal deposits in scattered areas with the location names and a road map of the areas visited. However, no stratigraphic correlation or connection was given. Freidrich Alexander Bühse, the Latvian botanist (1821– 1898), visited northern Iran (Guilān to Astarābād–Setārehābād, modern Gorgān) during his mountain hikes in 1847–1849, and his writings contain a few geological observations (Bühse, 1849, 1892). Based on notes collected by Bühse (1849) and Lieutenant Colonel Woskoboinikow (1846) from northern Iran, Dr. C. Grewingk (1853, 1881) published some geological description of routes from Araxes to Shāhrud Rivers, Astarābād [modern Gorgān], and other places with descriptions of various rock units such as hippuritic limestone, Silurian–Devonian outcrops near Qazvin and Shāhrud River (W Alborz), Liassic coal seams (the Shemshak Formation) and Cretaceous rocks in the Alborz, Jurassic rocks at Māsuleh and Shāhrud (W Alborz), and nummulitic limestone in Shirkuh of Yazd and at Kuhrud (Qohrud) near Kāshān (in Central Iran). He also referred to the volcanic rocks of Sahand and Sabalān volcanoes, Qazvin, and Damāvand, and metamorphic rocks and granites near Lake Urumieh, Karadāgh (Arasbārān), Māsuleh, and Tehran. The report contains a geological map of the Āzarbāijān, Guilān, and Māzandarān Provinces of northwestern and northern Iran. The British geologist and archaeologist William Kenneth Loftus (1820–1858) was a member of the British Commission under Lieutenant Colonel Williams in 1849–1852 working at the Susā (Shush) archaeological mound excavation in western Zāgros, SW Iran. In his paper on the geology of the TurkeyIran border area, Loftus (1855) presented a map showing broad geological features of the area, including granite and metamorphic rocks and fossiliferous limestones, the oldest of which was regarded as Silurian. Loftus (1855) applied the term “Nummulitic Series” for the Lower Tertiary beds up to and including the Oligocene–Miocene Āsmāri Formation of James and Wynd

Development of geological perceptions and explorations on the Iranian Plateau (1965) in the Zāgros as well as in northwest Iran. Loftus (1855) named and described “gigantic species of Alveolina,” which later Carpenter and Brady (1869) proposed as the genus Loftusia of the Eocene. Loftus also named the Late Tertiary rocks of the Zāgros Mountains of the southwest Iran as “Gypsiferous Series” (later the Miocene Fārs Group and Pliocene Bakhtiāri Formation; and Upper Red Formation of Central Iran) at the Zāgros foothills. Loftus (1855) used the term “Supranummuliten Kalke” for the modern Qom Formation (Middle–Upper Oligocene–Lower Miocene) in Central Iran. Three years later, Abich (1858) called the rock unit the “Urmi Series.” In 1862, R.G. Watson climbed the Damavand volcano in central Alborz, N. Iran, and described a summit of the volcano being covered by snow with sulfur deposits, sulfur fumes, and hot ground (Watson, 1862). In 1888, in a paper on the geology of the Sabalān volcano in NW Iran, Sjögren described a cirque glacier at an elevation of 4740 m that descended to ~3800 m on the Sabalān Quaternary volcano. From 1885 to 1925, an Austrian and French expedition continued to explore and collect samples of the late Miocene vertebrate fauna of the “Marāgheh Bone Beds” at the southern foot of the Sahand volcano, southeast of Lake Urumieh, in northwest of Iran (Fig. 1). The studies were carried out by Kittl (1885), Pohlig (1886), Lydekker (1886), Rodler and Weithofer (1890), and de Mecquenem (1905, 1906, 1908, 1911, 1924–1925). 8.1.1. Otto Wilhelm Hermann von Abich’s Time (1857–1882 CE) The German geologist Otto Wilhelm Hermann von Abich (1806–1886), who spent much of his professional life in the Caucasus and Armenian Highlands (Abich, 1878; Milanovsky, 2007), wrote about the sedimentary, metamorphic, and volcanic rocks in Āzarbāijān (NW Iran), including those on the Sabalān volcano, near Tabriz and Mero Dāgh, outcrops of Devonian and Carboniferous limestones in northern Āzarbāijān, outcrops of Jurassic and Cretaceous limestones in the area south of the Araxes River, and his “Urmi Series,” which later became the Miocene Qom Formation of Central Iran (Abich, 1858). In the same report, Abich (1858) used the term “Supranummulitenkalke” (first applied by Loftus, 1855) for the Upper Oligocene– Lower Miocene limestones above the nummulite-bearing beds in Āzarbāijān, which is now known as the Qom Formation. A year later, Abich (1859) used the German term “Bergkalk” (lit. “Mountain Limestone”) for the Upper Permian limestone and interpreted shales in Transcaucasia and Iran as a Carboniferous limestone. In 1878, he found some brachiopods and ammonites in these limestones. From 1833 to 1878, Abich wrote 23 reports and books. Abich also communicated the reports written by the Russian explorer and Council at Tabriz, Nikolai Vladimirovich de Khanikoff/Khanykof, containing geological observations in Āzarbāijān referring to Miocene limestone in the Shāhi Island of Lake Urumieh, as well as notes on earthquakes recorded in Tabriz (Khanikoff, 1858, 1861; Berberian and Arshadi, 1976;

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084715/spe525-03.pdf by University of Wisconsin Madison user

45

Berberian, 1994, 2014). Khanikoff discovered the late Miocene vertebrate fauna of the “Marāgheh Bone Beds” at the southern foot of the Sahand volcano in northwest of Iran in 1840, and the area was studied by Abich (1859), von Brandt (1870), and Grewingk (1881). 8.1.2. William Thomas Blanford’s Time (1872–1876 CE) Blanford (1832–1915 CE), the British geologist and naturalist who was appointed as a member of the Persian Boundary Commission of the British Empire, spent some time in Iran and wrote a few reports about his expeditions (Blanford, 1872, 1873, 1876). Blanford (1872) was possibly the first geologist to name the Miocene molasse deposits of the remote Baluchestān and Makrān (Fig. 1) the “Makrān Series” and dated the beds as postnummulitic (Eocene) in age. His study was later followed by Duncan and Sladen (1880) and Pilgrim (1908). Blanford (1872) also introduced the term “Hormuz [sic, in almost all literature, ‘Hormoz’] Salt” in the Persian Gulf islands, which is still used by geologists today. Pilgrim (1908) used the term “Hormuz Series” for the Hormoz Salt domes in the Zāgros. The challenge of dating the evaporites and sediments, and igneous rocks of the Hormoz Salt complex was followed by Pilgrim (1922, 1924), Richardson (1926), Harrison (1930), and many more. In 1873, Blanford gave a brief description of the sands, clays, and gravels covering a large part of the Iranian playas. He found some early Paleozoic fossils and regarded them as Silurian (Blanford, 1872). Blanford (1876) briefly summarized the geological studies in Iran from 1822 to 1874, followed by a section in his book on general geological features and formations, as well as mineral resources. His information was categorized into the following headings: (1) Metamorphic and Granitic Rocks; (2) Paleozoic (Carboniferous and Devonian); (3) Secondary [Mesozoic] (Liassic and Jurassic; Cretaceous Hippuritic Limestone); (4) Tertiary (Nummulitic Series, Gypsiferous Series, Makrān Group); and (5) Quaternary and Recent. 8.1.3. Emile Tiezte’s Era (1875–1886 CE) One of the most important nineteenth-century geological publications on Iran was written by the Austrian geologist Emil Ernest August Tietze (1845–1931 CE), titled Die Mineralreichtümer Persiens (Vienna, 1879), among his 11 publications from 1875 to 1886 (Tietze, 1875a, 1875b, 1875c, 1875d, 1877a, 1877b, 1877c, 1878, 1879, 1881, 1886) on the geology of the Siāhkuh region, tectonics of the Alborz, Damāvand volcano region, mineral resources, young sediments of northern Iran, and contribution to the geology of Iran, with references to earlier geological publications on Iran. It was Tietze (1877b) who designated the volcanic and tuffaceous rocks of the Eocene Karaj Formation (Dedual, 1967) of the Alborz Mountains (Fig. 1) north of Tehran as “Grüne Schichten” (i.e., the “Green Beds”), but his assumed Triassic age was not correct. The term “Green Beds” was used in all reports until 1967. The complex was later visited by Stahl (1911) and Rivière (1934b, “Couches Vertes”), who assigned the formation to the Oligocene. Later,

46

M. Berberian

Schenck (in Clapp, 1940), Rieben (1942), and Schröder (1945a, 1965) gave an Eocene age for the formation. Tietze (1879) named the Miocene Upper Red Formation of Central Iran the “Salz-oder-Gips formation.” Tietze (1881) was possibly the first person to describe the loess deposits at the northern foothills of the Alborz Mountains. His study was later followed by Alfons Gabriel (1939), Clapp (1940), and Sedlacek (1956). Tietze (1877b) also noticed the “Gorgān Schists” and named it as “Schiefergesteine der Gegend von Asterabad.” He wrote about the Devonian–Carboniferous beds in the Alborz (later the Mobārak Limestone of Assereto, 1963), and the “Lias Sandstein” (Tietze, 1877b, 1878). The latter was named in 1897 and 1911 as the “Rhät, Lias und Dogger Sandsteine Schieferthone,” which was formally named in 1966 as the Upper Triassic to Lower–Middle Jurassic Shemshak Formation by Assereto (1966a). Later, Tietze (1881) recognized the fossiliferous Sarmatian rocks of the Caspian beds along the coastal Caspian Sea. 8.1.4. Alexander von Stahl’s Era (1893–1933 CE) The German geologist, engineer, and surveyor Alexander Friedrich von Stahl was one of the most important figures in the earlier geological work in Iran (Stahl, 1893a, 1893b, 1893c, 1894a, 1894b, 1894c, 1895, 1897, 1903a, 1903b, 1904a, 1904b, 1905, 1907, 1909, 1911, 1925, 1927a, 1927b, 1928, 1933), creating the initial stratigraphic outline of northern Iran (Jājrud and Lār Valleys). Stahl (b. 1850) served as postmaster-general of the Persian Telegraph Company (possibly started working in 1877?), and in this capacity, he traveled extensively throughout the country, thereby making numerous geological observations, which he described and published between 1893 and 1933 in 22 articles in the German language, some with excellent maps. He was very interested in surveying, mapping, and improving the roads, geology, and mineral resources, and he prepared a surveyed map of the Tehran region in 1900 on a scale of 1:210,000, along with the surrounding mountains and tributaries (Stahl, 1900), as well as maps of other parts of Iran. About mineral resources Stahl (1894a) wrote hat: “About ten kilometers north of Weshnave [Veshnoveh], between Kum [Qom] and Kashān, out of dark brown volcanic rock of aphanitic structure, there comes up a vein of calcite, dipping vertically and copper pyrite and black copper ore mingled within” (Stöllner, 2005, p. 196). Stahl also studied the Bāfq iron ore deposit in southeast Central Iran (Fig. 1) and estimated ~200 million tons of 60% iron reserves and >2% of phosphorus. On his way from Yazd to Bāfg, a group of bandits tried to attack his camp during night, but after the loud cry of the tall Stahl, the thieves escaped. Probably Stahl’s most important report is a general review titled Persien: Handbuch der Regionalen Geologie (Heidelberg, 1911), including two folded plates, one being the first, smallscale (1:840,000) geological map of Iran (excluding the easternmost part of the country). Stahl (1911) considered the schists and marbles of the Darreh Anjir Range, and the Anārak gneisses

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084715/spe525-03.pdf by University of Wisconsin Madison user

overlying granite, which he viewed as the oldest rocks in Central Iran, to be of Archean age, and hence he developed his concept of the “Median Mass” covered by shallow-marine and continental deposits. Stahl (1911) also considered the “sillimanite gneiss” of the Hamédān area to be Archean as well (which, in fact, is Mesozoic; Berberian and Alavi-Tehrani, 1977a). For Stahl (1911), Central Iran was a vast Archean stable plateau (“Median Mass”) surrounded by the northern and southern fold belts of the Alborz and Zāgros (Fig. 1). Stahl (1911) argued against the Triassic age for the “Green Beds’’ suggested by Tietze (1877b) and assigned it to the Oligocene (later known as the Eocene Karaj Formation; Dedual, 1967). He visited the “Gorgān Schists” of the northern Alborz Mountains (Berberian et al., 1973), named the Middle–Upper Oligocene–Lower Miocene Qom Formation as “Miozäne Kalke,” and called the Miocene Upper Red Formation as “Basal Red Group of the Neogene” (Stahl, 1911). Stahl’s 1911 concept of “Median Mass” was accepted and established by geologists of the oil companies, where they stopped their geological investigations at the southwestern edge of the supposed “Median Mass” along the Main Zāgros reverse fault (Fig. 1). The “Median Mass” also became an established concept by de Böckh et al. (1929) and Bogdanoff et al. (1964). The concept was partly shattered in 1938 when Baier (1938), by relying mainly on the occurrence of graphite, wrote that the greater part of the Anārak metamorphic rocks of Stahl’s (1911) “Median Mass,” as well as metamorphics of the Sirjān belt (Fig. 1), were metamorphosed equivalents of the Lower Jurassic coal-bearing shales (later known as the Shemshak Formation by Assereto, 1966a). Baier (1938) also thought the metamorphosed Kharānaq limestones were Triassic in age (the rocks were in fact Cretaceous). Baier (1938, see Harrison, 1968, p. 183; Stöcklin, 1968a, p. 1232) stated that the Alborz and the Zāgros Mountains in the north and south were parts of two enlarged “double orogens” covering the whole plateau, with Central Iran as a “mobile internal zone.” This idea was also expressed by Holmes (1944). Baier compared the Iranian structures with those of the Penninic zone of the European Alpine belt. Stahl (1911) also used the term “vieux grès rouge” (the obsolete ‘Early Devonian old red sandstone’ used by Abich, 1859), for the first time (see Stöcklin, 1972, p. 178, 143) in Iran for the present Lower Cambrian Lālun Sandstone (Assereto, 1963, p. 19). The term, with its incorrect Devonian age, was widely used until 1963 by Rivière (1934b), Furon (1941), Bailey et al. (1948), and many more. Following Tietze (1877b), Stahl (1897, 1911) reported on the Devonian–Carboniferous rocks in the Alborz that later became known as the Mobārak Limestone (Assereto, 1963). In 1927, Stahl published an account of the physiography and drainage along his line of traverse from Ardebil through the Boghrov Dāgh section of the Tālesh Mountains and on to the Māsuleh, Shāhrud, and Sefidrud areas of northwestern Iran (Stahl, 1927b).

Development of geological perceptions and explorations on the Iranian Plateau 8.1.5. Other Activities (1885–1908 CE) In 1885, Rodler described the late Miocene vertebrate fauna of the “Marāgheh Bone Beds” at the southern foot of the Sahand volcano, southeast of Lake Urumieh, in northwest Iran (Fig. 1). Later, Rieben (1935) correlated the “Marāgheh Bone Beds” with the “Couche à Lignite” (“Lignite Beds”) and the “Couche Lacustres à Poisson” (“Lacustrine Fish Beds”) of Tabriz. The latter was discovered in 1908 by Mecquenem (Brachylebias Persicus). Apparently, Lieutenant H. Vaughan of the seventh Bengal Infantry was the first European to describe the surficial geology of the Great Kavir playa in northern Central Iran from first-hand observation (Goldsmid, 1890; Vaughan, 1893; Jackson et al., 1990). At the beginning of the twentieth century, Ernest Watson Vredenburg (1870–1923), who was working on the Tertiary paleontology of India for the Geological Survey of India, used the term “Flysch” for the Late Cretaceous to Oligocene rocks exposed in the modern Iran-Pakistan Makrān border region (Fig. 1), considered as Late Cretaceous (Vredenburg, 1901). He also wrote a paper on the occurrence of the genus Orbitolina in India and Iran (Vredenburg, 1908). A year later, Vredenburg (1909) published a paper on the Upper Cretaceous hippuritic limestone of Sistān in eastern Iran. The French geologist, mining engineer, and archaeologist Jean-Jacques de Morgan (1857–1924), who excavated the Susā (Shush) archaeological mound in the western Zāgros, SW Iran (Amiet, 1994; Perrot, 2013), made significant contributions to the geology of Iran (Morgan, 1892, 1894, 1895, 1900, 1905a, 1905b, 1907; Douvillé, 1900, 1904a, 1904b, 1905). He studied the geology of northwestern Iran, the Caspian coastal region, Kordestān, Lorestān, and Khuzestān. Apparently, Morgan was the first geologist to recognize and report the presence of oil at Qasr-e Shirin in the northwestern part of the Zāgros (Morgan, 1892). Seemingly, it was Morgan who later recommended to the Anglo-Persian Oil Company (APOC) to start exploration at the Masjed Solaymān area, where the first oil well hit the trap (Gabriel, 1952). Volume 3 of his Scientific Mission in Iran series deals with the geology and stratigraphy of the country (Morgan, 1905a). Morgan also studied the Kurdish [Kordi] language and dialect and the Mandaean texts. H. Douvillé, another French geologist, carried out paleontological studies in Iran and published nine reports from 1900 to 1910 (Douvillé, 1900, 1902, 1904a, 1904b, 1905, 1910). The British geologist J.A. Douglas studied the paleontology of the country, especially associated with Carboniferous, Permian, Triassic, and Miocene–Pliocene deposits of SW and SE (Baluchestān) of Iran, and published seven reports during the interval 1927–1950 (Douglas, 1929, 1936, 1950). Douvillé reported fragments of trilobites, orthocerides, and brachiopods of possibly Silurian age collected from Bur Kuh north of the Jāz Muriān Depression. Another British paleontologist, L.R. Cox (1897–1965), published three reports during 1934–1938 on the paleontology of Iran, with a description of new species (Cox, 1936).

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084715/spe525-03.pdf by University of Wisconsin Madison user

47

Concurrently, the Austrian geographer Hans Bobek (1903– 1990) wrote about regional geography and morphology of the country, with some geological references, in his 12 reports from 1934 to 1953 covering the Alborz, Takht-e Solaymān, Great Kavir of Central Iran, and other parts of the country. Apparently, Busk (1934, 1935, 1937) was the first European recognizing glaciers on the High Alborz Mountains (Fig. 1) of north Iran; he was followed by Bobek (1934) who made sketch maps of the area. As a member of the German climbing expedition of 1936, Bobek (1937) mentioned glaciers on Damāvand volcano in central Alborz Mountains, Northern Iran. In 1957, Bobek produced a 1:100,000 scale map of the Takht-e Solaimān region showing the glaciers (Bobek, 1957). In 1963, Bobek wrote a report on the implications of the Quaternary climate change in Iran (Bobek, 1963). Bobek reported red/violet limestones interbedded with sandstones and red/green shales containing Cruziana and Ordovician trilobite fossils in Hezār Cham of ‘Alamkuh in the Alborz (Bobek, 1934; Dietrich, 1937). He also studied the Great Kavir of northern Central Iran (Bobek, 1959, 1961, 1969). Elles (1930) reported some dark shale with Lower Valentian graptolites on the lower slopes of the Faraghān Mountain in the High-Zāgros of southeast Iran. Although petroleum was discovered in Iran in 1908, the Iranian share and income were very low.19 Hence, the country needed additional sources of energy and revenue. Therefore, in the early 1930s, Rezā Shāh Pahlavi directed his government to sponsor the Swiss and German geologists to survey the Iranian mineral resources, especially coal, iron, copper, and lead deposits and improve the mining techniques and safety.20 During this period, Böhne (1932a, 1932b), Diehl (1944), and Ladame (1945) studied the distribution of mineral resources and presented catalogues covering the country with several maps and geological cross sections. The scattered scientific activities were halted during World War II and the Anglo-Soviet invasion and occupation of the country.

19 The Iranian annual oil revenue was £0.5 million in 1919, £1.3 million in 1930, and £16 million in 1950. The oil revenues dropped to £7 million in 1950, £0.0 in 1951, and £0.1 million in 1953, and the country suffered budget deficit (Amirsadeghi, 1977). Iran’s budget deficit in 1959 was US$101 million (Fisher, 1968). 20

Alfons Gabriel (1894–1976), the Austrian geographer and travel writer who made several trips to Iran and its deserts (1927–1928, 1933, 1937), wrote five books about Iran. After his 1934 visit to the Nakhlak lead mine in the Central Iranian desert, he described the problems of the mining technique and the perilous life of the miners (Gabriel, 1935):

Here [at Nakhlak] lead is mined and smelted from the ore in a primitive way and transported on camel back to Anārak in ingots of 30 kg weight. About 100 people work in the nearby mines. They are all without their families, come from Anārak, Chupānān, Jandaq, and other places and usually stay until they become victims of lead poisoning and are thus forced to give [up] their work. Young and old find work in Nakhlak; the wages are four to seven Qerān a day [~US$0.70 to $1.12]. They work at daybreak. In the afternoon you can see the tired people coming out of the mine with their heavy picks and an oil lamp. (Gabriel, 1935, quoted by Wulff, 1966, p. 16)

48

M. Berberian

8.1.6. Anglo-Persian (Iranian) Oil Company (APOC/AIOC) Activities in the Zāgros Since geological investigations in the Zāgros (Fig. 1) have been well documented elsewhere (Ferrier, 1982; Bamberg, 1994, 2000; Berberian, 1997), I will only briefly mention the works of a few geologists. In 1908, Dr. Henry Guy Ellcock Pilgrim (1874– 1943; Fermor, 1943), who was appointed to the Geological Survey of India in 1902, coined the terms “Fars Series” (later named as the “Fars Group” by James and Wynd, 1965) and “Bakhtiari Series” (Pilgrim, 1908; the “Bakhtiāri Formation” of James and Wynd, 1965) in his report on the geology of the Persian Gulf and the adjoining regions. In 1924, Pilgrim applied his 1908 term of “Hatat Series” (the metamorphic rocks of Siāh Hatāt in Oman) to the rocks outcropping in the Zendān Range of the western Makrān as well as to the rocks near Neyriz in the High Zāgros (Fig. 1). He extended the use of the “Oman Series” for the Neyriz ophiolites. The name was later replaced by “coloured mélange” (Gansser, 1955). Pilgrim (1924) proposed the term “Khān-i Surkh [Khān-e Sorkh] Volcanic Series” for the Miocene lavas and tuffs, as well as the “Panj Intrusive Series” in the Panj Mountains of the Kermān area, SE Iran. Pilgrim, who arrived in Iran in 1904, spent 20 yr working in the country. Sir Arnold Talbot Wilson (1844–1940), the British colonial administrator of Mesopotamia, and the resident director of the APOC in the Persian Gulf, oversaw the discovery of the first oil trap at Masjed Solaymān in the Zāgros, SW Iran in 1908. Wilson (1930, p. 130) wrote about the “Median Mass” that:

With regard to the remaining records of the shocks scattered over the Median Mass, it must be remembered that this is an irregular complex of horsts and depressions and it seems if the movements, which are irregular in direction, have continued since Cretaceous times, accompanied since Eocene by considerable volcanic action which has gone until fairly recent times, but the association of the recorded shocks with recently depressed blocks is of interest.

Following the first discovery of petroleum on 26 May 1908 at the Masjed Solaymān oil well of the Zāgros fold-and-thrust belt in southwest Iran, the London-based Anglo-Persian Oil Company (APOC) was founded for petroleum extraction, and so systematic geological investigation in the Zāgros became a fundamental task (the British government gained control of the Iranian oil industry by purchasing a majority of the company’s shares in 1914). The company was renamed the Anglo-Iranian Oil Company (AIOC) in 1935, and later, after the nationalization of the Iranian oil industry by nationalist Prime Minister Dr. Mohammad Mosaddeq, Esq., in 1954, it became British Petroleum (BP; Ferrier, 1982; Bamberg, 1994, 2000; Berberian, 1997; Badakhshan and Najmabadi, 2004). During this interval systematic geological investigations in the Zāgros were carried out, mostly by the British geologists from the ground during land surveys. Among the early geologists of the APOC/AIOC, George Martin Lees (1895–1955; Arkell, 1955), John Verron Harrison

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084715/spe525-03.pdf by University of Wisconsin Madison user

(d. 1972; Vincent, 1972), and Norman Leslie Falcon (1904– 1966) made the most important contributions by conducting geological mapping of nearly 77,700 km2 in the Zāgros. Harrison arrived in 1919 and spent 7.5 yr out of his 8 yr stay in Iran working in the field for the APOC. Lees arrived in Iran in 1922 and worked 2.5 yr in the field, out of his 4.5 yr stay in the country. Falcon joined APOC in late 1927 and stayed 7 yr in the country, out of which he spent 4.2 yr in APOC field investigations (British Petroleum Co., Ltd., 1956a; Berberian, 1997). Although little of their work has been published, they certainly laid the foundations of our present knowledge of the Zāgros Mountains. In 1924, Lees and Hugo De Böckh toured the Zāgros Mountains and identified several productive oil fields. Later De Böckh, Lees, and Richardson (Böckh et al., 1929) published the theoretical side of their findings, where they accepted the “Median Mass” concept of Stahl (1911) for Central Iran. Lees found Cambrian trilobite fossils in dark shale blocks brought up by a Hormoz Salt dome northeast of Bandar Lengeh (Lees and Richardson, 1940). Harrison and Falcon worked on the mechanism of salt dome intrusion and the role of gravitational forces in tectonics (Harrison and Falcon, 1934a, 1934b, 1935, 1936b) and together studied the colossal Saimareh landslide (misspelled “Saidmareh” in the literature) in the northern flank of the Kabirkuh anticline in the Zāgros (Harrison and Falcon, 1937a, 1937b, 1938). Harrison also prepared an undated and unpublished lexicon of Iran containing old stratigraphic names, which was later utilized by Stöcklin (1972) in preparation of his Stratigraphic Lexicon of Iran, and by Ata’ollāh Setudehnia (1972) for the Zāgros. Harrison, Falcon, Huber, Stöcklin, and their associates, with the aid of camels and mules, carried out the first geological mapping of the uninhabitable, remote, and unknown Makrān and Jāz Muriān in southeast Iran (Harrison et al., 1935–1936; Harrison and Falcon, 1936a; Harrison, 1943; Huber, 1952; Stöcklin, 1952). J.V. Harrison, who, at the age of over 80, went on an excursion to Mazandarān (northern Iran) in 1958, told Jovan Stöcklin that “out of his 90-months assignment in Iran in the 1920s and early1930s, he just spent one day in Tehran, about a month at AIOC’s headquarters in Masjed Solaymān, and all the rest on expeditions in the Company’s Concession Area” (Jovan Stöcklin, 30 May 2003, personal commun.). Harrison’s colleague in Iran, P.T. Cox, told professor E.A. Vincent of Oxford that, “hard work was probably J.V.’s prime passion and that it was defined as something involving evident physical discomfort” (Vincent, 1972, p. 1; for more information and selected bibliography of J.V. Harrison see Kummel, 1972). Beginning in 1938, aerial photographs became available for the Zāgros from the London-based company. Geological investigation was gradually followed by the Iranian Oil Company (IOC; later National Iranian Oil Company [NIOC]) in 1950 (discussed later). During this period, Laurence Lockhart (1890–1975), who worked for APOC (later AIOC) from 1919 to 1930, wrote an article about Iranian petroleum in ancient times, the summary

Development of geological perceptions and explorations on the Iranian Plateau of which was presented at the II Congrès Mondial du Pétrole in Paris in 1937 (Lockhart, 1939). Lockhart traveled extensively in Iran, and the 1920s–1950s photographs taken by him were published in 2002 (Bamberg and Melville, 2002). 8.1.7. Alborz and Central Iran Investigation in the 1930s and 1940s During the early 1930s, Ovtsinnikov (1930), Rieben (1930, 1935, 1942), Böhne (1932a, 1932b), and Rivière (1934a, 1934b) provided valuable contributions on the geological knowledge of the central Alborz Mountains (Fig. 1) with emphasis on lithostratigraphic sequences, cross sections, and geological mapping (Ovtsinnikov, 1930), coal deposits, along with a good geological sketch map of the Shemshak area on a scale of 1:100,000 (Böhne, 1932b), and the geology of Āzarbāijān in northwest Iran (Rieben, 1930, 1935, 1942). Böhne (1929, 1932a, 1932b), Diehl (1944), and Ladame (1945) studied ore deposits of the country within their geological context and presented catalogues covering the whole country, with mineral distribution maps and scaled geological profiles. Böhne reviewed the country’s ore deposits, state of smelting, and their economic importance and urbanization (1932b). Eugène Hubert Rieben (1889–1972) lived in Tabriz for many years. Apparently he (or his father?) was teaching at an American missionary (Presbyterian) school in Tabriz or Urumieh. Between 1925 and 1932, Rieben made important studies on the geology of Āzarbāijān in northwest Iran (Rieben, 1930, 1935), the Quaternary terraces around Lake Urumieh, and the south Caspian region (Rieben, 1942). Rieben (1935) named the Pliocene Lacustrine “Fish Beds of Tabriz” (originally studied by Mecquenem, 1908) as “Couche lacustres à Poisson,” and the “Tabriz Lignite Beds” as “Couche à Lignite” (first discovered by Murray, 1859). French geologists André Rivière and Raymond Furon (1898–1986) worked for a while as professors at the University of Tehran; Rivière for 6 yr starting in 1929, and Furon for 2 yr in 1936–1937. Rivière (1930, 1931a–1931h, 1932a–1932e, 1933, 1934a, 1934b, 1935a–1935c, 1936; Rivière et al., 1973) was the first geologist to create a detailed unified stratigraphic framework of the central Alborz Mountains, in addition to a geological sketch map on a scale of 1:300,000, and numerous geological cross sections showing strongly folded and faulted Green Beds (the Eocene Karaj Formation) and the Jurassic schuppen (imbricate) structure (Rivière, 1934b). In 1936, Rivière coined the term “anti-Alborz” for the range between east Tehran and Semnān. Furon taught at the University of the Musée National d’Histoire Naturelle in Paris and also taught 2 yr at the Lycée FrancoAfghan in Kabul (Kābol), Afghanistan. Furon wrote ~15 publications on Iran dated from 1931 to 1954 about salt extrusion, the “Ural axis” in the Iranian Plateau, the geology of the Great Desert (Kavir), the geology and structure of the Iranian Plateau, fusulina and nummulitic limestones in the Hablehrud valley, the geology of the Khārk Island in the Persian Gulf, and microfauna of the Acquitanian limestone in the Qom area. In 1961, he became the president of Société Géologique de France.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084715/spe525-03.pdf by University of Wisconsin Madison user

49

During the summer of 1933, Count Ardito Desio (1897– 2011) led an expedition to Iran. The expedition climbed some of the highest peaks of the Zāgros Mountains of SW Iran. Desio also developed a new route to the summit of Iran’s highest peak, Damāvand volcano (5670 m) in the Alborz. He published some scientific reports on these trips (Desio, 1934a, 1934b). In the early 1940s, the geological knowledge of Iran was summarized by Clapp (1940) and Furon (1941); the latter postulated a “Hercynian Ural-Oman-Madagasgar axis” (Furon, 1941). The existence of the lineament was accepted by Gansser (1955); however, he rejected the Hercynian age of the formation of the lineament in Iran. The imaginary lineament is still erroneously used in Iranian geological reports. American geologist Fredrick G. Clapp wrote about five reports from 1930 to 1940 on the geology and tectonics of the Iranian Plateau. In his geology of eastern Iran, covering ~100 pages, Clapp (1940) introduced the Iranian geology and mountain belts of the Alborz (Elburz [sic] in Clapp, 1940), Zāgros, and Central Iranian ranges, indicating faults and overthrusts bounding the Iranian ranges. He described the strata from Cambrian through Recent with major unconformities—those of pre-Triassic, preEocene, pre-Pliocene, and post-Pliocene unconformities being the most important. Clapp emphasized that the Iranian Plateau was still in the “process of elevation as evident from recurrent earthquakes.” The idea of uplift of the Iranian Plateau with each earthquake as stated by Clapp is very interesting because during that time, the seismicity and active faulting of Iran were not being studied or documented. In 1944, in his paper on the geological structures of Iran, J.W. Schröder (1) named the linear volcanic belt of the southwest Central Iran as “Urmia-Dokhtar Zone” (Karkas belt in Fig. 1); (2) named the metamorphic belt of Sirjān as “la Zone de Hamedan” (Sirjān belt in Fig. 1); and (3) accepted Baier’s (1938) concept of the “Central Iranian eugeosyncline” (Schröder, 1944, 1945b). He also wrote about the age of the “Couches Vertes” (the ‘Green Series,’ later the Eocene Karaj Formation; Schröder, 1945a, 1965) and the geology of the Lārak Island in the Persian Gulf, and compared the salt domes with those of the Salt Range, Pakistan (Schröder, 1946). Concurrently, based on the available data, Arthur Holmes (1944) presented a NE-SW cross section of Iran showing the Asiatic and Arabian forelands underthrusting the Central Iranian geosyncline with the bordering folded and overthrust ranges in the NE (Alborz-Kopeh Dāgh) and SW (Zāgros). The central Iranian basin was portrayed devoid of any deformation. Alfons Gabriel (1894–1976), who was an Austrian geographer and travel writer, and his wife Agnes were probably the first Europeans who crossed the Lut Desert of eastern Iran; and also crossed the Great Kavir of northern Central Iran. He published twelve reports from 1935 to 1964 (Gabriel, 1929, 1934, 1935, 1938, 1939, 1952, 1957a, 1957b, 1961, 1963, 1964). In 1935, he described the production at the Nakhlak lead-silver mine (Gabriel, 1935). Gabriel (1952) wrote a very interesting book describing the geological exploration of Iran since antiquity,

50

M. Berberian

including occasional remarks on geological, geomorphological, and archaeological observations. His German geographer and geologist counterpart, Gustav Stratil-Sauer (1894–1975), mainly studied geomorphology of the Lut Desert and Kuh-e Hezār, as well as geomorphology and paleoclimate of Iran with seven publications from 1937 to 1957 (Stratil-Sauer, 1934, 1937a, 1937b, 1941, 1953, 1957; Stratil-Sauer and Weise, 1974). In 1948, Bailey, Jones, and Asfiā (the latter established the Geological Survey of Iran in 1959, discussed later) wrote a paper on the geology of the Alborz Mountains covering a few geological observations and paleontological determinations on the Areh Kuh, Āb ‘Ali, and Lar Valley (Bailey et al., 1948). They pointed out of two phases of compression in the southern Alborz in the Cretaceous and in the late Miocene. The authors criticized Rivière’s work, but they were not correct in their criticism. 8.2. The 1950–1980 Period During the second half of the twentieth century, more attention was paid to the geological exploration in Iran, mostly by European geologists. After the establishment of the IOC (later NIOC) in 1950 (to explore eight regions of Iran), and the Geological Survey of Iran in 1962 (to systematically map the country outside the Zāgros), the Iranian geologists were trained by the European geologists and started working in different parts of the country under their supervision. In 1951, the Alborz Foundation of the Presbyterian Mission invited the Swiss geologist Eugène Hubert Rieben (1889–1972) to give lectures on geology to the students of the University of Tehran. In spring 1952, together with Dr. J.W. Schröder of University of Geneva (invited by the Food and Agriculture Organization of the United Nations [FAO]), they started studying the groundwater and water-bearing alluvial deposits of Iran. In 1953, Rieben was invited by Dr. Ahmad Hossein ‘Adl (1889–1962), head of the Seven Years Plan, Plan and Budget Organization of the Imperial Government of Iran, to focus on the Tehran alluvial deposits for groundwater evaluation. In 1956, at the request of the Irrigation Department of the Ministry of Water and Power, the FAO offered Rieben the opportunity to study the groundwater aquifers of northern Iran, which lasted to 1960. Rieben for the first time divided the alluvial deposits of the Tehran basin into four different stratigraphic formations of Hezārdarreh (A), Kahrizak (B), Tehran (C), and the Recent deposits (D) (Rieben, 1953a, 1953b, 1955, 1966). Rieben’s work was later followed and completed by H. Laurent (Laurent, 1967, for the Karaj plain) and M. Engalenc (1968, for the Tehran plain). In 1956, British Petroleum Co. published geological maps and numerous cross sections of the Zāgros Mountains of southwest Iran. Later, in 1964, the explanatory notes for the maps with columns and cross sections were published (British Petroleum Co., Ltd., 1956a, 1956b, 1964a, 1964b). Late in 1959, Dr. André Vatan of the French Petroleum Institute (Institut Français du Pétrole [IFP]) was invited to Iran to supervise the Iranian students in the field geology course at the

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084715/spe525-03.pdf by University of Wisconsin Madison user

University of Tehran. Vatan had a yearly program to take the geology students of the Faculty of Sciences and mining students of the Faculty of Engineering for field work in the Qom basin area, south of Tehran. He wrote an internal geological guide report of the Qom area with some maps and cross sections describing stratigraphy and tectonics of the region (Guide Géologique de la Regions de Qum; mimeograph copy, University of Tehran). André Vatan continued his work toward the end of the 1960s (when I was student, an older Vatan took our class in 1967 for field study). Vatan spoke in French and he had a French-Persian translator with him in the class and during field study. In 1959–1960 and in 1962, respectively, Italconsult and the geologists of the General Italian Oil Company (Aziende Generale Italiana, AGIP) and (Société Irano-Italienne des Pétroles, SIRIP) carried out photogeological and field mapping of the coastal Makrān, SE Iran (Fig. 1), for assessment of hydrocarbon potential, along with preparation of excellent quality maps and reports (Italconsult, 1959–1960; AGIP/SIRIP, 1962). In 1965, Gerry S. James (who later became the chief exploration geologist of the Iranian Oil Operating Companies) and John G. Wynd published their comprehensive classic work on the stratigraphic nomenclature of the Iranian Oil Consortium Agreement Area in the Zāgros fold-and-thrust belt of southwestern Iran (Fig. 1), with correlation charts of the Triassic to Pliocene– Pleistocene sediments of the Zāgros, Iraq, Kuwait, and Saudi Arabia (James and Wynd, 1965). 8.2.1. 1950–1961 IOC/NIOC Period: The Era of Heinrich Martin Huber (1950–1979), Augusto Gansser (1955–1962), and Jovan Stöcklin (1950–1961) In 1949, the Iran Oil Company (IOC) was founded by the Seven-Year-Plan Organization of the Imperial Government of Iran to explore eight regions of Iran. At this time, ~80% of the country, excluding the British Oil Concession area in the Zāgros, was geologically unknown. The Government of Iran requested Professor Arnold Heim to assemble a team of eight Swiss geologists to start exploration and training of young Iranian counterparts in the newly created IOC, a forerunner of the NIOC. On 1 February 1950, Drs. Heinrich Huber, Augusto Gansser, Jovan Stöcklin, Pierre Albert Soder (1922–2008; Lehner, 2009), Karl Theodor Goldschmid (1896–1982; Soder, 1983), Ernst Frei, and Max Furrer (with Arnold Heim making up the eight-man Swiss team) joined the newly established IOC to explore for oil and map the country that had not been geologically surveyed earlier (Fig. 2). During this time, the Iranian oil industry was still in the hands of the Anglo-Iranian Oil Company (later BP). After nationalization of the Iranian oil industry by Dr. Mosaddeq, Esq., the IOC became eventually integrated into the new NIOC. When Professor Albert Arnold Heim (1882–1965) and the eight Swiss geologists arrived in Tehran as an advance party and began work on 1 February 1950, IOC still had less than a dozen employees, among them Rezā Fakhrā’i and Hushang Tarāz (1923–2012), who were IOC’s first Iranian geologists. Stöcklin noted that in 1950, after IOC’s Director Bāgher Mostofi

Development of geological perceptions and explorations on the Iranian Plateau had introduced Heim to his small Iranian staff, he pointed to a large wall map of the studied oil-prospective zone of the “British Concession Area” of the Zāgros at 1:1,000,000 scale, showing the country’s geology that was reliably known at that time. This map represented roughly 10% of Iran’s territory, the remaining 90%, disregarding a few widely scattered spots, was blank sheet. Mostofi told the eight Swiss geologists that “it is now your task to fill this map!” (Jovan Stöcklin, 30 May 2003, personal commun. from Seuzach, Switzerland). Sure enough, Dr. Stöcklin accomplished that mammoth task during his 27 yr of working and living in Iran as the chief geologist both at the IOC/NIOC and later as director of the Geological Department of the Geological Survey of Iran (GSI). Arnold Heim served as chief geologist only from 1950 to 1952 in Iran. Jovan Stöcklin (1921–2008), the father of the modern and systematic geological investigation in Iran, lived and worked for 27 yr in Iran; beginning in 1957, his wife Elisabeth shared 20 yr in Iran, and their four children all were born and grew up in Tehran. At first, the exploration was conducted under the supervision of Swiss geologists led by professor Arnold Heim (1882–1965),

from 1952 by Dr. Augusto Gansser (1910–2012), and later by Dr. Karl Theodor Goldschmid (1896–1982) as chief geologists. In 1960, the work was gradually taken over by the first generation of Iranian geologists at the NIOC, and soon Huber remained the only Swiss, highly respected for his commitment and dedication at the NIOC. The “first generation geologists,” including Rezā Kalhor, Gholāmrezā Mohājer, Yusef Paran, ‘Amānollāh* Ja’fari, Rezā Fakhrā’i*, Parviz Minā, Mohammad Taghi Razāghniā, Bāgher Mostofi, Akbar Dibāj, and others, conducted reconnaissance surveys of primarily Central and East Iran, under supervision of the Swiss geologists. Only one month after the arrival of the Swiss geologists, on 1 March 1950, two field parties were ready to start on the IOC’s first expedition to Baluchestān, SW Iran, assisted by newly employed drivers, cooks, camp-workers, and provisionally equipped with vehicles, camping outfits, food supplies, personal field clothing, etc., all rapidly selected in the old Bazaar in *Errata: In the 23 Nov. 2016 online version, ‘Amānollāh was given as ‘Amrollāh and Fakhrā’i was given as Fakhāri.

Figure 2. Invaluable historic photograph of Bāgher Mostofi (managing director, IOC; middle) with the eight Swiss geologists arriving in Tehran, February 1950, to start working at the Iranian Oil Company (IOC, later National Iranian Oil Company [NIOC]). From left to right: Pierre Albert Soder (1922–2008), Augusto Gansser (1910–2012), Heinrich Martin Huber (1917–1992), Arnold Albert Heim (1882–1965), Jovan Stöcklin (1921–2008), Karl Theodor Goldschmidt (1896–1982), Ernst Frei, and Max Furrer. Courtesy of Milena Pika-Biolzi (ETH, Zürich, Switzerland, 30 July 2014).

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084715/spe525-03.pdf by University of Wisconsin Madison user

51

52

M. Berberian

downtown Tehran (Fig. 3). The few roads passable by car at the time were bumpy dust trails, and it took 16 days for the party to reach Chāh Bahār port in Baluchestān, covering a distance of ~1750 km (Fig. 4). It took a full day alone for the field party to work their way with pick and shovel through the 11 km of Tang-e Sarteh south of Bampur. On two expeditions totaling five months, the entire Makrān coastal range between Jāsk and the Pakistan border (a distance of 290 km) was reconnoitered, with all transportation done on camelback (~10 camels to carry the field facilities and food, and local Baluchi camel drivers and guides, and gendarmes for security); each party had to carry ~40 canisters of petrol on board since there was no gasoline station in the area. For food, the party depended largely on local dates, camel milk, and fish from the

Gulf of Oman. One driver became seriously ill with malaria and was sent back to Tehran. The coastal Makrān ranges to the Pakistani border were surveyed during 45 consecutive days. It took the party a whole week in a forced day-and-night march to return to Jāsk in the western Makrān, 1237 km straight line from SE Tehran (Jovan Stöcklin, 30 May 2003, personal commun.). Early reconnaissance work by IOC led soon to the selection of two Neogene basins for detailed surface exploration and drilling for petroleum: the Qom Basin in northern Central Iran, south of Tehran, and the Māzandarān Basin on the Caspian coast of northern Iran. Stöcklin in cooperation with one or two Iranian colleagues prepared the first detailed plane-table surveys of the Alborz and Sarājeh (Qom Basin) anticlines (Fig. 5), carried out semidetailed mapping of the central and eastern parts of the

Figure 3. Historic photograph of the Iranian Oil Company (IOC) geologists with their cars getting ready to start the Baluchestān expedition on 1 March 1950, Tehran. From left to right: Karl Theodor Goldschmidt (1896–1982); Dr. Kayhān; Fatollāh Naficy (1908–2002; managing director of Sherkat Sahāmi Naft-e Melli, the forefather of the National Iranian Oil Company [NIOC] who became a member of the board of directors of the NIOC); unknown; Arnold Albert Heim (1882–1965); Bāgher Mostofi; Elisabeth (?) and Jovan Stöcklin (1921–2008); Heinrich Martin Huber (1917–1992); and Hushang Tarāz (1923–2012). Courtesy of Milena Pika-Biolzi (ETH, Zürich, Switzerland, 30 July 2014). Hormoz Naficy (London) kindly identified his father and Dr. Kayhān in the photograph.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084715/spe525-03.pdf by University of Wisconsin Madison user

Development of geological perceptions and explorations on the Iranian Plateau

53

Figure 4. Dr. Jovan Stöcklin at his first Iranian Oil Company (later National Iranian Oil Company) field trip in Iran, coastal Makrān between Chāhbahār and Gwādar (Gavāter), in March 1950. For the caption of this picture Stöcklin wrote: “Me with my camel; I am the one in the foreground!” Courtesy of Jovan Stöcklin (30 May 2003, personal commun., Seuzach, Switzerland).

Figure 5. Historic photograph of the Iranian Oil Company’s Alborz-1 oil well location (ground stones laid down) in mid-1950s in the Qom basin of Central Iran. From left to right: Heinrich Martin Huber, Bāgher Mostofi, Reza Fakhrā’i*, Hushang Tarāz, ? Freeman, Augusto Gansser, Hormoz Nafici, and Jovan Stöcklin. Courtesy of Jovan Stöcklin (30 May 2003, personal commun., Seuzach, Switzerland) and Milena Pika-Biolzi (ETH, Zürich, Switzerland, 30 July 2014). *Erratum: In the 23 Nov. 2016 online version, Fakhrā’i was given as Fakhāri.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084715/spe525-03.pdf by University of Wisconsin Madison user

54

M. Berberian

wooded Māzandarān Basin, and participated in well-site geological characterization of the first bore holes. On 26 August 1956, the uncontrolled oil at well No. 5 in the Alborz oil field north of Qom gushed to the height of 53 m at a rate of 120,000 barrel (19,000 m3)/d and caught fire. The oil gusher was controlled after 90 d of work by Myron Kinley of the United States with the cooperation of Bāgher Mostofi of NIOC. This was the largest environmental pollution by petroleum products in Iran, which was never remediated. Nearly all of the work in the dense jungles of Māzandarān in northern Iran was carried out on foot and on horseback, including a traverse of the Alborz Mountains between Nekā and Dāmghān (100 km straight line) in 1954. At this time, reliable topographic maps or aerial photographs were not available. Stöcklin had to prepare his own road map, using a compass, pocket altimeter, and a wrist watch for converting marching or riding time into distance. Other mapping covered the mountains between Qom and Esfahān, where the party vehicles sometimes appeared in places where people, having never seen an automobile, ran away in panic. On a special mission in 1953, Huber and Stöcklin, assisted by Rezā Fakhrā’i and commander Dāneshvar (later director of the Geographic Department of the Iranian Army) as surveyor, did the first exploration and reserve estimate of the Hodjedk coal deposits (ESE of Zarand in Kermān Province, SE Iran), which later served as a starting point for the large team from Soviet Union, including imbedded Komitet Gosudarstvennoy Bezopasnosti (KGB) agents, who developed the important Hojedk coal mining complex supplying the Iranian steel industry of Esfahān (Jovan Stöcklin, 30 May 2003, personal commun.). After the old Anglo-Iranian oil field in Naft-e Shāh had become the exclusive domain of NIOC for the local oil consumption of Iran, Gholam ‘Ali Mohājer and Jovan Stöcklin became, in 1956, the first NIOC geologists to resume exploration in this area, extending it on horseback as far southeast as Ilām, SE Iran. In 1955, in cooperation with Rezā Kalhor (a later exploration and production director of NIOC), they surveyed much of the northern, eastern, and southern frame of the Great Kavir (desert), what seems to have been the first motorized complete traverse of the Kavir, using a single Jeep with an experienced driver and an old caravan leader guiding along a camel path from Hosseinān to Jandaq across this most inhospitable Great Kavir (desert) in north Central Iran. A reconnaissance of the northwestern Lut Desert between Tabas, Behābād, and Rāvar in SE Iran was carried out in 1960 with ‘Ali Moshtāghiān (later NIOC director), including a 200 km traverse of the hitherto completely unknown, lifeless, and waterless desert stretch between Parvadeh and Behābād, using four camels for transport of a minimum of equipment and food but a maximum water supply, accompanied by a camel driver and a local guide, and during 10 d not encountering a single house, tree, or another human. Among many surprises, this survey led to discovery of the first pre-Tertiary salt domes in Central Iran and the identification of their salt source beds with the Hormoz Salt

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084715/spe525-03.pdf by University of Wisconsin Madison user

complex and the Upper Triassic evaporites of the Persian Gulf region. This had a considerable impact on the paleogeographic reconstructions of Iran in the light of the new plate-tectonic concepts that came into being at just about that same time (Jovan Stöcklin, 30 May 2003, personal commun.). Dr. Heinrich Martin Huber (1917–1992; Fig. 6), who worked at the Swiss Federal Institute of Technology (ETH), University of Basel, Switzerland, and the Texas Oil Company (Texaco, in Colombia), continued his work in the IOC and NIOC as the head of the Surface Geology Department until 1978, just before the regime change, when he left the country. With his wife Dorli Walser and their two daughters, they lived in the Shemirān affluent quarter in northern Tehran (Marianne Huber Glunez and Eva Glunez, 6 June 2003, personal commun.; Soder, 1992). Immediately upon arrival in Iran, Dr. Huber traveled to northern Makrān, a first taste of many journeys to come through the desert, by jeep, on a camel, or donkey back, in insufferable heat or chilling winds. Makrān also was to remain his last area of work, for Paragon Contech in the late 1970s, after returning to Switzerland. For IOC and NIOC, Huber worked in many different regions across the vast country. Huber contributed significantly to exploration of coal in Kermān as well as groundwater investigation, and to engineering issues, and made his expertise in petrography available to his fellow geologists. With his colleague Gansser, they went on several mountaineering expeditions in their spare times (i.e., the ‘Alamkuh Range among others; Marianne Huber Glunez and Eva Glunez, 6 June 2003, personal commun.; Soder, 1992). Huber prepared the first Geological Map of Iran on a scale of 1:2,500,000 in 1960 (Huber, 1960), and the six sheets of the Geological Map of Iran on a scale of 1:1,000,000 with

Figure 6. Dr. Heinrich Martin Huber at the Kāl-e Reshm tent camp in November 1956 in the Great Kavir (Desert) of Central Iran. Courtesy of Marianne Huber Gluenz and Eva Gluenz (London, 6 June 2003, personal commun.).

Development of geological perceptions and explorations on the Iranian Plateau explanatory notes published by the NIOC in 1978 (Huber, 1978). Jovan Stöcklin, in admiring Huber’s work in Iran and his dedication to the country, mentioned that “Huber worked like a horse in the field and office!” (Jovan Stöcklin, 30 May 2003, personal communication) (Fig. 7). Although Augusto Gansser’s main work and interest was the Himalaya (Sorkhabi, 2012), in 1951 he joined H. Huber and A. Heim working for the IOC. Gansser (1910–2012) investigated the general stratigraphy of Iran for the NIOC and systematically mapped a portion of Central Iran (Fig. 8; Gansser, 1955). He was apparently one of the first geologist who mapped the Great Kavir and Lut Deserts for the then-IOC. Gansser coined the term “Colored Mélange” describing the chaotic mixture of ophiolite-radiolarite belts (Gansser, 1955), and later he distinguished the mélange from ophiolite zones without mélange character (Gansser, 1960). The term has been widely used in Iran by almost all the geologists since then and has been erroneously labeled as a stratigraphic unit on almost all the maps and reports. Gansser was also involved in the discovery of the Alborz oil field near Qom in Central Iran (mentioned earlier); the well hit the oil trap on 26 August 1956 (Mostofi and Gansser, 1957; Sorkhabi, 2012). Together with Huber, they published a paper on geological observations in the central Alborz, describing detailed stratigraphy of the ‘Alamkuh region and introducing the Precambrian basement and the overlying Cambrian–Ordovician deposits (Gansser and Huber, 1962).

Figure 7. Dr. Heinrich Martin Huber after retirement (not dated) in Switzerland. Courtesy of Marianne Huber Gluenz and Eva Gluenz (London, 6 June 2003, personal commun.).

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084715/spe525-03.pdf by University of Wisconsin Madison user

55

8.2.2. 1962–1979 NIOC Period During the second half of the twentieth century, the “second” and “third” generations of Iranian geologists worked at the NIOC and the Iranian Operating Companies (IOOC) in Central Iran, Kopeh Dāgh (NE), Alborz and Caspian (N), Moghān (NW), and Zāgros (S, Fig. 1). The “second generation” of geologists included: ‘Abbās Afshār-Harb, Bizhan Allāhyār, H. Amini, Homāyun Ansāri, Sāleh Banāfti, Fatollāh Bozorgniā, Heshmat Bozorgniā, Manuchehr Fotuhi, ‘Ali Golestāneh, Amir Kalāntari, ‘Emād Kavāri, Ahmad Kheradpir, Siāvosh Mohāfez, ‘Ali Moshtāghian, Mehdi Pārsi, Hassan Modarressi, ‘Atā Mogharebi, Iraj Sālehi, and Parviz Shahidi, among others. Some of these geologists conducted independent mapping projects in regions primarily outside of the Zāgros concession area, and some did subsurface work, including planning and supervising of exploration and production drilling. The “third-generation” geologists included: Siāmak Āgāh, Vahid Amjadi, ‘Ali Ārtinmehr, Ardeshir Āzarpay, Ninus* Benyāmin, Gholāmrezā Dashti, Hassan Kāveh, Fereidun Mālekiān, Gholām ‘Ali Mohājer, Majid Mohājer, Siāvosh Najmābādi, Hamid Nārāni, Akbar Rahaqi, ‘Atāollāh Setudehniā, Mohammad Fakhāri, Manuchehr Takin (through *Errata: In the 23 Nov. 2016 online version, Ninus was given as Nimous, and Parviz Shahidi was included in this list of geologists.

Figure 8. Dr. Augusto Gansser (1910–2102). Courtesy of Milena Pika-Biolzi (ETH, Zürich, Switzerland, 30 July 2014).

56

M. Berberian

the Consortium), Iraj Yāsini, and more. At the beginning, they assisted either the second-generation geologists or Dr. Heinrich Martin Huber (1917–1992), head of the Surface Geology Department of NIOC, conducting field mapping; later, some became party leaders (Siāmak Āgāh, ‘Emād Kavāri, and Manuchehr Takin, 3–13 August 2015, personal commun.). 8.2.3. The GSI (1962–1978) and Era of Dr. Jovan Stöcklin (1962–1975) In early 1962, the Imperial Government of Iran established the Geological Survey of Iran (GSI) for systematic study of geology and mineral resources under the United Nations (UN) Development Programme (UNDP Special Fund Project) and with the hard work of Dr. Safi Asfiā (1916–1999) and Dr. Nasrollāh Khādem (1910–1999), both graduates of the 1938 École des Mines de Paris, France, members of a large group sent by Rezā Shāh Pahlavi for graduate studies to Europe for development and modernization of Iran. The UN project manager of the GSI was Dr. David A. Andrews from the USGS (Fig. 9), who served from 1962 to 1968. Dr. Andrews was a very skilled person, especially in dealing with the Iranian government authorities in establishing and running the GSI. He was a nice person and greatly admired by all members of the GSI. The UN contributed US$1,425,300.00, and the Iranian government contribution was US$2,079,000.00 for the establishment of the GSI (Berberian, 1997). After implementation of the law that established the GSI, Dr. Asfiā became the nominal director of the GSI from 1959 to 1961; however, the office was mostly directed by Dr. Khādem, who became the managing director of the GSI from 1962 to 1974. Both Asfiā and Khādem also taught at the University of Tehran as adjunct professors. Dr. Asfiā then became the consulting minister and director of the Plan and Budget Organization of the Imperial Government of Iran. After the regime change in 1979, Dr. Asfiā was imprisoned by the revolutionary court for

4 yr (1979–1983); Dr. Khādem was summoned to the court, and his retirement pension was cut drastically by demoting his rank (both were accused of having positions in the Imperial Government); Dr. Tarāz’s house and bank accounts were confiscated and he fled to the United States, where he passed away. When Dr. Khādem started his duties at the GSI, except for the Zāgros and the areas covered by IOC/NIOC, there was not much geological information on Iran, and he had to start from scratch. It was under the directorship of Dr. Khādem and with the help of the UN geologists, especially Dr. Jovan Stöcklin, that the GSI employed young Iranian valedictorian geology graduates to be trained and was able to substantially increase the geological knowledge of Iran. As a member of the second generation of young Iranian geologists, and the fourth Armenian ethnic-religious minority member, Dr. Khādem employed me at the Geological Survey of Iran in April 1971 without conducting an interview; however, the strict selection criteria were relaxed when Dr. Khādem retired in 1974. Jovan Stöcklin joined the UNDP and was appointed the chief geologist and director of the Geological Department (1962–1975) in this project. His main responsibilities were systematic geological investigation and mapping of the country, coordination and supervision of all the geological activities of the GSI, from field work to publication of the maps and reports, and field training of the first-generation young Iranian geology graduates hired by the GSI (Jamshid Eftekhārnezhād, ‘Abdolrahim Hushmandzādeh, Viguen Issākhāniān (1932– 2016), Mohammad Hassan Nabavi, Hārun Nicolās, Mansur Samimi-Namin, Shāhen Tātāvussiān, Nāser Vāleh, Mostafā Zāhedi (1933–2005), and many more). Soon (in 1963), Dr. Anton Wolfgang Ruttner (1911–2006; Cernajsek, 2007, 2009), director of the Geological Survey of Austria, joined Stöcklin as a UNDP coworker to establish the Geological Department of the GSI. Ruttner (Fig. 10) had a very close relation with Iranian

Figure 9. Dr. David A. Andrews (U.S. Geological Survey), the United Nations Project Manager at the Geological Survey of Iran (1962–1968), with Dr. Hushang Tarāz, director of the Geological Department of the Geological Survey of Iran (1975–1978), at the Economic Commission for Asia and the Far East (ECAFE) Conference in Bangkok, Thailand, 1973. Courtesy of Dr. Hushang Tarāz (20 May 2009, personal commun., San Diego, California).

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084715/spe525-03.pdf by University of Wisconsin Madison user

Development of geological perceptions and explorations on the Iranian Plateau geology since 1956, when he arrived in Iran and started working at the Mināk Mining Company in the Tuyeh and Ozbak Kuh areas in Central Iran and the Āqdarband coal mine in the Kopeh Dāgh in northeast Iran (Fig. 1). He studied the Cambrian rocks of east Central Iran and mapped the Shirgesht and Ozbak Kuh areas, as well as training young Iranian geologists (Mohammad Hasan Nabavi, Hushang Tarāz, Mansur ‘AlaviNā’ini, Javād Hājiān), and he wrote a book on the geology of the Āqdarband area (Ruttner, 1991). Ruttner also studied the geology along the future track of the Iran-Turkey railroad in the Qotur Valley, near the northwest Iranian border. Ruttner was extremely nice and friendly with everybody; he always took the same cook for his field party (Shir Āghā) and even helped him when he was ill, when he took Shir Āghā to Austria and paid the expenses. During phase 1 of the GSI development (1962–1968), the GSI recruited over 50 young Iranian geologists, chemists, cartographers, and supporting personnel, while the UN supplied all field and laboratory equipment, as well as 18 international experts employed for periods between 1 and 6 yr, whose task it was to train the young Iranian geologists in their special fields of field mapping, paleontology, petrography, mineral resources, and cartography. About half of the Iranian technical/scientific staff and eight of the UN experts were assigned to the Geology Department of Jovan Stöcklin.

Figure 10. Dr. Anton Wolfgang Ruttner (1911–2006). Courtesy of Thomas Hofmann (Geologische Bundesanstalt, Wien, Austria, 30 July 2014).

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084715/spe525-03.pdf by University of Wisconsin Madison user

57

Field parties were organized as independently operating teams of two to three geologists with support personnel (a cook, two drivers, locally hired guides, etc.) and camping facilities. In the beginning, the party chief was always a UN expert, and later a sufficiently trained and experienced Iranian geologist. Additional assistance was received from several European universities, who sent doctoral candidates to Iran and in exchange accepted Iranian candidates, all contributing with their doctoral dissertations on Iranian themes to the GSI’s scientific and mapping projects. Under an associated bilateral project, the Swiss government detached two experts from its Federal Cartographic Service and provided all engraving and photomechanical equipment necessary for the cartographic preparation of the GSI maps for offset color printing. A partial target of the Geology Department in phase 1 was initiating the publication of GSI reports with 1:100,000 scale maps of selected areas of special interest, and of a systematic 1:250,000 scale geological quadrangle map series with explanatory texts (Jovan Stöcklin, 30 May 2003, personal commun.; Berberian, 1997). Since there were no trained and experienced Iranian geologists, Stöcklin invited Dr. Augusto Gansser (who had been recently appointed as professor at Zurich ETH, Switzerland) and Alpine research students mostly from: (1) Technische Hochschule, University of Zurich, Switzerland (Jean Dellenbach, 1964, from Strasbourg, France; Christoph Lorenz, 1964; H. Flügel, 1964; Martin Glaus, 1965; Peter Allenbach, 1966; René Steiger, 1966; Eduard Dedual, 1967; Stephan Paul Meyer, 1967; P. Stalder, 1971; E.G. Cartier, 1971; Peter Ernst Süssli, 1974); (2) Istituto di Geologia dell’Università di Milano, Italy (Riccardo Assereto, 1963, 1966a, 1966b; Nerina Fantini Sestini, 1965a, 1965b, 1965c, 1966a, 1966b; Fantini Sestini and Glaus, 1966; Maurizio Gaetani, 1964, 1965, 1968); (3) Bundesanstalt für Bodenforschung, Germany (Reinhold Huckriede, Martin Kürsten, and Helmut Venzlaff, 1962; Kermān region), and many more, to help in mapping Iran, especially the central Alborz Mountains near Tehran. Furthermore, Otto Thiele (Austria, 1963–1965), S. Iwao (Tokyo, Japan, 1963–1965), and K. Hirayama (Tokyo, Japan, 1962–1965) started mapping and training the young Iranian geologists at the GSI in the Golpāyegān, Central Alborz, the Soltānieh range at Tārom, and the Angurān zinc-mine areas. In early 1963, Professor Dimitrij L. Stepanov, the UN paleontologist from Leningrad University, USSR, who specialized in Permian macrofossils, joined the GSI. Stepanov was a very knowledgeable scientist and an expert on the Djulfian strata (Upper Permian) of Armenia. He studied and dated numerous samples and strata, and trained young Iranian paleontologists (Farokh Golshani). Stepanov worked for 5 yr at the GSI and left the country in 1968. In 1969 and 1970, Stepanov published three reports on the Paleozoic in general, and the Carboniferous and Permian–Triassic (with Golshani and Stöcklin) stratigraphy of Iran. In late 1970s, Stepanov was chosen as the first chairman of the Subcommittee on Permian Stratigraphy.

58

M. Berberian

More than 30 foreign geologists from Australia, Austria, Canada, Czechoslovakia, England, Finland, France, Germany, Japan, New Zealand, Netherlands, Norway, Philippines, Soviet Union, Sweden, Switzerland, and the United States worked and trained the Iranian geologists at the GSI from 1962 to 1973. Riccardo Assereto (1937–1976; Figs. 11 and 12), the student of Professor Count Ardito Desio (1897–2001), whom Desio considered like a son (Maurizio Gaetani, 28 September 2006, personal commun.), made a major contribution to the stratigraphy of Iran during his short postgraduate years in Iran. Assereto together with three other students, Maurizio Gaetani, Ignazio Ippolito, and Domenico (Mimmo) Fornaro, arrived at Tehran by bus in early July 1962. Assereto was one of the brilliant Alpine geologists of the period who introduced the Paleozoic stratigraphy with lithostratigraphic nomenclature of the central Alborz Mountains north of Tehran, which was submitted to, and approved by, the Stratigraphic Names Committee of Iran (Assereto, 1963, 1965, 1966a, 1066b). The nomenclature has been employed in numerous publications since then and is valid today. After receiving his doctorate, Assereto became a professor at Milan University. Prior to the historical and classic work of Assereto (1963), the Lower to Upper Cambrian Soltānieh, Bārut, Zāgun [sic, Zāigun], and Milā Formations were considered as “preDevonian” in age, because the Lower–Middle Paleozoic Lālun Sandstone Formation of Iran was compared to the “Old Red Sandstone” of England and was dated as Devonian (Ovtsinnikov, 1930; Böhne, 1932a; Rivière, 1934b; Bailey et al., 1948; Dellenbach, 1964; among many others). Sadly, Dr. Assereto (Figs. 11, 12) together with his 10-yr-old son Andrea and a professor of regional geology of the University of Bologna, Dr. Giulio Pisa, were killed by a rock avalanche triggered by the 15 September 1976 Ms 6.5 Friuli earthquake in the Carnia Mountains, northeast Italy (Gaetani, 1978; Maurizio Gaetani, 28 September 2006, personal commun.). All the maps prepared in the early days of the GSI were based on aerial photographs of the World-Wide Aerial Survey Corporation (joint project of the Iranian Imperial Army and the U.S. Army) at the scale of ~1:55,000. No sufficiently reliable topographic base maps for the country were available; therefore, the aerial photographs were tied in to special ground triangulations, usually carried out by A. Gerānpay and Hushang Tarāz (1923–2012; Figs. 13 and 14). The number of SHORANcontrolled (short-range navigation) checkpoints supplied by the Geographic Department of the Imperial Army enabled the GSI to adjust the triangulations to the official grid system. For a complete list of the Iranian geologists working at the GSI during this period, see Berberian (1997). The Soltānieh range of Tārom (Stöcklin, Nabavi, Eftekhārnezhād, and Samimi-Namin) resulted in the first identification in north Central Iran of fossiliferous Cambrian carbonates and of a conformably underlying sequence of red clastic and dolomites (the Soltānieh Dolomite) of Early Cambrian– Infracambrian age (Stöcklin et al., 1964). These rock sequences

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084715/spe525-03.pdf by University of Wisconsin Madison user

subsequently were followed through the Alborz Mountains and Central Iran to the Shotori-Shirgesht-Ozbakkuh areas of Tabas (Stöcklin et al., 1965; Ruttner et al., 1968), where also the most complete Paleozoic section known so far in Iran was found, and where the Infracambrian beds link up with their Hormoz-type evaporitic and diapiric equivalents discovered in the northwestern Lut Desert during Stöcklin’s NIOC period. By the end of phase 1 (1968), a dozen official GSI reports and eight colored maps, including the first results of the Mineral Resources Department, were printed in English and published, with many more still in preparation. This material was most useful for initiating a rapidly expanding exchange of publications with geological surveys and universities in neighboring countries and worldwide. After the completion of phase 1, the UN projects concentrated on mineral exploration, and Stöcklin remained the only UN expert in the Geology Department of the GSI (Fig. 15). As coordinator for Iran-Afghanistan-Pakistan for the United Nations Educational, Scientific and Cultural Organization (UNESCO)–sponsored International Tectonic Maps of the Middle East and Asia and the Far East, Stöcklin organized in 1968 an International Tectonic Symposium at the GSI Headquarters in Tehran, with excursions to Central and south Iran. Such international activities, which included cooperation in an Iran-Himalaya Working Group of the International Geodynamics Project, called for Stöcklin’s participation in working sessions and field meetings in Turkey, the Caucasus, Oman, Afghanistan, Tajikistan, the Urals, Siberia, the Himalaya, and other Asian regions. Professor Hollis Dow Hedberg (El Doctor Hedberg; 1903– 1988) of Princeton University invited Jovan Stöcklin to join the International Union of Geological Sciences (IUGS) Subcommission on Stratigraphic Classification and his five-person editorial committee for the first edition of the International Stratigraphic Guide. This in turn led to creation of a national Stratigraphic Names Committee of Iran, which revised the entire stratigraphic nomenclature of the country. Supported by his vast knowledge of Iranian geology since 1950, proficiency in English, French, German, and Russian (he also spoke fluently in Persian), and the irreplaceable archives of the old geological reports, maps, and books of the country at the GSI (which unfortunately was looted in the chaotic years of 1978–1979) and NIOC, Jovan Stöcklin prepared the mammoth stratigraphic lexicon of Iran with 403 formal, informal, and old rock-unit names (excluding the Zāgros) with full descriptions and references for all the entries (Stöcklin, 1972). Together with Jamshid Eftekhārnezhād and ‘Abdolrahim Hushmanzādeh, Stöcklin resumed exploration in the central Lut, which had begun as early as 1963, when this remote desert region was geologically still virtually untouched. It was believed to be formed exclusively of Cenozoic extrusive rocks, but was then revealed to have a high-grade metamorphic nucleus, granite intrusions, lacunary Permian–Mesozoic platform sediments, and even ophiolitic and flysch assemblages. The result of this expedition was published in 1972 with a 1:500,000 scale geological

Development of geological perceptions and explorations on the Iranian Plateau

59

Figure 11. Riccardo Assereto (1939– 1976) with cap (left) and Pierre Allenbach (right) at the tent camp in the Lar valley in August 1966. The location is just down valley of the ChehelCheshmeh Mountain, which was later submerged by the Lar dam reservoir (1974–1981). The two young geologists are discussing the junction zone of their geological maps. Courtesy of Maurizio Gaetani (19 December 2006, personal commun., Milano, Italy). Because of the unfortunate sudden death of Riccardo Assereto at the young age of thirtyseven during the 1976 Friuli earthquake rock avalanche, this historic photograph reminds me of Death Playing Chess, a medieval church painting from the 1480s in the Täby Church fresco by Albertus Pictor (~1440–~1507); and the iconic scene in Ingmar Bergman’s 1957 movie (The Seventh Seal), where the Swedish knight (Antonius Block) challenges Death to a chess match for his life. Unlike Antonius, Riccardo could not evade death in the geological chess game by a geological event long enough to transfer his valuable knowledge and experience to his students.

Figure 12. Riccardo Assereto (1939–1976) on top of the Tizkuh ridge in summer 1963 on the Upper Jurassic–Lower Cretaceous succession beds, with the Damavand volcano in the background right. Courtesy of Maurizio Gaetani (13 September 2006, personal commun., Milano, Italy).

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084715/spe525-03.pdf by University of Wisconsin Madison user

Figure 13. Dr. Hushang Tarāz with the University of Tehran geology students during fieldwork in the 1960s near Tehran. Courtesy of Dr. Hushang Tarāz (20 May 2009, personal commun., San Diego, California).

60

M. Berberian

Figure 15. Dr. Jovan Stöcklin in May 2003 in Seuzach, Switzerland. Courtesy of Jovan Stöcklin (30 May 2003, personal commun.).

Figure 14. Dr. Hushang Tarāz (right) with Dr. Ken-Ichi Ishii (left) at the Permian–Triassic Ābādeh Formation outcrop in SW Central Iran. Courtesy of Dr. Hushang Tarāz (20 May 2009, personal commun., San Diego, California).

map of the Lut Desert covering an area of 320 × 200 km (Stöcklin et al., 1972). In 1974, the National Iranian Committee on Geodynamics was created, composed of representatives of the GSI, NIOC, Iranian universities, the mining industry, and some individuals. The committee organized the Tehran Symposium on the Geodynamics of Southwest Asia during 8–15 September 1975, and the result of the symposium was published in 1975 as a special publication by the GSI (Stöcklin, 1981a). After the unexpected retirement of Dr. Khādem as managing director of the GSI, in 1975 Stöcklin handed over his function as department head to his Iranian counterpart Dr. Hushang Tarāz and assumed the role of a geological adviser, assisting in planning, training, and supervision (Dr. Tarāz concurrently was teaching at the University of Tehran as adjunct professor; Fig. 13). He then joined a UN mineral exploration project in Nepal for a 2 yr term. Stöcklin’s final assignment in Iran was from 1977 to 1979, when he returned to Iran from his Nepal assignment at the request

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084715/spe525-03.pdf by University of Wisconsin Madison user

of the Imperial Government of Iran as a UN geological advisor to the newly established Atomic Energy Organization of Iran (AEOI) to organize and supervise within the AEOI’s Exploration Department at Regional Geology Section, which was tasked to define potential uranium-prospective areas that deserved more detailed radiometric, geochemical, and drilling follow-up investigations by respective specialists. Several field parties were organized after the GSI model. His activities continued until autumn of 1978, when well-organized growing opposition and strikes by the Iranian staff of AEOI merged with the organized opposition movement that rapidly spread throughout the country and ended up in a complete standstill in all government agencies and in numerous private enterprises. On 6 February 1979, after the United Nations decided to suspend all their Iranian projects and evacuate all their non-Iranian personnel, Dr. Jovan Stöcklin left Iran for good, the country that had become a second homeland for him and his family. In a personal communication dated 30 May 2003, Jovan Stöcklin wrote me that: “National pride, overwhelming hospitality, politeness and civility in human relations—the characteristics of the inheritors of an ancient culture—are the dominant impression left on me from the Iranian people.” For his contributions to the geological exploration of Iran and the Himalayas, the German Society for Geosciences awarded Dr. Jovan Stöcklin its Leopold-von-Buch Medal; he was an

Development of geological perceptions and explorations on the Iranian Plateau Honorary Member of the Société Geologique de France, German Geological Society, and the Geological Society of Nepal. Stöcklin published over 30 reports and numerous geological maps of Iran while working at the GSI, and many geological internal reports for the IOC/NIOC. He passed away on 30 April 2008 just after submission of his paper, which he presented as a keynote address to the Fifth Nepal Geological Congress (Stöcklin, 2006; Steingruber Stöcklin, 2011; Jovan Stöcklin, 20 December 2001, 5 January 2003, 29 and 30 May 2003, personal commun., Seuzach, Switzerland). Regrettably, the NIOC, GSI, AEOI, and the Geological Societies and universities of Iran did not honor his genuine lifetime dedication to the progress of the country, exploration, and educating the first generation of the Iranian graduates from 1950 to 1978. In 1976, the Shah’s government with the increased petroleum price at the time, started an extensive geological mapping and mineral exploration project to map the remote and geologically unknown east and southeast Iran, which had not been systematically surveyed (the East-Iran Project). Therefore, the GSI was instructed to conceive a large project to source out the mapping of east and southeast Iran to foreign companies with Iranian partners; none of the Iranian partners was a geologic mapping firm, and they did not provide geologists to map the area. The East-Iran Project was therefore launched under the general direction and supervision of the GSI. Firms and scientists from France, Canada, and Australia became involved in this comprehensive project: (1) Mapping of the northern East Iranian Ranges, north of latitude 32°N, was granted to the Bureau de Recherches Géologiques et Minières (BRGM) and the Iranian partner of Geometal (Tehran). One of the French geologists was killed at the oasis city of Tabas-e Golshan during the 16 September 1978 Mw 7.3 Tabas-e Golshan earthquake. (2) A Canadian group of geologists, assembled by Watts, Griffis, and McOuat, Ltd., from Toronto, Canada, working in partnership with Āb va Khāk Consulting Engineers of Tehran, obtained the contract for mapping eastern Iran south of latitude 32°N in Sistān. The project was executed by a Canadian Consortium consisting of Wright Engineers, Ltd. (Vancouver), Bondar-Clegg and Co., Ltd. (Ottawa), by Watts, Griffis, and McOuat Ltd. (Toronto) with Āb va Khāk (Tehran). A field crew of six Canadian geologists mapped an area of ~30,000 km2 of Sistān in east Iran and introduced the Sistān suture zone into the understanding of Iranian geology. (3) Finally, the Makrān area of southeast Iran (Area No. 1; mostly south of latitude 27.30°N, and between longitudes 57°N and 62.50°N) was granted to Contech Pth., Ltd. (Australia), working in partnership with Paragon Consulting Engineers (Tehran). A crew of 25 field geologists, four field petrologists, and two field micropaleontologists, utilizing helicopters, mapped six 1:250,000 scale geological quadrangles (150 × 100 km) and twentyone 1:100,000 sheets (50 × 33 km). Dr. Gerald Joseph

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084715/spe525-03.pdf by University of Wisconsin Madison user

61

Home McCall, fellow of the Geological Society of London, worked as field and office supervisor for Contech. (In 2011, Joe McCall received the Distinguished Service Medal of the Geological Society of London.) In total, 19 geological quadrangle maps on a scale of 1:250,000 (with selected 30 maps on a scale of 1:100,000) covering an area of 330,000 km2 (20% of the country) in east and southeast Iran were carefully mapped by the Canadian, French, Australian, and British geologists in only 3 yr from 1976 to 1978. By late 1978, the greater part of the field work had been completed when the rapidly deteriorating political and social structure in Iran and related financial difficulties led to complete deadlock in the project, and all employees of the foreign companies left the country (Stöcklin, 1981; Joseph McCall, 6 October 2006, personal commun.). As a result of the 1978–1979 turmoil in Iran, an Islamic regime replaced the Shah’s monarchic government; this was followed by the 1980–1988 Iran-Iraq War and long-term closure of the country, which imposed a severe drawback on scientific research in earth sciences. All the international contacts were cut off, and no technical journals or books were purchased from abroad for nearly a decade. 9. TECTONIC CONTEXT The present political boundaries of Iran cover an area of ~1,648,000 km2 in the center of the larger Iranian Plateau, with various tectonic and topographic features of active mountain belts resulting from a juvenile continental collision with intense seismic activity, various suture zones, quasi-rigid blocks, active volcanoes, and a subduction zone. Several major structural and mineral provinces on the Iranian Plateau are characterized by different structural history, style, and evolution (Fig. 1). The crusts of diverse structural provinces, ranging in age from the late Neoproterozoic to Tertiary, have undergone numerous collisional orogenies, with the final orogeny during the Neogene subsequent to the complete closure of the Neo-Tethyan Ocean and its branches during late Paleogene–early Neogene (Berberian, 2014, p. 151–171). Most of the principal data on the regional Iranian geology and tectonics prior to 1990 were covered in classical landmark reports prepared by James and Wynd (1965), Stöcklin (1968a, 1968b, 1972, 1974a, 1974b, 1977), Falcon (1967), Ricou (1971), Braud (1978, 1988), Huber (1978), Berberian and King (1981), Berberian and Berberian (1981), F. Berberian et al. (1982), Berberian (1983a, 1983b, 1983c, 1984, 1989), Tirrul et al. (1983), McCall (1985), and references therein. Most of the previous work has concentrated on relatively distinct regions within the Iranian Plateau, dealing with a certain geologic period, and/or was mainly based on relative geological correlation with little or no radiometric age constraints. Interpretation of the tectono-magmatic and metamorphic evolution of Iran during this period was poorly constrained due to little or no robust geochronological and geochemical data; this was later improved during the twenty-first century through cooperation

62

M. Berberian

by foreign institutions and individuals, mainly by the graduate students working abroad. 9.1. Revised Classical Context Following Baier (1938), Clapp (1940), and others mentioned earlier, Holmes (1944) showed five tectonic units in Iran (composed mainly of two marginal fold belts in the north and south and a median central Iran) as the: Arabian foreland (alluvial plain of Mesopotamia); folded and faulted overthrust SW border ranges (Zāgros); median geosyncline (Iranian Basin); folded and overthrust NE border ranges (Alborz); and Asiatic foreland (Turkmen Steppe). Ābdāliān (1962) published a tectonic map of Iran on a scale of 1:4,000,000 dividing the country into eight structural zones (Fig. 16): (1) Arabian-Syrian Shield; (2) marginal folds (Zāgros and Kopeh Dāgh); (3) Paleozoic cores (between the marginal folds and Iranid); (4) Iranid (including Makrān flysch, radiolarite-ophiolite, and Hamédān zones); (5) Urumieh-Sarāvān volcanic belt; (6) Central Iran; (7) Alborz; and (8) Russian Platform. This was followed by Harrison (1968), who divided the country into eight units: the Alborz System (including the Kopeh Dāgh); Alborz foredeep; the Tabas wedge (including Anārak and Kermān); East Iranian quadrangle (covering the Lut Desert, southern Khorāsān, and Sistān to the Afghan-Pakistan border line); volcanic belt (Karkas in Fig. 1); complex belt (Sirjān in Fig. 1); zone of normal folding (Zāgros); and Makrān (Fig. 1). Despite the data that have become available since 1968, Harrison’s 1968 tectonic units of Iran were used by Eckart Ehlers (2001). Based on his 18 yr of continuous work in Iran at IOC/ NIOC and GSI, Stöcklin (1968a) in his landmark paper thoroughly summarized the new Iranian geological data and their interpretations. He emphasized that the traditional concept of Iran as a pair of orogenic belts (the Alborz in the north and the Zāgros in the south), with a “median mass” between, was not correct, and Central Iran cannot be regarded as a eugeosyncline. He noticed that Iran was a relatively stable platform during the Paleozoic, and the Iranian platform was an extension northward of that of Arabia. During the Mesozoic and Cenozoic Eras, an Alpine-type orogeny affected all of Iran, but typical geosynclinal development took place only in the border regions, in the Zāgros, Kopeh Dāgh, Makrān, and the eastern Iranian ranges along the Afghan frontier (Fig. 1). Central Iran, though equally mobile, remained essentially epicontinental or continental. This was followed by fracturing of the Iranian platform during the Triassic along what is known as the Zāgros thrust zone, along the northern Zāgros Mountains. Stöcklin (1968a), in a review of the structural history and tectonics of Iran, divided the country into two main domains: (1) stable areas (Turān, Arabia, Lut) and (2) folded areas (Kopeh Dāgh, Alborz, Central Iran, Sanandaj-Sirjān, Zāgros thrust zone, Zāgros folded belt, and East Iran-Makrān). Based on differences in their structural history and tectonic style, Stöcklin (1968a) distinguished the following nine major structural zones in Iran:

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084715/spe525-03.pdf by University of Wisconsin Madison user

(1) The Plain of Shatt al-Arab (Arvandrud); (2) folded belt of Zāgros; (3) Zāgros thrust zone; (4) Sanandaj-Sirjān Ranges; (5) Central Iran sensu stricto (comprising the whole area between the northern and southern Iranian ranges); (6) Alborz Mountains; (7) Kopeh Dāgh Range; (8) Lut block; and (9) East Iran and Makrān Ranges (Fig. 1). Stöcklin’s interpretations (1968a) were based on five fundamental facts: (1) Stratigraphically and tectonically, the Alborz Mountains are related much more closely to Central Iran (Fig. 1). (2) No orogenies affected Iran in the Paleozoic Era, except for gentle epeirogenic movements. (3) The Iranian Plateau went through all stages of a fully developed Alpine-type orogeny during the Mesozoic and Tertiary Eras, with typical geosynclinal development in the Zāgros, Kopeh Dāgh, East Iranian Ranges, and Makrān, whereas Central Iran shows only incomplete geosynclinal development. (4) Important trends in the Alpine structural plan clearly were inherited from Precambrian orogeny. (5) A true “median mass” is limited to a part of the East Iran Lut Desert. One of the important discoveries in the early 1970s since Stahl’s 1911 “median mass,” Baier’s 1938 “double orogens with Central Iran eugeosyncline,” and Furon’s 1941 “Ural-Oman axis” was that the majority of the metamorphic rocks of Iran were not Precambrian in age, as was reported in numerous publications, but were Paleozoic rocks metamorphosed during the hitherto unknown Late Triassic and Late Cretaceous orogenic

Figure 16. Professor Setrāk Ābdāliān in 1960 at his home in Pāstor Avenue, Tehran. This is the only photograph that I was able to find; it was cropped out from a group family photograph. Courtesy of Yervānd Sābunchiān (London, 3 February 2014, personal commun.).

Development of geological perceptions and explorations on the Iranian Plateau movements. The studied metamorphics were: the Gorgān Schists, northern Alborz (Berberian et al., 1973); Sirjān Belt, south and southwest Central Iran (Berberian, 1972a, 1972b 1973, 1977; Sabzehei and Berberian, 1972; Hushmandzadeh et al., 1972, 1973; Hushmandzadeh and Berberian, 1972, 1973; Berberian and Nogol, 1974; Berberian and Alavi-Tehrani, 1977a, 1977b); Deh Salm in the eastern Lut Desert (Berberian and Soheili, 1973); and Kuh-e Āgh Bābā, Māku, in the NW (Variscan?; Berberian and Hamdi, 1977). 9.2. Plate-Tectonics Approach The revised interpretation of the structural evolution of Iran in a wider geological and geophysical context emerged in the light of the plate-tectonic theory and papers on plate tectonics of the Mediterranean region (McKenzie, 1970) and Red Sea and Africa (McKenzie et al., 1970). In the late 1960s and early 1970s, Luc-Emmanuel Ricou (1971) and Jean Braud (1978, 1988) started detailed mapping of the Neo-Tethyan ophioliteradiolarite belts of the Neyriz and Kermānshāh areas as their “state doctorate” projects (“thèse doctorat d’État,” Université de Paris, France). Ricou (1971), in his regional analysis of the NW Syria–SE Turkey–Zāgros–Oman ophiolite belt, stressed the importance of thrusting and nappe structures during the late Maastrichtian without going deeply into the cause of the major contraction. Nonetheless, plate tectonics was not taught at the Iranian universities until 1981 when I wrote the curriculum and started teaching at the University of Tehran and the University of Tarbiat Modarres in Tehran. Manuchehr Takin (1972) was possibly the first Iranian who tried to broadly analyze the Iranian geology in the context of continental drift and focused on the birth of the Tethyan Ocean between the Zāgros and Central Iran in the Triassic, giving rise to a trench-Cordillera and trench–island arc system and its closure during the Late Cretaceous. Takin (1972) considered that eastern Central Iran broke up into a microcontinent (a triangular-shaped “Central-and-East Iranian microcontinent,” CEIM) surrounded by narrow oceans, with ophiolites (“colored mélange”) marking their boundaries. Despite being influenced by the strong shadow of fixism and geosyncline ways of thinking in the Middle East, some authors showed that the Iranian Plateau has undergone different intense deformational phases caused by successive collisional orogenies. Hence, the traditional focus on local areas evolved in a broader regional context. The similarities between the Lower Paleozoic sedimentary units of the Zāgros, Central Iran, and the Alborz, together with the Mesozoic–Tertiary magmatic assemblages, led Stöcklin (1974a, 1974b, 1977), Berberian and King (1981), Berberian and Berberian (1981), Berberian et al., 1982, and Berberian (1983a, 1983b, 1983c, 1984, 1989) to the notion that the central-north Iranian structural units were once part of a broad Paleozoic marginal continental platform attached to Gondwana. They also recorded that the crust is composed of a set of continental blocks that detached from Gondwana and

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084715/spe525-03.pdf by University of Wisconsin Madison user

63

collided with Laurasia during the Triassic, with closure of the “Alborz Paleo-Tethys” in the north, and opening of the “Zagros Neo-Tethys” in the south (Stöcklin 1974a, 1977; Berberian and King, 1981; Berberian and Berberian, 1981; Berberian et al., 1982, Berberian, 1984, 1989). Stöcklin (1974a, 1977) interpreted the Mesozoic Sevān ophiolites to be the continuation of the northern Iran PaleoTethys. Although this was not the case, the southeastern continuation of the Mesozoic Sevān ophiolites in northwestern Iran was discovered in the Arasbārān region of the Qaradāgh (Arasbārān) range (Berberian et al., 1981). In their complete review of the Iranian geology and tectonics, Berberian and King (1981) presented correlation charts for the stratigraphic units with paleogeographic maps and evaluated the structural evolution of the Iranian Plateau since the late Precambrian in the context of the plate tectonics. The plate-tectonic aspect of the 1981 paper was corrected and completed by the papers of F. Berberian et al. (1982) and Berberian (1983a, 1983b), followed by two International Geological Conference papers in Moscow and Washington, D.C. (Berberian, 1984, 1989). The problem common to almost all the articles prior to 1982 (Takin, 1972; Stöcklin, 1974a, 1977; Berberian and King, 1981; and many more) was that because the ophiolite belt between the Zāgros and Central Iran was covered by the Maastrichtian sediments, almost all the authors stated that the Neo-Tethys (Zāgros) Ocean in the south was closed in the Late Cretaceous. It was only after detailed geochemical and geochronological analyses of the two calc-alkaline plutonic complexes of Bazmān (Makrān) and Natanz (Sirjān belt; Fig. 1) that for the first time it was proven that: (1) subduction of the Oman oceanic crust was well established by the Late Cretaceous; and (2) the Arabian-Central Iranian collision along the Neo-Tethys (Zāgros) did not occur during the late Cretaceous (Maastrichtian), but took place during the late Paleogene or early Neogene time (F. Berberian et al., 1982; Berberian, 1983b, 1983c, 1984, 1989). The plutons were dated by the Rb-Sr whole-rock isochron method at 74 + 2 Ma (Bazmān, Makrān) and 24 ± 4.5 Ma (Natanz, Karkas). Detailed trace-element studies together with low 87Sr/86Sr initial ratios ( Deformation & uplift along the Southern Alborz

7



Covered high-intensity magnetic patches (aeromagnetic maps)

8



S o u t h - v e r g i n g f o l d s w i t h n o r t h - d i p p i ng t h r u s t s

crust along its southern boundary. Further geochemical and geochronologic work on the south Alborz igneous rocks is needed to prove this scenario. 10. HONORING IRANIAN EARTH (AND PLANETARY) SCIENTISTS Since ancient times, Iran has been blessed and honored with great scholars and scientists working under difficult conditions and unusual circumstances in various fields. Nonetheless, very few Iranian earth scientists have gained the respect and honor they deserve for their outstanding scientific contributions and achievements. The following are a few examples introduced here in the hope that it will encourage future young Iranian earth scientists. Azophi, 1837: The Azophi lunar impact crater and the minor planet 12621 Alsufi were named by the astronomical com-

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084715/spe525-03.pdf by University of Wisconsin Madison user

munity honoring Abdolrahmān Sufi Rāzi (al-Sufi, Azophi), the famous Iranian scientist and astronomer (903–986 CE), who was born in the great city of Ray (ancient Rhagae), south of modern-day Tehran, lived in Esfahān, and passed away at Shirāz. The Azophi lunar impact crater, which is located at 22.1°S, 12.7°E, has a diameter of 47 km and a depth of 3.7 km (Beer and Mädler, 1837; Andersson and Whitaker, 1982; Berberian, 1997; Whitaker, 2003; Blue, 2007; IAU, 2015a, 2015b). Alfraganus, 1935: A Moon crater at 5.4°S, 19°E (diameter 20 km) is named after Ahmad Farghāni (ca. 798–865 CE), who calculated the diameter of the Earth by measuring the meridian arc length and wrote Kitāb fi Jawāmi’ ‘Elm al-Nujum, A Compendium of Science of the Stars in 833 (Andersson and Whitaker, 1982; Whitaker, 2003; Blue, 2007; IAU, 2015a, 2015b).

Figure 19. Schematic map of the Paleo-Tethys suture zone in northern central Alborz, and deformed Paleozoic rocks with equal area plot of b-lineation (B) and schistosity (S) in the Gorgān Schist. Figure is modified from Berberian et al. (1973). B.Gaz—Bandar Gaz (port).

68 M. Berberian

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084715/spe525-03.pdf by University of Wisconsin Madison user

Figure 20. Schematic map of the northern and southern Alborz Mountains in (top) Late Paleozoic–Triassic with Paleo-Tethyan assemblage and highly deformed Paleozoic rocks along the northern belt (ZZ); and (bottom) Cretaceous–Eocene with highly deformed southern belt. Filled circles with ages denote geochemical and geochronological samples located in the Eocene Karaj Formation volcanics (left) and the Mobārakābād gabbro (right) in the southern central Alborz (Verdel et al., 2011). The igneous rocks with arc-like trace elements are subduction products of either (top) the southern branch of the Neo-Tethys (Zāgros Ocean), or (bottom) the continuation of the northern branch of the Neo-Tethys (Sevān-Arasbārān-Sabzévār) Ocean (?). More research is needed to prove the latter case (if any). ZZ—intense compressional deformation, crystallization, and metamorphism.

Development of geological perceptions and explorations on the Iranian Plateau

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084715/spe525-03.pdf by University of Wisconsin Madison user

69

70

M. Berberian

Nasireddin, 1935: A 52-km-wide lunar impact crater at 41°S, 0.2°E on the southern hemisphere of the Moon is named honoring Nasir al-Din Tusi (1201–1274 CE), the Iranian polymath and prolific writer (astronomy, mathematics, biology, chemistry, mineralogy, and philosophy) who was born in the city of Tus of Khorāsān Province of eastern Iran (Andersson and Whitaker, 1982; Whitaker, 2003; Blue, 2007; IAU, 2015a, 2015b). Birunite, 1957: A new mineral of hydrated calcium silicate [Ca18(SiO3)8.5(CO3)8.5SO4·15H2O, orthorhombic system] was discovered in Uzbekistan and named in honor of Abu Rayhān Biruni (973–1048 CE), the genius Iranian polymath and one of the greatest scholars (natural scientist, mathematician, astronomer, geodesist, historian, and chronologist) who was born in “Birun” in the ancient Iranian Khārazm Province of Greater eastern Iran and lived in Ghazni (modern-day Afghanistan), the capital city of the Ghaznavid Turkic dynasty (977–1163) ruling in central and east Iran. Biruni wrote a book on the physical characteristics of minerals and for the first time discussed the motion/drift of the continents (plates in modern earth science), as well as earthquakes and crustal faulting, mountain building/orogeny, uplift and change in river base level, mineralogy, weathering, erosion and roundness, and artesian wells, among other subjects (Badalov and Golovanov, 1957; Berberian, 1997, 1992, 2014). Avicennite, 1958: A new mineral of thallium and iron oxides [Tl2O3, isometric] was discovered in Zirābulāk Mountain in Samarkand, modern-day Uzbekistan (Soghd Province in ancient Greater Iran; 1917–1990 Soviet Union) and named in honor of Avicennā (Pursinā, ebn Sinā, Abu ‘Ali Sinā, 980–1037 CE), the great Iranian polymath and one of the most significant scholars (scientist, physician, and philosopher). Avicenna was born in Bokhārā of the ancient Iranian Greater Khorāsān Province (Bukhara in modern-day Uzbekistan), the capital city of the Iranian Sāmānid dynasty (819–999), settled in the city of Ray (ancient Rhagae, south of Tehran), and died in Hamédān. Avicenna classified the minerals for the first time and wrote about the theory of earthquakes, principles of earth science, mountain building/ orogeny, and meteorites (Karpova et al., 1958; Kon’kova and Savel’ev, 1960; Berberian, 1997, 2014). Ulugh Beig, 1963: The Iranian mathematician, astronomer, and Timurid ruler (Mirzā Muhammad Tāraghay [Tāreq] b. Shāhrokh, 1394–1449 CE) was honored with the name of a crater on the Moon at 32.7°N, 81.9°W, with 54 km diameter (Andersson and Whitaker, 1982; Whitaker, 2003; Blue, 2007; IAU, 2015a, 2015b). Avicenna, 1970: A Moon crater at 39.7°N, 97.2°E (diameter 74 km) was named after the great Iranian polymath and jurist Avicenna (ca. 980–1037 CE; Andersson and Whitaker, 1982; Whitaker, 2003; Blue, 2007; IAU, 2015a, 2015b). Abulwafa, 1970: A Moon crater (1.0°N, 116.6°E, diameter 55 km) was named after the Iranian mathematician and

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084715/spe525-03.pdf by University of Wisconsin Madison user

astronomer Abulwafā Buzjāni (940–998), the author of Kitāb al-Majisti, Almagest (Andersson and Whitaker, 1982; Whitaker, 2003; Blue, 2007; IAU, 2015a, 2015b). Al-biruni, 1970: The Iranian mathematician, astronomer, and geodetist Biruni (ca. 973–1048 CE) was honored with the naming of a Moon crater with 77 km diameter located at 17.9°N, 92.5°E (Andersson and Whitaker, 1982; Whitaker, 2003; Blue, 2007; IAU, 2015a, 2015b). Omar Khayyam, 1970: A Moon crater located at 58°N, 102.1°W (diameter 70 km) was named after Omar Khayyām Neishāburi (ca. 1048–1131 CE), the Iranian mathematician, astronomer, philosopher, and poet. In 1079, Khayyām reformed the Iranian calendar, now known as the Jalāli or Khayyāmi Calendar (Andersson and Whitaker, 1982; Whitaker, 2003; Blue, 2007; IAU, 2015a, 2015b). Isculitoides seyedemamii, 1972: A Lower Triassic (Middle Spathian) ammonoid (phyllum: Mollusca, class: Cephalopoda, subclass: Ammonoidea, genus: Isculitoides) from the Nakhlak Group, north of Anārak in Central Iran, was named after Dr. Kāzem Seyed-Emāmi (Tozer, 1972). Metadagnoceras amidii, 1972: A Lower Triassic (Middle Spathian) ammonoid (pyllum: Mollusca, class: Cephalopoda, subclass: Ammonoidea, genus: Metadagnoceras) from the Nakhlak Group, north of Anārak in Central Iran, was named after Mehdi Amidi, who discovered the fossil (Tozer, 1972). Eophyllites davoudzadehi, 1972: A Lower Triassic (Middle Spathian) ammonoid (phyllum: Mollusca, class: Cephalopoda, subclass: Ammonoidea, genus: Eophyllites) from the Nakhlak Group, north of Anārak in Central Iran, was named after Monir Dāvudzādeh, who collected the fossil (Tozer, 1972). Al-Khwarizmi, 1973: A Moon crater at 7.1°N, 106.4°E (diameter 65 km) was named after the Iranian mathematician and astronomer Musā Khārazmi (Algoritmi/Algaurizin, ca. 780–850 CE; Andersson and Whitaker, 1982; Whitaker, 2003; Blue, 2007; IAU, 2015a, 2015b). Khademite, 1973: A new mineral of aluminum sulfate hydrate [Al(SO4)F·5H2O, orthorhombic], which was discovered in Sāghand area of east Central Iran, was named in honor of Dr. Nasrollāh Khādem (1910–1999), the founder and the first managing director of the Geological Survey of Iran (1962– 1978; Briand et al., 1973, 1977). Iranorhynchus seyedemamii n. gen., sp. nov., 1978: A new fossil of a lobe-finned fish (phylum: Chordata, class: Sarcopterygii, subclass: Dipnoi, genus: Iranorhynchus) of Late Devonian (Frasnian) age from the Kermān area was discovered and dedicated to Dr. Kāzem Seyed-Emāmi (Janvier and Martin, 1978). Tusi, 1979: A main-belt minor planet 10269 Tusi (1979 SU11) was discovered and named by the Soviet astronomer Nikolai Stepanovich Chernykh (Crimean Astrophysical Observatory) in 1979 in honor of the Iranian polymath Nasir al-Din Tusi (1201–1274 CE; Schmadel, 2012). Omarkhayyam, 1980: Minor planet 3095, which was discovered in 1980 by the Ukrainian astronomer Lyudmila Vasileyevra

Development of geological perceptions and explorations on the Iranian Plateau Zhuravelva, was named “3095 Omarkhayyam” honoring the Iranian polymath (IAU, 2015a, 2015b). Paranutilus emamii, 1981: An Upper Triassic Nautiloidea (phylum: Mollusca, class: Cephalopoda, subclass: Nautiloidea) discovered in the Ābādeh area of Central Iran was named in honor of Dr. Kāzem Seyed-Emāmi (Bando and Taraz, 1981). Senowbaridaryana raretrabeculata (Boiko), 1991: A NorianRhaetian sponge fossil from Caucasia was named honoring Dr. Baba Senowbari-Daryan, who collected the fossil (Boiko et al., 1991). Cadomites (Polyplectites) bozorgniai, sp. nov., 1998: A Middle Jurassic (late Bathonian) ammonoid (phylum: Mollusca, class: Cephalopoda, subclass: Ammonoidea, family: Stephanoceratidae, genus: Cadomites, subgenus: Polypectites) from the lower Baghamshāh Formation in the southwestern part of Tabas of eastern Central Iran was named honoring Dr. Fathollāh Bozorgniā (1922–2002) of the National Iranian Oil Company (Seyed-Emami et al., 1998). Umbroostrea emamii, 2001: A bivalve (phyllum: Mollusca, class: Bivalva, family: Ostreidae, genus: Umbrostea) from the Upper Triassic Nāyband Formation of the Tabas area, Central Iran, was named after Dr. Kāzem Seyed-Emāmi (Hautmann, 2001). Neseuretus ghavideli, 2004: An Early Ordovician trilobite (family: Calymemidae, subfamily: Reedocalymennidae, genus: Neseuretus) from the Shirgesht Formation of the Derenjāl Mountain in Central East Iran was discovered and named in honor of Dr. Mohammad Ghavidel-Syooki (Bruton et al., 2004). Welteria hamedanii sp. nov., 2005: A new sponge fossil (phylum: Porifra, class: Demospongea, family: Solenolmiinae, genus: Welteria) from the Upper Triassic (Norian–Rhaetian) of the Nāyband Formation of Central and east Iran was named after ‘Ali Hamèdāni, who discovered the fossil (Senowbari-Daryan 2005). Ghavidelia damghanensis, 2008: An Early to Middle Ordovician brachiopod (family: Ephippelasmatidae, genus: Ghavidelia) from the Lashkarak Formation of the Simeh Mountain, northwest of Dāmghān in the eastern Alborz Mountains was discovered and named after Dr. Mohammad GhavidelSyooki (Popov et al., 2008). Manberodus fortis sp. nov., 2008: A new fossil species of Chondrichthyes (jawed fish) Manberodus fortis gen. nov. (genus: Manberodus fortis, family: Aztecondotidae, order: Omalodontiformes) with age range of 383.7–382.4 Ma (early Frasnian; Late Devonian) was discovered in the Kaftar Mountain northeast of Esfahān and named in honor of Dr. Manuel Berberian in recognition of his contribution to the tectonic evolution of the Iranian Plateau (Hairapetian et al., 2008; Ginter et al., 2008, 2010). Amblysiphonella torabii sp. nov., 2011: An Upper Triassic sponge fossil (phylum: Mollusca, class: Bivalva, family: Ostreidae, genus: Umbrostrea) from the Nāyband Formation south of

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084715/spe525-03.pdf by University of Wisconsin Madison user

71

Bāgherābād was named after Hossein Torabi, who discovered the fossil (Rashidi and Senowbari-Daryan, 2011). 11. CONCLUDING REMARKS It is widely held that all the Western scientific and engineering thoughts and writings have their origins in the ancient Greek and Roman scholars and their great civilizations, which was followed by Islamic science and civilization during the Medieval Age of Europe. The review addressed in the scope of this essay shows that scientific concepts of the natural world were also found in the ancient Iranian civilization (ca. 1200 BCE–634 CE), with violent disruptions in 330 BCE and 637–652 CE. During the tenth and eleventh centuries, the greatest Iranian scholars of all time, with their monumental works benefiting from the ancient Iranian knowledge and civilization (as well as benefiting from the ancient Indian, Greek, and Roman civilizations), analyzed and discussed the physical changes of Earth’s surface through their detailed and time-consuming studies, experiments, and observations in various parts of the Iranian Plateau, India, and other parts of the ancient East. These Iranian scholars were definitive in their findings and well focused in their investigations, analyses, and interpretations prior to writing the results. Their observations and ideas were adequately widespread enough to address processes of change on Earth’s crust, the deep roots of the mountain belts and their growth and uplift during earthquakes, formation of rocks from sediments or by intense heat, the meaning of fossils, and much more. Hence, they were able to put forward a concept of the dynamism of Earth’s phenomena. Their findings and thoughts, indeed, have a modern tone and represent a great contribution to earth science. Their works influenced the European scholars during and after the Renaissance. Numerous social, political, and economic events that have caused severe drawbacks, decline, and cessation of scientific activities, including earth science in Iran, mainly began ca. 1031– 1040 CE during invasion and expansion by Turkic nomadic tribes, and especially after the Mongol invasions (ca. 1218–1220, 1231, 1370 CE), which coincided with the gradual rise of Europe from its so-called “Dark Ages” (ca. 500–1000 CE). Sporadic activities were recorded in excavation of mineral resources and some geological ideas, which mainly restated the ideas of the tenthand eleventh-century “Iranian intermezzo intellectual zenith” and geological perceptions. For example, the mineralogical book of Nasir al-Din Tusi (1256–1259 CE) was mainly based on the book of precious stones by Biruni (1040–1048 CE), and whatever Tusi allegedly took from previous scholars such as Abu Ya’ghub ibn Eshāq al-Kindi (810–873 CE) and Zakaryā Rāzi (Rhazes, Rasis; 854–925 CE) is similar to the writings of Biruni, who himself benefited from them. From the sixteenth through nineteenth centuries, the historical landscape of Iran was marked by invasions and political rivalry of the Ottoman Turks (12 wars with Iran from 1514 to 1821), Russia (five wars from 1651 to 1828), and Great Britain

72

M. Berberian

during their expanding hegemony on the Iranian Plateau and the Persian Gulf, coupled with decadence of the autocratic rulers in Iran, which expedited the decline and demise of science in the country. Modern geological studies were carried out by the European explorers and geologists beginning in the nineteenth and early twentieth centuries for general reconnaissance of natural resources. This culminated in discovery of the rich oil fields in southwest Iran (the first being the Masjid Solaimān field in 1908), which shaped the politics of the region due to its worldwide economic and geopolitical significance. After the establishment of the National Iranian Oil Company (NIOC) in 1950, and especially of the Geological Survey of Iran (GSI) in 1962, the Iranian geologists were trained by the Western experts and in Western universities, and conducted geological mapping and scientific research of their own. Sadly, during the last millennium, Iran was unable to reattain the status of the tenth–eleventh-century level of the intellectual zenith in sciences, and never became able to nurture a new generation of world-class scientists such as Khārazmi, Rāzi, Fārābi, Biruni, and Avicenna. From a sociological point of view, science and scientific thought, research, and problem solving could not become a native entity during the last millennium within a system devoid of democracy, transparency, and encouragement. Furthermore, the few elite statesmen and scholars who occasionally had the opportunity to change the status quo of the system were either assassinated (Nezām al-Molk Tusi, 1092; Shams al-Din Jovaini, 1284; Fazlollāh Hamédāni, 1317; Qā’em Maqām Farahāni, 1835; Amir Kabir, 1852) or imprisoned (Dr. Mosaddeq, Esq., 1953–1967) by the incompetent and corrupt tyrants. It is hoped that young Iranians of the present and coming centuries do a better job and build a brighter future in the advancement of earth sciences in their country. ACKNOWLEDGMENTS To set forth this history has been a matter of personal gratification for me after 46 yr of research mentioning the names and monumental works of the geniuses and remarkable Iranian and foreign earth science masters of the past who worked for Iran. The valuable contribution of these scholars, who were not driven by money, power, or religion, surpassed their own lifetime and reached to everybody and everywhere; they do not belong to a single country or nation. They were formed by every element of the early cosmos, and at the end of their time, their minds, knowledge, hearts, and souls flew and plunged into the world’s ocean of science and technology; they returned as exceptional rare elements to the planet Earth and, thus, attained perpetual lives. A section of this paper was presented on 29 October 2013 at the Colorado Convention Center, Denver, Colorado, during the 125th Anniversary Annual Meeting of the Geological Society of America (GSA, 2013), T188—Sessions 214 and 291, “Tethyan Evolution and Seismotectonics of Southwest Asia: In Honor of 40 Years of Manuel Berberian’s Research Contribu-

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084715/spe525-03.pdf by University of Wisconsin Madison user

tions.” I must tender my thanks to GSA and the session chair, Rasoul Sorkhābi, for organizing this special session and for his warm hospitality in Denver. I am grateful to Jovan Stöcklin (20 December 2001, 5 January and 29 May 2003, Seuzach, Switzerland); Mrs. Elisabeth Steingruber Stöcklin (3 October 2016, Seuzach, Switzerland); Marianne Huber Gluenz and Eva (Huber) Glunez (20 May 2003, London); Tillfried Cernajsek (15 February 2007, Geologische Bundesanstalt, Austria); Maurizio Gaetani (9 August, 13, 14, and 28 September 2006, 4 January 2007, September 29, 2016, University of Milan, Italy); Hushang Tarāz (20 May 2009, San Diego, California); Joseph McCall (6 October 1981, Geological Society of London); Milena Pika-Biolzi (30 July 2014, ETH Zürich, Switzerland); and Thomas Hofmann (30 July 2014, Bibliothek & Verlag, Geologische Bundesanstalt, Wien, Austria), who generously shared their precious stories, data, and invaluable historic photographs with me and who granted permission to use them in this essay. Seyyed Hosseisn Nasr kindly gave permission to use extracts from his publications in English and Persian. I also thank Nour Zaghi (California) for permission to use short excerpts from his e-book in this work. Likewise, I acknowledge Emād Kavāri (Tehran), Siāmak Āgāh (Houston, Texas), Hormoz Naficy (London), Manuchehr Takin (London), and Michael A. ‘Alā (London) for providing brief information regarding the National Iranian Oil Company (NIOC) Iranian geologists during 1961–1979, as well as Yervānd Sābunchiān (London), Sirus Arshadi (Germany), Kāmrān Ansari (Tehran), and Prince M.H. (Mickey) Kadjar (Qājār; Dallas, Texas) for retrieving some data for me throughout this research. Ārkādi Kārākhānyān (Yerevān, Republic of Armenia) kindly reminded me of the geological work of Ānāniā Shirākātsi (610–685 CE) and Mikhitār Herātsi (ca. 1120–1200 CE) and provided the relevant material. Comments and corrections by Jack Schroder Jr., Rasoul Sorkhābi, and Ezat Heydari, which contributed to improvement of the manuscript, are greatly appreciated. However, I exclusively remain responsible for the content of this paper. Special thanks go to the GSA staff for their efforts during the final editing and printing process. Research on this time-consuming work and digging through numerous ancient and old publications (the first ever documentation for Iran) was not supported by any grant, organization, or individual; it was merely supported by love and family pocket. It is my earnest hope that this essay stimulates interest in young Iranian geologists, and introduces them to the old masters of geology, who dedicated their lifetimes to a cause, and to some non-Iranian earth scientists. This report is dedicated to the memory of the commitment of Dr. Jovan Stöcklin (1921–2008), the Swiss geologist who spent 27 years of his productive life in Iran studying, mapping, and publishing on the unknown Iranian geology, tectonics, and mineral resources, as well as training the first generation of geologists both at the Iranian Oil Company/NIOC and the Geological Survey of Iran. Despite being a member of the second generation of geologists at the GSI while Jovan Stöcklin was at the GSI, I never had the honor or opportunity of working with Dr. Jovan Stöcklin in the field or in the office. On a personal Christmas

Development of geological perceptions and explorations on the Iranian Plateau

73

Figure 21. Jovan Stöcklin’s handwriting on a December 2001 Christmas card sent from Seuzach, Switzerland. Published with kind permission from Dr. Stocklin’s widow, Elisabeth Steingruber Stöcklin. Elisabeth wrote me that: “It is with pleasure that I give you the permission to publish the Christmas card written to you by my husband. I appreciate it very much that your report about geological exploration in Iran is dedicated to my husband. And I am also looking for reading your paper” (3 October 2016, personal commun., Seuzach, Switzerland).

card postmarked 20 December 2001 and mailed from Seuzach, Switzerland (Fig. 21), Jovan Stöcklin wrote me that:

I remember our last meeting: it was at the 1984 IGC in a waiting lounge of a Moscow airport; we had a brief chat, and I felt sorry for you, because you sat apart from your Iranian colleagues and seemed very depressed. … I understand that you are now living in the USA. … I very much hope that you have found a new way of life satisfactory to you and your family. … I have reached the age of 80, am in good health, have become 4-fold grandfather, am no longer working about geological problems but enjoying various other aspects of life such as reading, travelling, meeting family and friends, and often remembering the many years of fine cooperation with Iranian colleagues in that wonderful country. Your publications on seismotectonics of Iran continue to hold a prominent place in my library.

I cherish Jovan Stöcklin’s good thoughts, good words, and good deeds for the country and earth science at large, as well as his 2001 Christmas postal card (Fig. 21) and later communications regarding his scientific activities in Iran. REFERENCES CITED Aaboe, A., 1958, On Babylonian planetary theories: Centaurus, v. 5, no. 3–4, p. 209–277. Ābdāliān, S., 1962, La Tectonique de l’Iran: Tehran, Institute of Geophysics, Tehran University, no. 8, 76 p., with tectonic map of Iran. Abich, H. von, 1858, Vergleichende grundzüge der geologie des Kaukasus wie der Armenischen und nordpersische gebrige (prodromus einer geologie

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084715/spe525-03.pdf by University of Wisconsin Madison user

der Kaukasischen lander): Memoire of the Academy of Sciences St. Petersburg, ser. 6, no. 7, p. 359–534. Abich, H. von, 1859, Vergleichende grundzüge der geologie des Kaukasus wie der Armenischen und Nordpersischen gebirge (Prodromus einer geologie der Kaukasischen Länder): Memoire of the Academy of Sciences St. Petersburg, ser. 6, no. 7, p. 359–534. Abich, H. von, 1878, Geologische forschungen in den Kaukasischen Länden, Theil 1, Eine bergkalkfauna aus der Araxes-Enge bei Djulfa in Armenien: Wien, A. Hölder, 126 p. Adamiyat, F., 1975, Amir Kabir va Iran: Tehran, Kharazmi Publishers, 1354/1975, Volume 1, 357 p. Adams, F.D., 1938, The Birth and Development of the Geological Sciences: Baltimore, Maryland, The Williams and Wilkins Company, 506 p. (Reprinted in 1954 by Dover Publishers, New York. Also at archive.org/details/ birthanddevelopm031745mbp#page/n361/mode/1up/search/avicenna.) Afshar, 1961. See Estakhri, 951. Afshar, I., 1996, ‘Arāyes al-Javāher wa Nafāyes al-Atāyeb of ‘Abolqāsem ‘Abdollāh Kāshāni: Tehran, Anjoman Āthār Melli, no. 53, 406 p. [in Persian]. Afshar, I., 2004, Javāher-Nāmeh Nezāmi Mohammad ebn abi al-Barakāt Jowhari Neyshāburi dar Sāl-e 592 H [Encyclopaedia of Precious Stones of Jowhari Neyshāburi in the Year of 1162 C.E.]: Tehran, Iran Mirāth Maktub Publishers, 1383/2004, 445 p. [in Persian]. AGIP/SIRIP (Aziende Generale Italiana/Société Irano-Italienne des Pétroles), 1962, Geology of Southern Makran: Tehran, filed at the Geological Survey of Iran, unpublished Internal Report. ‘Ali, 1967. See Biruni, 1025. Alipur, K., 1993, History of Geology and Mining in Iran [“Tārikh-e Dānesh-e Zamin-shenāsi va Ma’adan dar Irān”]: Tehran, Geological Survey of Iran, 266 p. [in Persian]. Allenbach, P., 1966, Geologie und Petrographic des Damavand und seiner umgeburg (Zentral-Elburz), Iran: Geologisches Institut, ETH-Zürich Mitteilung 63, 114 p. Amidi, M, 1978, Geological Quadrangle Maps of Ahar: Tehran, Geological Survey of Iran, map no. D2, scale 1:250,000. Amiet, P., 1994, De Morgan, Jacque: Encyclopaedia Iranica, v. VIII, fasc. 2, p. 175–177; online at www.iranicaonline.org.

74

M. Berberian

Amirsadeghi, H., ed., 1977, Twentieth Century Iran: London, Heinemann, 299 p. Ānāniā Shirākātsi, ca. 660–681, Cosmography, translated from the Old Armenian and with commentary by Ter-Davtyan, K.S., and Arevshatyan, S.S., 1962: Yerevan, Yerevan University Publishing, p. 100–107. Anawati, G.C., 1976, Le Kitāb al-Saydana d’al-Bārānā, in The Commemoration Volume of Beruni International Congress in Tehran, 1355/1976, p. 75–96; online at iranicaonline.org. Anawati, G.C., 1979, The Kitāb al-jamāhir fi maārifat al-jawāhir of al-Bārānā, in Said, H.M., ed., Al-Bārānā Commemorative Volume, Proceedings of the International Congress held in Pakistan on the Occasion of Millenary of Abu Raihan Muhammad ibn Ahmad al-Biruni (C. 973–1051 AD), November 26, 1973 through December 12, 1973: Karachi, Hamdad National Foundation, 1979, p. 437–453. Anawati, G.C., 1989, Bārānā, Abā Rayhān, in Pharmacology and Mineralogy: Encyclopaedia Iranica, v. IV, fasc. 3, p. 281–282, http://www .iranicaonline.org/articles/biruni-abu-rayhan-v. Andersson, L.E., and Whitaker, E.A., 1982, NASA Catalogue of Lunar Nomenclature: National Aeronautics and Space Administration (NASA) Reference Publication AP-1097, 183 p., http://planet4589.org/astro/lunar/ RP-1097.pdf. Anklesaria, B.T., 1956, Zand-Akasih, Iranian or Greater Bundahishn, English transliteration and translation: Bombay, Rahnume Mazdayasnan Sabha, 196 p., digital version by J.P. Peterson available at http://www.avesta.org/ mp/grb.htm. Ārām, 1973. See Biruni, 1025. Arkell, W.J., 1955, George Martin Lees 1898–1955: London, Biographical Memoirs of Fellows of the Royal Society, Volume 1, p. 161–173, doi:10.1098/rsbm.1955.0012. Arrian of Nicomedia, second century CE (before 145), The Anabasis of Alexander, or, The History of the Wars and Conquests of Alexander the Great (E.J. Chinnock, tr., 1884, EBook #46976), Chapters XVI (Escape of Darius into Media—March of Alexander to Babylon and Susa), Chapter XVIII (Defeat of Ariobarzanes and Capture of Persepolis): http://www .gutenberg.org/files/46976/46976-h/46976-h.htm (accessed 2015). Asatryan, G., Danelian, T., Sosson, M., Sahakyan, L., Person, A., Avagyan, A., and Galoyan, G., 2010, Radiolarian ages of the sedimentary cover of the Sevan ophiolite (Armenia, Lesser Caucasus): Ofioliti, v. 35, no. 2, p. 91–101. Ashtiani, ‘A.E., 1984, Mirzā Taqi Khān Amir Kabir (Afshār, I., ed.) (3rd ed.): Tehran, Tus Publishers, 1363/1984, 158 p. Assereto, R., 1963, The Paleozoic Formations in Central Elburz (Iran): Rivista Italiana di Paleontologia e Stratigrafia, v. 69, no. 4, p. 503–543. Assereto, R., 1965, Sui caratteri stratigrafici e tettonici dei terreni del Mesozoico e del Terziario nei dintorni di garmabdar (Elburz Centrale, Iran): Rivista Italiana di Paleontologia e Stratigrafia, v. 71, no. 1, p. 167–174. Assereto, R., 1966a, The Jurassic Shemshak Formation in Central Elburz (Iran): Rivista Italiana di Paleontologia e Stratigrafia, v. 72, no. 4, p. 1133–1182. Assereto, R., 1966b, Geological Map of Upper Djadjerud and Lar Valleys (Central Elburz, Iran), Scale 1:50,000 with Explanatory Notes: Istituto di Geologia dell’ Università di Milano, ser. G., no. 232, 86 p. Avery, P., Hambly, G.R.G., and Melville, C., eds., 1991, The Cambridge History of Iran, Volume 7, From Nader Shah to the Islamic Republic: Cambridge, UK, Cambridge University Press, 1072 p. Avestā. See Poure Davoud, 1977a, 1977b. Avicenna (Ibn Sinā; Pur Sinā), Abu ‘Ali Hussain ibn Abdallāh, 1020, Fann-e Samā’-e Tabi’i az Ketab Shafā [The Book of the Remedy; Sufficientia; Mohamad ‘Ali Forughi, tr]: Tehrān, Vezārat Ma’āref va Owqāf va Sanāye’ Mostazrafeh Publishers, v. 1, 1316/1937, 684 p.; v. 2, 1319/1940, 779 p. [in Persian]. Also: Chāpkhāneh Majles Publishers, 1316/1937, 1949; and Amir Kabir Publishers, 1361/1982, 675 p. See also Holmyard, E.J., and Mandeville, D.C., 1927, eds. ’Abdal-Halim Mostair, Sa’id Zāyed, and ‘Aballāh Ismā’il, and 1965, Al-Shifa of Ibn Sina, National Sciences, Part 5 on Minerals and Meteorology: Cairo, Egypt, Amiri Publishers, 94 p. [in Arabic with summary in French]. Badakhshan, A., and Najmabadi, F., 2004, Oil Industry, II—Iran’s oil and gas resources, in Encyclopaedia Iranaica: http://www.iranicaonline.org/ articles/oil-industry-ii (accessed 2014). Badalov, S.T., and Golovanov, I.M., 1957, Birunite, a new mineral of the thaumasite group: Doklady Akademii Nauk Uzbekistan SSR, v. 12, p. 17–21 [in Russian]. Badian, E., 1985, Alexander in Iran, in Gershevitch, I., ed., The Cambridge History of Iran: Volume 2. The Median and Achaemenian: Cambridge, UK, Cambridge University Press, p. 420–501.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084715/spe525-03.pdf by University of Wisconsin Madison user

Bahār, 1990, Bondahésh Farnbagh Dādagi (Bundahishn by Farnbagh Dādagi), translated by M. Bahār: Tehran, Tus Publisher, 1369/1990, 237 p. [in Persian]. Bahrami, E., 1990, Dictionary of the Avesta: Tehran, Neyshabur Foundation Publishers, 1369 (original publication), Volume 4, 3553 p. Baier, E., 1938, Ein Beitrag zum thema zwischengebirge: Zentralblatt für Mineralogie und Geologie Abteilung B, v. 11, p. 385–399. Bailey, E.B., Jones, R.C., and Asfia, S., 1948, Notes on the geology of the Elburz Mountains, north-east of Teheran, Iran: Quarterly Journal of the Geological Society, London, v. 104, no. 1, p. 1–42. Bamberg, J., 1994, The History of the British Petroleum Company: Volume 2. The Anglo-Iranian Years, 1928–1954: Cambridge, UK, Cambridge University Press, 668 p. Bamberg, J., 2000, Bristish Petroleum and Global Oil, 1950–1975, The Challege of Nationalism (History of British Petroleum, Volume 3): Cambridge, UK, Cambridge University Press, 666 p. Bamberg, J., and Melville, C., 2002, Images of Persia, Photographs by Laurence Lockhart 1920s–1950s: Cambridge, Massachusetts, The Centre of Middle Eastern and Islamic Studies, University of Cambridge, 86 p. Bando, J., and Taraz, H., 1981, Upper Triassic cephalopods from Abadeh, Central Iran: Tehran, Geological Survey of Iran, Report no. 49, p. 105–116. Beaumont, P., Bonine, M., and McLachlan, K.S. eds., 1989, Qanat, Kariz and Khattara; Traditional Water System in the Middle East and North Africa: Edgeware, Middlesex, Means Press Ltd., 305 p. Beer, W., and Mädler, J.H., 1837, Der Mond nach seinen kosmischen und individuellen verhältnissen: Oder Allgemeine vergleichende Selenographie mit besonderer beziehung auf die von den verfassern herausgegebene Mappa selenographica: Berlin, Simon Schropp, 454 p. Bell, C.M., 1840, Geological notes on part of Mazundaran: Transactions of the Geological Society of London, ser. 2, v. 5, p. 577–581. Berberian, F., and Berberian, M., 1981, Tectono-plutonic episodes in Iran, in Gupta, H.K., and Delany, F.M., eds., Zagros–Hindu Kush–Himalaya Geodynamic Evolution: American Geophysical Union Geodynamics Monograph 3, p. 5–32, doi:10.1029/GD003p0005. Berberian, F., Muir, I.D., Pankhurst, R.J., and Berberian, M., 1982, Late Cretaceous and early Miocene Andean-type plutonic activity in northern Makran and Central Iran: Journal of the Geological Society, London, v. 139, no. 5, p. 605–614. Berberian, M., 1972a, Two Important Deformational and Metamorphic Phases in the Belt Northeast of the Zagros Thrust Line (Iran): Tehran, Geological Survey of Iran, Internal Report, 13 p. Berberian, M., 1972b, Progress Report on the Khabr and Deh-Sard Geological Maps (Two 1:100,000 Sheets from Hadji-Abad Quadrangle Map): Tehran, Geological Survey of Iran, Internal Report, 6 p. Berberian, M., 1973, Two Important Deformational and Metamorphic Phases in the Belt Northeast of Zagros Thrust Line (Iran)—A Brief Structural Review of the Sanandaj-Sirjan Belt: Tehran, Geological Survey of Iran, Internal Report, 27 p. Berberian, M., 1977, Three phases of metamorphism in the Hadji-Abad quadrangle (southeastern extremity of the Sanandaj-Sirjan zone); a paleotectonic discussion, in Berberian, M., ed., Contribution to the Seismotectonics of Iran, Part III: Geological and Mining Survey of Iran, Report no. 40, p. 239–262. Available at researchgate.net, academia.edu, and http:// manuelberberian.com. Berberian, M., 1983a, The Southern Caspian, a compressional depression floored by a trapped, modified oceanic crust: Canadian Journal of Earth Sciences, v. 20, no. 2, p. 163–183. Berberian, M., 1983b, Generalized tectonic map of Iran, in Berberian, M., ed., Contribution to the Seismotectonics of Iran, Part IV: Geological Survey of Iran, Report no. 52, scale 1:1,500,000 [in English]. Berberian, M., 1983c, Continental deformation in the Iranian Plateau, in Berberian, M., ed., Contribution to the Seismotectonics of Iran, Part IV: Geological Survey of Iran, Report no. 52, 625 p. [in English plus 74 p. in Persian]. Available at researchgate.net, academia.edu, and http:// manuelberberian.com. Berberian, M., 1984, Structural evolution and tectonics of the Iranian Plateau, a plate tectonic approach, in Proceedings of the 27th International Geological Congress, Moscow, 4–14 August 1984, Volume IV. Tectonics of Continental Fold Belts, Moscow, Section 07, p. 140: Utrecht, VNU Science Press. Berberian, M., 1989, Tectonic evolution of the Iranian mountain belts, in Abstracts of the 28th International Geological Congress, July 9–19, 1989, Washington, D.C.: Washington, D.C., the Congress, v. 1 of 3, p. 129–130.

Development of geological perceptions and explorations on the Iranian Plateau Berberian, M., 1991, Is the theory of earthquake Greek or Iranian?: Iranshenasi (Bethesda, Maryland), v. II, no. 4, p. 834–845 [in Persian] and p. 120 [abstract in English]. Berberian, M., 1992, Biruni Khārazmi and the theory of continental drift: Irānshenāsi (Bethesda), Maryland, v. IV, no. 1, p. 139–147 [in Persian]. Berberian, M., 1994, Natural Hazards and the First Earthquake Catalogue of Iran: Volume 1. Historical Hazards in Iran Prior to 1900, A UNESCO/ IIEES Project during the United Nations International Decade for Natural Disaster Reduction (IDNDR: 1900–2000): Tehran, International Institute of Earthquake Engineering and Seismology (IIEES), 603 p. [in English] plus 66 p. [in Persian]. Berberian, M., 1997, An Investigation into the History of Cosmogony and Earth Science in Ancient Iran [Jostāri Dar Dānesh-e Kayhān va Zamin Dar Irānveij]: Tehran, Bonyād Neyshābur Publishers, 551 p. [in Persian]. Berberian, M., 2014, Earthquakes and Coseismic Surface Faulting on the Iranian Plateau; A Historical, Social, and Physical Approach: Amsterdam Elsevier, Developments in Earth Surface Processes 17, 714 p., ISBN: 9780-444-63292-0. Berberian, M., and Alavi-Tehrani, M., 1977a, Structural analysis of the Hamadan metamorphic tectonites; a paleotectonic discussion, in Berberian, M., ed., Contribution to the Seismotectonics of Iran, Part III: Geological and Mining Survey of Iran, Report no. 40, p. 263–280. Available at researchgate.net, academia.edu, and http://manuelberberian.com. Berberian, M., and Alavi-Tehrani, M., 1977b, Structural pattern of the Hamadan metamorphic tectonites, in Berberian, M., ed., Contribution to the Seismotectonics of Iran, Part III: Geological and Mining Survey of Iran Report 40, 1 map. Berberian, M., and Arshadi, S., 1976, On the evidence of the youngest activity of the North Tabriz fault and the seismicity of Tabriz City, in Berberian, M., ed., Contribution to the Seismotectonics of Iran, Part II: Geological Survey of Iran, Report no. 39, p. 397–418. Available at researchgate.net, academia.edu, and http://manuelberberian.com. Berberian, M., and Hamdi, B., 1977, First Discovery of the Ordovician Beds and Conodonts in the Slightly Metamorphosed Rocks of the Kun-e-Agh Baba, Maku Quadrangle, Azarbaidjan: Tehran, Geological Survey of Iran, Internal Report, 7 p. Berberian, M., and King, G.C.P., 1981, Towards a paleogeography and tectonic evolution of Iran: Canadian Journal of Earth Sciences, v. 18, no. 2, p. 210–265. Berberian, M., and Nogol, M., 1974, Preliminary Explanatory Text of the Geology of Deh Sard and Khabr Maps with Some Remarks on the Metamorphic Complexes and the Tectonics of the Area (Two 1:00,000 Sheets from Hadji-Abad Quadrangle Map): Tehran, Geological Survey of Iran, Internal Report, 60 p. Berberian, M., and Soheili, M., 1973, Structural History of Central Lut— Consolidation of the Supposed Lut Block during the Early Kimmerian Orogeny; a Preliminary Field Note: Tehran, Geological Survey of Iran, Internal Report, 20 p. Berberian, M., Hushmandzadeh, A., and Lotfi, M., 1973, Deformational Phases and Related Metamorphism in the Gorgan Area: Tehran, Geological Survey of Iran, Internal Report, 14 p. Berberian, M., Amidi, S.M., and Babakhani, A., 1981, Discovery of the Qaradagh Ophiolite Belt, the Southern Continuation of the Sevan-Akera (Little Caucasus) Ophiolite Belt in Northwestern Iran (Ahar quadrangle), a Preliminary Field Note: Tehran, Geological Survey of Iran, Internal Report, 15 p. Biruni, Abu Rayhān Mohammad ebn Ahmad, 1000, Al-Āthar al-Baqiya ‘an alQorun al-Khaliya [Book of Vestige from Past Centuries, 390/1000, Akbar Dānāseresht, tr.]: Tehran, Khayyām Publishers, 1320/1941; Ebn-e-Sina Publishers 1352/1973; Amir Kabir Publishers, 1363/1984. Biruni, Abu Rayhān Mohammad ebn Ahmad, 1025, Tahdid Nihāyāt al-Amākin Li Tashih Masāfāt al-Masākin [Determination of the Coordinates of Positions for Correction of Distances between Cities, 416/1025, Ahmad Ārām, tr.]: Tehran, Tehran University Publishers, 1392, 1352/1973 (English tr. by Jamil ‘Ali: The American University of Beirut, Centennial Publications, Lebanon, 1967), 278 p. Also al-Biruni, Kitab al-Tahdid (F. Krenbow, tr.), in Al-Biruni Commemoration Volume: Calcutta, India, Iran Society, 1951. Biruni, Abu Rayhāan Mohammad ebn Ahmad, 1030, Kitāb fi Tahqiq mā li’lHind men Maqula Ma’qula fi al-‘Aql om Mazdula [Critical Study of What India Says; Whether Accepted by Reason or Refused, 421/1030], Tahqiq Mal al-Hend by Master Abu Raihan Biruni (Manuchehr Soduqi-Sahā, tr., 1362/1983): Tehran, Mo’asseseh Motāle’at va Tahqiqāt Farhangi, no.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084715/spe525-03.pdf by University of Wisconsin Madison user

75

521, v. 1, 260 p. [in Persian]. Also, Sachau, E.C., tr., 1910, Alberuni’s India, An Acount of the Religion, Philosophy, Geography, Chronology, Astronomy, Customs, Laws, and Astrology of India, 2 volumes: London, Kegan Paul, Trench Trubner & Co. Also see archive.org/details/ TahqiqMaliAlHind. Biruni, Abu Rayhāan Mohammad ebn Ahmad, 1050, Kitāb al-Jamāher fi Ma’refat al-Jawāher [The Sum of Knowledge about Precious Stones], Krenkow, F., ed.: Hayderabad, 1355/1963. Also in Anawati, G.C., The Kitāb al-Jamāhir fi Mar’rifat al-Jawāhir of al-Biruni, Al-Biruni Commemorative Volume: H. M. Said, Karachi, 1979, p. 437–453. English tr. by Hakim Mohamad Sa’id, 1989: Pakistan Hijra Council, Islamabad, Pakistan, 355 p. Blair, S.S., and Bloom, J.M., 1995, The Art and Architecture of Islam 1250– 1800: New Haven, Connecticut, Yale University Press, 348 p. Blanford, W.T., 1872, Note on the geological formations seen along the coast of Biluchistan and Persia from Karachi to the head of the Persian Gulf, and on some of the Gulf islands: Records of the Geological Survey of India, v. 5, no. 2, p. 41–45. Blanford, W.T., 1873, On the nature and probable origin of the superficial deposits in the valleys and deserts of Central Persia: Quarterly Journal of the Geological Society, London, v. 29, part 1, p. 493–503. Blanford, W.T., 1876, Eastern Persia, An Account of the Journeys of the Persian Boundary Commission, 1870–1872, Volume 2, Zoology and Geology: London, Macmillan, 516 p. Blue, J., 2007, Gazetteer of Planetary Nomenclature: planetayrnames.wr.usgs .gov (accessed 2015). Bobek, H., 1934, Reise in Nortwest Persien 1934: Zeitschrift der Gessellschaft für Erdkunde, Koner zu Berlin, v. 9/10, p. 359–369. Bobek, H., 1937, Die Rolle der Eiszeit in Nordwestiran: Zeitschrift für Gletscherkunde, v. 25, p. 130–185. Bobek, H., 1957, Die Takht-e Sulaimangruppe im mittleren Alburzgebirge, Nordiran, in Wiche, K., ed., Festschrift zur Hundertjahrfeier der Geographischen Gesellschaft in Wien 1856–1965: Vienna, Geographische Gesellschaft Vienna, p. 236–264. Bobek, H., 1959, Features and Formation of the Great Kavir and Masileh: Tehran, University of Tehran Arid Zone Research Center Publication no. 2, 63 p. Bobek, H., 1961, Die Salzwüsten Irans al Klimazeugen: Kommissionsverlag der Österreichische Akademie der Wissenschaften: Sonderabdruck aus dem Anzeiger der philosophischen-histtorisch Klasse, no. 3, p. 7–19. Bobek, H., 1963, Nature and implications of Quaternary climate changes in Iran, in Changes of Climate, Proceedings of Symposium on Changes of Climate with Special Reference to Arid Zones: Rome, 1961, UNESCO, p. 403–413. Bobek, H., 1969, Zur Kenntnis der Südlichen Lut, ergebnisse einer luftbildanalyse: Wien, Mitteilungen de Österreichischen Geographischen Gesellschaft, v. 111, hft. II/III, p. 155–192. Bogdanoff, A.A., Mouratov, M.V., and Schatsky, N.S., eds., 1964, Tectonics of Europe, Explanatory Note to the International Tectonic Map of Europe, scale 1:2,500,000, Subcommission for Tectonic Map of the World: Mosco, Akademiya Nauk, SSSR, 360 p. Böhne, E., 1929, Üeberblick ueber erzlagerstäetten Persiens und den gegenwärtigen stand von gewinnung und verhüettung: Erzmetall, v. 26, no. 3, p. 57–61. Bostock, J., and Riley, T.H., eds., 1855. See Pliny the Elder, ca. 79. Böhne, E., 1932a, Die steinkohlenvorcommen Persien: Zeitschrift fuer Praktische Geollogische, v. 40, p. 113–119, 132–141. Böhne, E., 1932b, Die wirtschafliche bedeutung und ersccliessung der steinkohlenvorkommen Persiens: Zeitschrift fuer Praktische Geologische, v. 40, no. 10, p. 145–148. Boiko, E.V., Belyaeva, G.V., and Zhuravleva, I.T., 1991, Phanerozoic Sphinctozoans from the Territory of the USSR: Moscow, Nauka Academy of Sciences USSR, Siberian Department, Institute of Geology and Geophysics, Academy Sciences Tajikistan, USSR, Institut Tajikistan, 233 p. [in Russian]. Bondéhéshn. See Bundashishn. Bosworth, C., 1973, The heritage of rulership in early Islamic Iran and the search for dynastic connections with the past: Tehran, Iran, Journal of the British Institute of Iranian Studies, v. XI, p. 51–62. Boucharlat, R., 2013, Mercury in the walls of the Shaur Palace, in Perrot, J., ed., The Palace of Darius at Suza; The Great Royal Residence of Achaemenid Persia: London, I.B. Tauris & Co Ltd., p. 464–467.

76

M. Berberian

Boyce, M., 1982, A History of Zoroastrianism, Volume II: Leiden, Netherlands, E.J. Brill, p. 289–292. Boyce, M., 1989, A History of Zoroastianism: Volume I. The Early Period: Leiden, Netherlands, E.J. Brill, 350 p. Boyle, J.A., 1965, The Cambridge History of Iran: Volume 5. The Saljuq and Mongol Periods: Cambridge, UK, Cambridge University Press, 763 p. von Brandt, J.F., 1870, Uber die von Herrn Magister Adolph Goebel auf Seiner Persischen Reise bei der Stadt Maragha in der Provinz Aderbeidjan gelfundenen Saugethier-Reste: Denkschrift Naturforscher-Vereins zu Riga, v. 187, p. 1–8. Braud, J., 1978, Les principales étapes de la formation du Zagros d’agrès l’étude structurale du secteur Kurde (transversale de Kermanshah): Paris, Société Géologique de France, Articles, Sommaires, Analyses, Correspondances, v. 4, p. 3–11. Braud, J., 1988, La Suture du Zagros au Niveau de Kermanshah (Kurdistan Iranien); Reconstitution Paléogéographique, Évolution Géodynamique, Magmatique et Structurale [Thèse Doctorat d’État]: Paris, Géodiffusion Mémoire 5, 489 p., ISBN: 2-907089-04-8 [from the unpublished thèse Doctorat d’État, 1978, Université de Paris-Sud, Faculté des Sciences d’Orsay (Essonne), no. national de thèse: 1987PA112506, 489 p.; online at http://www.sudoc.abes.fr/DB=2.1//SRCH?IKT=12&TRM=0435779 54&COOKIE=U10178,Klecteurweb,D2.1,E90c353d0-d26,I250,B341 720009+,SY,A%5C9008+1,,J,H2-26,,29,,34,,39,,44,,49-50,,53-78,,80 -87,NLECTEUR+PSI,R71.170.79.243,FN. Briand, P., Berthelon, J.P., Cesborn, F., and Sadrzadeh, M., 1973, New aluminum sulphate hydrate, Khademite of Saghand (Iran): Compte Rendu Hebdomadaires des Seances de l’Académie des Science, ser. C, v. 277, p. 1585–1588. Briand, P., Cesborn, F., and Berthelon, J.P., 1977, Iran Sulfates of Saghand near Yazd (Iran): Memoire Hors: Série de la Société Géologique de France 8, p. 77–85. British Petroleum Co, Ltd., 1956a, Geological maps and sections of southwest Persia, Proceedings of the 20th International Geological Congress, Mexico, Based on the Work of the Geological Staff of the Anglo-Persian Limited 1909–1033 and Anglo-Iranian Oil Company Limited 1933–1951, Staff of the British Petroleum Company (BP): London Edward Stanford Ltd., 17 color 1:1,000,000 scale maps and transects in folder. British Petroleum Co, Ltd., 1956b, Oil and gas in southwest Iran: Mexico City, Proceedings of 20th International Geological Congress, Mexico, Symposium sobre yacimientos de Petoleo y Gas, the International Geological Congress (IGC), v. II, p. 33–72. British Petroleum Co, Ltd, 1964a, Notes to accompany 1:250,000 geological maps of southwest Iran: London, the British Petroleum Company Ltd. British Petroleum Co, Ltd, 1964b, Geological Maps, Columns, and Sections of High Zagros of South-West Iran, scale 1:250,000: London, British Petroleum Company, Ltd. Bruton, D.L., Wright, A.J., and Hamedi, M.A., 2004, Ordovician trilobites from Iran: Palaeontohraphica Abteilung A, v. 271, no. 5–6, p. 111–149. Bühse, F.A., 1849, Bergreise von Gilan nach Asterabad (Asterabad, October 1848), in Beiträge zur Kenntniss des Russischen Reiches, und der angrenzenden Länder Asiens v. 13 and 14, p. 217–236: St.-Petersburg Akademie der Wissenschaften, books.google.com/books?id=JD8ZAQAAIAAJ&pg =PA215&dq=Bergreise+von+Gilan+nach+Asterabad&source=gbs_toc _r&cad=4#v=onepage&q=Bergreise%20von%20Gilan%20nac. Bühse, F.A., 1892, Die grosse Persische Salzwüste und ihr umgebung: Leipzig, Deutsche Rundschau für Geographische und Statistik, v. 15. Bundahishn [Bondéhéshn, Bondahésh]: a) in English: Pahlavi texts, The Sacred Books of the East in 5 Volumes, translated by E.W. West: Delhi, India, Motilal Banarsidass, 1962, vol. 5, 151 p. b) in Persian: Bondahésh by Farnbagh Dādagi, translated by M. Bahār: Tehran, Tus Publisher, 1369/1990, 237 p. c) in Persian: Bondehesh Hendi, translated by R. Behzādi: Tehran, Mo’asseseh Motāle’āt va Tahqiqāt Farhangi Publishers, no. 593, 1368/1989, 392 p. Burnett, C., 2015, The transmission of science and philosophy, in Kedar, B.Z., and Wiesner-Hanks, M., eds., The Cambridge World History: Volume 5. Expanding Webs of Exchange and Conflict, 500 CE–1500 CE: Cambridge, UK, Cambridge University Press, ch. 13, p. 339–388. Busk, D.L., 1934, Climbing and ski-ing in the Elburz Range, North Persia, 1931–2: London, Alpine Journal, v. 45, no. 247, p. 334–341. Busk, D.L., 1935, Climbing in the Takht-i-Suleiman Group, N. Persia: London, Alpine Journal, v. 47, no. 251, p. 290–309.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084715/spe525-03.pdf by University of Wisconsin Madison user

Busk, D.L., 1937, The German expedition in the Elburz Range, N. Persia: London, Alpine Journal, v. 49, no. 255, p. 245–247. Busse, H., 1975, Iran under Buyids, in Frye, R.N., ed., The Cambridge History of Iran, Volume 4: From the Arab Invasion to the Saljuqs: Cambridge, UK, Cambridge University Press, p. 250–304. Carpenter, W.B., and Brady, H.B., 1869, Description of Parkeria and Loftusia, two gigantic species of arenaceous foraminifera: Philosophical Transactions of the Royal Society of London, v. CLIX, p. 739–754. Cartier, E.G., 1971, Die geologie des unteren Chalus talis, zentral-Alborz/Iran [Dissertation]: Zürich, Mitteilungen Geologischen Institut, ETH, University of Zürich, N.F., No. 164, 134 p. Cernajsek, T., 2007, Director i.R. Hofrat Dr. Anton Ruttner, 19. März 1911– 18. September 2006: Jahrbuch der Geologischen Bundesanstalt B.-A. (JB. Geol. B.-A., Wien), v. 147, no. 3–4, p. 517–522, ISSN: 0016-7800, https://opac.geologie.ac.at/wwwopacx/wwwopac.ashx?command=getco ntent&server=images&value=JB1473_517_A.pdf. Cernajsek, T., 2009, Geologen-Wege-ein Rāckblick (1930/95) [The Path of a Geologist—A Retrospective (1930–1995)]: Jahrbuch der Geologischen Bundesanstalt B.-A. (JB. Geol. B.-A., Wien), v. 149, no. 2–3, p. 241–256, ISSN: 0378-0864, http://www.zobodat.at/pdf/ JbGeolReichsanst_149_0241-0256.pdf. Chardin, J., 1811, Les Voyages du Chevalier Jean Chardin en Perse et Autres Lieux de l’Orient (1686): Paris, M.L. Langlés, 10 volumes. Also, ‘Arizi, H., tr., Safar Nāmeh Chardin (1664–1670/1074–1080, and 1673–1677/1084– 1088 in Iran): Tehran, 1362/1983; ‘Abbāsi, M., tr., Siyāhat Nāmeh-ye Chardin (1664–1670/1074–1080, and 1673–1677/1084–1088 in Iran): Tehran, Amir Kabir Publisher, 10 volumes, 1338/1959 [in Persian]. Circle of Ancient Iranian Studies (CAIS)–School of Oriental Studies (SOAS), 2006, Archaeological News: The Circle of Ancient Iranian Studies, University of London: cais-soas.com/News/2006/April2006/19–04.htm (19 April 2004) (accessed 2014). Circle of Ancient Iranian Studies (CAIS)–School of Oriental Studies (SOAS), 2009, Archaeological News: The Circle of Ancient Iranian Studies, University of London: cais-soas.com/News/2009/August2009/14–08.htm (14 August 2009 and 4 October 2009) (accessed 2014). Clapp, F.G., 1940, Geology of eastern Iran: Geological Society of America Bulletin, v. 51, no. 1, p. 1–102, doi:10.1130/GSAB-51-1. Cohen, S., 1973, Enmerkar and the Lord of Aratta [Ph.D. thesis]: Philadelphia, Pennsylvania, University of Philadelphia. Cox, L.R., 1936, From southern Persia (Iran) and Bahrein Island: New Delhi, Palaeontologica Indica, Memoire of the Geological Survey of India, v. 22, no. 2, 69 p. Curtis, V.S., 1993, Persian Myths: London, The British Museum Press, 80 p. Curzon, G.N., 1892, Persia and the Persian Question: London, Longmans, Green, and Co, Volume II, 510 p. Dānāseresht, A., 1941, tr. See Biruni, 1000. Darmesteter, J., 1880, The Zend-Avesta: Part 1. The Vendidad: Oxford, UK, Oxford at the Clarendon Press, 240 p. Darmesteter, J., 1898, Yashts, The Sacred Books of the East: Avesta.org (accessed 2014). Daryaee, T., 2013, Sasanian Persia, The Rise and Fall of an Empire: London, I.B. Tauris, 256 p. Davies, R.G., compiler, 1977, Geological Map of Bandar-e Pahlavi: Tehran, Geological Survey of Iran Map D3, scale 1:250,000. Davies, R.G., Jones, C.R., Hamzepour, B., and Clark, G.C., 1972, Geology of the Masuleh Sheet, 1:100,000, Northwest Iran: Geological Survey of Iran, Report no. 24, 110 p. Dayton, J.E., 1971, The problem of tin in the ancient world: World Archaeology, v. 3, no. 1, p. 49–70. de Böckh, H., Lees, G.M., and Richardson, F.D.S., 1929, Contribution to the stratigraphy and tectonics of the Iranian ranges, in Gregory, J.W., ed., The Structure of Asia: London, Methuen, p. 58–176. Desio, A., 1934a, Appunti geografici e geologici sulla catena dello Zardeh Kuh in Persia [Geographic and geologic information concerning the Zard mountains in Persia]: Memorie Geologiche e Geographiche di G. Dainelli, 4, p. 141–167. Desio, A., 1934b, Sull’esistenza di piccoli ghiacciai nella Persia occidentale [Concerning the existence of small glaciers in western Persia]: Bollettino del Comitato Glaciologico Italiano, no. 14, p. 39–52. de Vries, J.J., 2006, History of groundwater hydrology, in Delleur, J.W., ed., The Handbook of Groundwater Engineering: Boca Raton, Florida, CRC Press, Taylor & Francis Group, p. 1-1–1-39.

Development of geological perceptions and explorations on the Iranian Plateau Dedual, E., 1967, Zur Geologie des mittleren und unteren Karaj-Tales, ZentralElburz (Iran): Zürich, Mitteilungen Geologischen Institut, ETH, University of Zürich, N.F. 79, 123 p. Dellenbach, J., 1964, Contribution à l’Étude Géologique de la Région Située à l’Est de Téhéran (Iran) [Ph.D. thesis]: Strasbourg, France, Faculty of Science, University of Strasbourg, France, 120 p. Didon, G., and Gemain, Y.M., 1976, Le Sabalan Vulcan Plio-Quaternaire de l’Azerbaïdjan Oriental (Iran), Étude Géologique et de Pétrographie de l’Édifice et de son Environnement Régional [Thèse 3eme cycle]: Institut de Géologie de Dolomieu, Grenoble University, France, 304 p. Diehl, E., 1944, Beitrag zur kenntnis des erzfundstellen Irans: Schweizerische Mineralogische und Petrographische Mitteilungen, v. 24, no. 2, p. 33–371. Dietrich, W.O., 1937, Ordoviz in Nordwest-Iran: Zentralblatt für Mineralogie und Paläontologie, Abteilung A, v. 10, p. 401–404. Dilek, Y., Imamverdiyev, N., and Altunkaynak, S., 2010, Geochemistry and tectonics of Cenozoic volcanism in the Lesser Caucasus (Azerbaijan) and the peri-Arabian region: Collision-induced mantle dynamics and its magmatic fingerprint: International Geology Review, v. 52, p. 536–578. Dilke, O.A.W., 1985, Greek and Roman Maps: London, Thames and Hudson, 224 p. Dodge, B., ed., 1970, The Fihrist of al-Nadim, A Tenth-Century Survey of Muslim Culture, Volume II. The Seventh Chapter: New York, Columbia University Press, www.muslimphilosophy.com/books/nad-phil.pdf (accessed 2013). Douglas, J.A., 1929, A marine Triassic fauna from eastern Persia: Quarterly Journal of the Geological Society, London, v. 85, pt. 4, no. 340, p. 624–650. Douglas, J.A., 1936, A Permo-Carboniferous Fauna from Southwest Persia (Iran): New Delhi, Palaeontologica Indica, Memoire of the Geological Survey of India, v. 22, no. 6, 55 p. Douglas, J.A., 1950, The Carboniferous and Permian Faunas of South Iran and Irania Baluchestan: New Delhi, Palaeontologica Indica, Memoire of the Geological Survey of India, v. 22, no. 7, 57 p. Douvillé, H., 1900, Les explorations géologiques de M.J. de Morgan en Perse: Proceedings of the 8th International Geologic Congress, v. 8, pt. 1: Paris, International Geologic Congress, p. 439–446. Douvillé, H., 1902, La lacune du terrain nummulitique de Soh (Perse) entre Kachan et Isphahan: Paris, Compte Rendu Société Géologique de France, v. 4, no. 2, p. 65–66. Douvillé, H., 1904a, Les explorations de M. de Morgan en Perse: Bulletin Société Géologique, ser. 4, no. 4, p. 539–553. Douvillé, H., 1904b, Mission J. de Morgan: Paris, Etudes Géologiques, v. 3, Paléontologique, pt. 2, Mollusques Fossils, Leroux, p. 191–380 + 50 plates. Douvillé, H., 1905, Les découvertes paléontologiques de M. de Morgan en Perse: Paris, Compte Rendu Société Géologique de France, v. 140, p. 891–893. Douvillé, H., 1910, Étude sur le Rudistes, Rudistes de Sicile, d’Algérie, d’Égypte, du Liban et de la Perse: Paris, Memoire Societe Géologique France, v. 18, no. 41, pt. 1, p. 1–84 + 77 figs. and 7 plates. Duncan, P.M., and Sladen, W.P., 1880 (1871–1895), The Fossil Echinoidea of Sind: Palaeontologia Indica, ser. 7 and 14, v. 1, 382 p. Durant, W., 1942, The Story of Civilization: Volume 1. Our Oriental Heritage: New York, Simon and Schuster, https://archive.org/stream/ storyofcivilizat035369mbp/storyofcivilizat035369mbp_djvu.txt (accessed 2014). Durant, W., 1950, The Story of Civilization: IV. The Age of Faith: New York, Simon and Schuster, 1196 p. Dyson, R.H., Jr., and Howard, S.M., eds., 1989, Tappe Hesar; Reports of the Restudy Project 1976: Florence, Italy, CASA Editrice le Lettere and Monographie di Mesopotamia, II, 151 p. Ebn Hauqal. See Hauqal. Ehlers, E., 2001, Geology: Encyclopaedia Iranica, v. X, fasc. 5, p. 456–460 (www.iranicaonline, 2014). Elgood, C., 1951, 2010, A Medical History of Persian and the Eastern Caliphate from the Earliest Times Until the Year A.D. 1932: Cambridge Library Collection, History of Medicine: Cambridge, UK, Cambridge University Press, 642 p., ISBN: 978-1-108-0188-2. Elles, G.L., 1930, The geology of salt plugs in Laristan southern Persia: Quarterly Journal of the Geological Society, London, v. 86, p. 463–520, doi:10.1144/GSL.JGS.1933.086.01-04.18. Engalenc, M., 1968, Contribution à la Géologie, Géomorphologie, Hydrogéologie de la Région de Teheran (Iran) [Thèse]: Montpellier, France, Centre d’Études et de Recherches Hydrogéologiques, 365 p. English, P.W., 1968, The origin and spread of qanats in the old world: Proceedings of the American Philosophical Society, v. 112, no. 3, p. 170–181.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084715/spe525-03.pdf by University of Wisconsin Madison user

77

Estakhri, Abu Eshāq Ebrāhim ibn Muhammad Fārsi, 951, Masālek va Mamālik, 350/951 [Roads and Countries, I. Afshār, tr.] (3rd ed.): Tehran, Bongāh Tarjomeh va Nashr Ketāb Publisher, 1368/1989, 356 p. [in Persian]. Falagas, M.E., Markadoulia, E.A., and Sammonis, G., 2006, Arab science in the Golden Age (750–1258 CE) and today: Federation of American Societies for Experimental Biosciences (FASEB) Journal, v. 20, no. 10, p. 1581– 1586, doi:10.1096/fj.06-0803ufm. Falcon, N.L., 1967, The geology of the north-east margin of the Arabian basement shield: British Association for the Advancement of Science, v. 24, p. 31–42. Fantini Sestini, N., 1965a, The geology of the Upper Djadjerud and Lar Valleys (north Iran): II. Paleontology; Bryozoans, brachiopods, and molluscs from Ruteh Limestone (Permian): Rivista Italiana di Paleontologia e Stratigrafia, Milano, v. 71, no. 1, p. 13–108. Fantini Sestini, N., 1965b, The geology of the Upper Djadjerud and Lar Valleys (north Iran): II. Paleontology, brachiopods from Dorud Formation: Rivista Italiana di Paleontologia e Stratigrafia, Milano, v. 71, no. 3, p. 773–788. Fantini Sestini, N., 1965c, The geology of the Upper Djadjerud and Lar Valleys (north Iran): II. Paleontology; On some Spinomarginifera from the Upper Permian of Mubarak-abad: Rivista Italiana di Paleontologia e Stratigrafia, Milano, v. 71, no. 4, p. 939–996. Fantini Sestini, N., 1966a, The geology of the Upper Djadjerud and Lar Valleys (north Iran): II. Paleontology; Brachiopods from Geirud Formation, Member D (Lower Permian): Rivista Italiana di Paleontologia e Stratigrafia, Milano, v. 72, no. 1, p. 9–60. Fantini Sestini, N., 1966b, The geology of the Upper Djadjerud and Lar Valleys (north Iran): II. Paleontology; Upper Liassic Molluscs from Shemshak Formation: Rivista Italiana di Paleontologia e Stratigrafia, Milano, v. 72, no. 3, p. 795–854. Fantini Sestini, N., and Glaus, M., 1966, Brachiopods from the Upper Permian Nissen (North Iran): Rivista Italiana di Paleontologia e Stratigrafia, Milano, v. 72, no. 4, p. 887–932. Fattahi, M., 2015, OSL dating of the Miam qanat (kariz) system in NE Iran: Journal of Archaeological Science, v. 59, p. 54–63. Fermor, L.L., 1943, Dr. G.E. Pilgrim, F.R.S: Nature, v. 152, no. 3862, p. 528– 529, doi:10.1038/152528a0. Ferrier, R.W., 1982, The History of the British Petroleum Company: Volume 1. The Discovery Years, 1901–1932: Cambridge, UK, Cambridge University Press, 698 p. Fisher, W.B., ed., 1968, The Cambridge History of Iran: Volume 1. The Land of Iran: Cambridge, UK, Cambridge University Press, 766 p. Flügel, H., 1964, The geology of the Upper Djadjerud and Lar Valleys (north Iran): II. Paleontology; Permian corals from Ruteh Limestone: Rivista Italiana di Paleontologia e Stratigrafia, Milano, v. 70, no. 3, 29 p. Forbes, R.J., 1966, Studies in Ancient Technology: Volume 7. Ancient Geology, Ancient Mining and Quarrying: Ancient Mining Techniques: Leiden, Netherlands, E.J. Brill Publishers, 259 p., second revised edition 1997, ISBN: 978-9004006270. Forbes, R.J., 1971, Studies in Ancient Technology: Volume 8. Metallurgy in Antiquity: Leiden, Netherlands, E.J. Brill Publishers, 295 p., ISBN: 9789004026520. Fraser, J.N., 1824, Notes made in the course of a voyage from Bombay to Bushire in the Persian Gulf (transmitted with a series of illustrative specimens): Transactions of the Geological Society of London, ser. 2, no. 1, p. 409– 412, doi:10.1144/transgslb.1.2.409. Frye, R.N., 1963, The Heritage of Persia: The World History of Civilization: Cleveland, Ohio, World Publishing Co., 301 p. Frye, R.N., 1965, The New Persian reconnaissance in western Iran, in Makdid, G., ed., Arabic and Islamic Studies in Honour of Hamilton A.P. Gibb: Leiden, Netherlands, E.J. Brill, p. 225–231. Frye, R.N., 1975a, The Cambridge History of Iran: Volume 4. The Period from the Arab Invasion to the Saljuqs: Cambridge, UK, University Press, 734 p. Frye, R.N., 1975b, The Golden Age of Persia, The Arabs in the East: London, Weidenfeld and Nicolson, 290 p. Furon, R., 1941, Géologie du plateau Iranien (Perse-Afghanistan-Bélouchistan): Memoire de Musée Nationale d’Histoire Naturelle, n.s., v. 7, no. 2, p. 177–414. Fürsich, F.T., Wilmsen, M., Seyed-Emami, K., and Majidifard, M.E., 2009, The mid-Cimmerian tectonic event (Bajocian) in the Alborz Mountains, northern Iran: Evidence of the break-up unconformity of the South Caspian Basin, in Brunet, M.-F., Wilmsen, M., and Granath, J.W., eds., South Caspian to Central Iran Basins: Geological Society, London, Special Publication 312, p. 189–203, doi:10.1144/SP312.5.

78

M. Berberian

Gabriel, A., 1929, Im Weltferenen Orient, Eim Reischbericht: Munich, Verlag von R. Oldenbourg, 365 p. Gabriel, A., 1934, Beobachtungen im Wuestenguertel Innerpersiens 1933: Wien, Mitteilungen der Geographischen Gesellschaft, v. 77, no. 1–6, p. 53–77. Gabriel, A., 1935, Durch Persiens Wüsten, neue Wanderungen in den Trockenräumen Innerirans: Stuttgart, Strecker und Schröder Verlag, 272 p., ASIN: B003WUGK40. Gabriel, A., 1938, The southern Lut and Iranian Baluchistan: Geographical Journal, London, v. 92, no. 3, p. 195–210. Gabriel, A., 1939, Aus den einsamkeiten Irans, Dritte Forschungsfahrt durch die Wüste Lut und Persisch-Balocistan mit einer Reise durch SüdAfghanistan: Stuttgart, Strecker und Schroder Verlag, 186 p. Gabriel, A., 1942, Die Lut und ihre Wege, Ergebnisse von drei Iranreisen: Frankfurt, Zelschrift für Erdkunde, v. 10, p. 423–442. Gabriel, A., 1952, Die Enforschung Persiens, Die Entwicklung der Abendlandischen Kenntnis der Geographie Persiens [Exploration in Persia, the Development of Western Knowledge of the Geography of Persia]: Vienna, Adolph Holzhausens Nfg., ASIN: B00C0BET4, 359 p. Gabriel, A., 1957a, Zur Oberflächengestalung der Pfannen in der Trockenäumen zentralpersiens: Wien, Mitteilugen Geographischer Gesellschaft, v. 90, no. 2–3, p. 146–160. Gabriel, A. 1957b, Ein Beiträge zur Gliederung und Landschaftskunde des innerpersischen Wüstegurtels: Wien, Geographisches Gesellschaft, Festschrift Hunderjahrfeier, p. 265–278. Gabriel, A., 1961, Die Wüsten der erde un ihre Erforschung: Berlin, SpringerVerlag, 167 p. Gabriel, A., 1963, Marco Polo in Persia: Vienna, Verlag Typographische Anstalt, 235 p. Gabriel, A., 1964, Die Auswirkungen vertikaler Luftströmungen und elektrischer Spannungsfelder in kahlen Sanden: Wien, Mitteilungen der Österreichischen geographischen Gesellschaft, v. 107, no 3, p. 125–127. Gaetani, M., 1964, Rossirhynchus adamantinus gen. n., sp. n., from the Tournaisian of central Elburz, Iran (Rynchonellida): Rivista Italiana di Paleontologia e Stratigrafia, Milano, v. 70, no. 3, p. 637–647. Gaetani, M., 1965, The geology of the Upper Djadjerud and Lar Valleys (north Iran): II. Paleontology; Brachiopods and molluscs from Geirud Formation, Member A (Upper Devonian and Tournaisian): Rivista Italiana di Paleontologia e Stratigrafia, Milano, v. 71, no. 3, p. 679–770. Gaetani, M., 1968, Lower Carboniferous brachiopods from central Elburz, Iran: Rivista Italiana di Paleontologia e Stratigrafia, Milano, v. 74, no. 3, p. 665–732. Gaetani, M., 1978, Riccardo Assereto: Bollettino della Società Geologica Italiana, v. 97, p. 327–328. Gaetani, M., Angiolini, L., Ueno, K., Nicora, A., Stephenson, M.H., Sciunnach, D., Rettori, R., Piece, G.D., and Sabouri, J., 2009, Pennsylvanian–Early Triassic stratigraphy in the Alborz Mountains (Iran), in Brunet, M.-F., Wilmsen, M., and Granath, J.W., eds., South Caspian to Central Iran Basins: Geological Society, London, Special Publication 312, p. 79–128, doi:10.1144/SP312.5. Galoyan, G., Rolland, Y., Sosson, M., Corsini, M., Bilo, S., Verali, C., and Melkonyan, R., 2009, Geology, geochemistry and 40Ar/39Ar dating of Sevan ophiolites (Lesser Caucasus, Armenia): Evidence for Jurassic back-arc opening and hot spot event between south Armenian block and Eurasia: Journal of Asian Earth Sciences, v. 34, no. 2, p. 135–153. Gansser, A., 1952, Geological Reconnaissance in the Kavir and Lut Deserts: Tehran, Iranian Oil Company/National Iranian Oil Company Geological Report 30, Internal unpublished report filed at NIOC. Gansser, A., 1955, New aspects of the geology in Central Iran, in Barbier, A., ed., Proceedings of the 4th World Petroleum Congress, Rome, June 6–15, 1955: Rome, Carlo Colombo publisher, section 1/A/5, paper 2, p. 279–300. Gansser, A., 1960, Ausseralpine ophioloth probleme: Eclogae Geologicae Helvetiae, v. 52, no. 2, p. 659–680. Gansser, A., and Huber, H., 1962, Geological observations in the Central Elbouz, Iran: Schweizerische Mineralogische und Petrographische Mitteilungen, v. 42, no. 2, p. 593–630. Geological Society of America, 2013, Tethyan Evolution and Seismotectonics of Southwest Asia: In Honor of 40 Years of Manuel Berberian’s Research Contributions (session), in Geological Society Abstracts and Programs, v. 45, no. 7, Session 291, p. 259 and 290. Gershevitch, I., 1985, The Cambridge History of Iran, The Median and Achaemenian Periods: Cambridge, UK, Cambridge University Press, v. 2, 946 p.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084715/spe525-03.pdf by University of Wisconsin Madison user

Gershevitch, I., 1995, Approaches to Zoroaster’s Gathas: Iran, British Institute of Persian Studies, v. 33, p. 1–29. Ghirshman, R., 1938–1939, Fouilles de Sialk près de Kashan, 1933, 1934, 1937, Volumes I & 2: Paris, Musée du Louvre, Département des Antiquités Orientales, Série Archeologiques, IV, Librairie Orientaliste Paul Geuthner, 152 p. Ghirshman, R., 1951, L’Iran des origines à l’Islam: Paris, Payot, 330 p. See also English translation: Iran from the Earliest Times to the Islamic Conquest: Harmondsworth, UK, Penguin Books Limited, 368 p. Gibb, H.A.R., 1993, Foreword, in Nasr, S.H., ed., An Introduction to Islamic Cosmological Doctrines; Conceptions of Nature and Methods Used for Its Study by the Ikhwan al-Safa, al-Biruni, and Ibn Sina: Albany, State University of New York Press, p. xiii–xv. Ginter, M., Liano, J.-C., and Valenzuela-Rios, J.-I., 2008, New data on Chondrichthyan microremains from the Givetian of the Renanué section in the Anagonian Pyrenees (Spain): Acta Geologica, v. 58, p. 165–172. Ginter, M., Hampe, O., and Duffin, C.J., 2010, Chondrichthyes Paleozoic Elasmohranchi, teeth, in Schultze, H.-P., ed., Handbook of Paleoichthyology, Volume 3D: Munich, Verlag Dr. Freidrich Pfeil, 168 p. Glaus, M., 1965, Die Geologie des Gebieted Nördlich des Kandevan-Passess (Zentral-Elburz, Iran): Mitteilungen Geological Institute ETH, University of Zürich, n.s., v. 48, 166 p. Gnoli, Gh., 1980, Zoroaster’s Time and Homeland, A Study on the Origins of Mazdaism and Related Problems: Naples, Italy, Istituto Universitario Orientale, 279 p. Gnoli, Gh., 1995, Once more Zoroaster’s time: A Manichean dating: East and West, IsIAO, v. 45(1/40), p. 313–319. Goldsmid, F., 1890, Lieutenant H.B. Vaughan’s recent journey in eastern Persia: Royal Geographical Society Proceedings and Monthly Record of Geography, v. 12, no. 10, p. 577–595. Golombek, L., and Wilber, D., 1988, The Timurid Architecture of Iran and Turan: Princeton, New Jersey, Princeton University Press, 510 p. (volume 1) + 481 p. figs. (volume 2). Greenwood, T., 2011, A reassessment of the life and mathematical problems of Anani Širakac’i: Revue de Études Arméniennes, v. 33, p. 131–186. Grewingk, C., 1853, Die geognostischen und orographischen verhältnisse des nördlichen Persiens: Verhhandl der Russische Mineralogische Gesellschat, St. Petersburg, p. 97–245. Grewingk, C., 1881, Ueber fossile saeugetiere von Maragha in Persien: Wien, Verhandlungen der K. K. Geoligischen Reichsanstalt, 296 p. Griffith, R.T.H., 1896, The Rig Veda, Sacred Writings: Volume 5. Hinduism: New Delhi, Motilal Banarsidass Publishers, PVT, Ltd., 656 p., http:// www.sacred-texts.com/hin/rigveda/index.htm. Guest, B., Horton, B.K., Axen, G.A., Hassanzadeh, J., and McIntosh, W.C., 2007, Middle to late Cenozoic basin evolution in the western Alborz Mountains: Implications for the onset of collisional deformation in northern Iran: Tectonics, v. 26, no. 6, TC6011, doi:10.1029/2006TC002091. Hairapetian, V., Ginter, M., and Yazdi, M., 2008, Early Fransian sharks from Central Iran: Acta Geologica Polonica, v. 58, no. 2, p. 173–179. Hakemi, A., 1972, Catalogue de l’Exposition, Lut, Xabis (Shahdad), Festival de la Culture et des Arts, in Premier Symposium Annuel de la researche archéologique en Iran: Tehran, Iranian Centre for Archaeological Research, 36 p. Hakemi, A., and Sajjadi, S.M.S., 1989, Shahdad excavations in the context of oasis civilization, in Ligabue, G., and Salvatori, S., eds., Bactria, An Ancient Oasis Civilization from the Sands of Afghanistan: Venice, Centro Studi Ricerche Ligabue, p. 141–156. Hallier, U.W., 1972, Fort, Atashgah und Chahar taq von Nakhlak, Überreste eine Sasanidischen Bergbausiedlung, Sonderdruck aus: Archaeologische Mitteilungen aus Iran [Reimer, Berlin]: Neue Folge, v. 5, p. 286–307. Hamédāni, Rashid al-Din Fazlollāh, 1311, Jāme’ al-Tavārikh Rashidi [The History of Rashidi, Karimi, B., ed.]: Tehran, Eqbāl Pul., 1338/1995 [in Persian]; also Rowshan, M., ed., Jāme’ al-Tavārikh (Tārikh Iran va Eslam): Tehran, Markaz Pazhuhesh Mirāth Maktub, 1392/2013, 3 volumes [in Persian]. Hansman, D.B., 1978, The question of Aratta: Journal of Near Eastern Studies, Chicago, v. 37, p. 331–336. Harrison, J.V., 1930, The geology of some salt plugs in Laristan, south Persia: Quarterly Journal of the Geological Society, London, v. 86, p. 463–522. Harrison, J.V., and Falcon, N.L., 1934a, The effect of gravity on rocks in weathered simple folds; a factor in tectonics [abs.]: Report of the British Association for the Advancement of Science, 1934 Report, p. 303.

Development of geological perceptions and explorations on the Iranian Plateau Harrison, J.V., and Falcon, N.L., 1934b, Collapse structure (Kuhgalu district, Persia): Geological Magazine, v. 71, no. 864, p. 529–539. Harrison, J.V., and Falcon, N.L., 1935, Gravity collapse structures and mountain ranges, as exemplified in southwestern Persia [abs.]: Proceedings of the Geological Society of London, no. 1299, p. 110–112. Harrison, J.V., and Falcon, N.L., 1936a, Geological Map, Coastal Makran: Tehran, filed at the Iranian Oil Company/National Iranian Oil Company, unpublished Internal Report. Harrison, J.V., and Falcon, N.L., 1936b, Gravity structures and mountain ranges, as exemplified in southwest Persia (with discussion): Quarterly Journal of the Geological Society, London, v. 92, no. 365, part 1, p. 91–102. Harrison, J.V., and Falcon, N.L., 1937a, The Saidmarreh landslip: Al-Naft Magazine, v. 13, no. 2, p. 2–8. Harrison, J.V., and Falcon, N.L., 1937b, The Saidmarreh landslip, southwest Iran: Geographical Journal, v. 89, no. 1, p. 42–47. Harrison, J.V., and Falcon, N.L., 1938, An ancient landslip at Saidmarreh in southwest Iran: Journal of Geology, v. 46, no. 3, pt. 1, p. 296–309. Harrison, J.V., 1943, The Jaz Murian Depression, Persian Baluchistan: The Geographical Journal, v. 101, p. 206–225. Harrison, J.V., 1968, Geology, in Fisher, W.B., ed., The Cambridge History of Iran: Volume I. The Land of Iran: Cambridge, UK, Cambridge University Press, p. 111–185. Harrison, J.V., Falcon, N.L., Allison, A., Hunt, J.A., Maling, B.P., and McCall, R.J.C., 1935–1936, Geological Map, Landward Makran, South Iran: Tehran, Iranian Oil Company/NIOC, Internal Report filed at NIOC. Hässig, M., Rolland, Y., Sosson, M., Galoyan, G., Müller, C., Avagyan, A., and Sahakyan, L., 2013, New structural and petrological data on the Amasia ophiolites (NW Sevan-Akera suture zone, Lesser Caucasus): Insights for a large-scale obduction in Armenia and NE Turkey: Tectonophysics, v. 588, p. 135–153. Hau, F.R., 1979, Gondeschapur - eine Medizinschule aus dem 6, Jahrhundert n. Chr: Zürich Gesnerus Swiss Journal of History of Medecine and Sciences, Ein Dienst der ETH-Bibliothek, v. XXXVI, p. 98–115. Hauqal (Haukal), ebn, 1800, Ketāb Masālek wa al-Mamālek [The Oriental Geography of Ebn Haukal, an Arabian Traveller of the Tenth Century, translated by W. Ouseley]: London, T. Cadell, Jun, and W. Davies, Strnad, chapter T1, Of the Province of Kirman, p. 138–146. Hautmann, M., 2001, Taxonomy and phylogeny of cementing Triassic bivalves (Family Prospondylidae, Plicatulidae, Dimydae and Ostreidae): Palaeontology, v. 44, no. 2, p. 339–373. Henning, W.B., 1951, Zoroaster, Politician or Witch-Doctor?: Oxford, UK, Oxford University Press, Ratanbai Katrak Lectures 1949, 51 p. Heratsi, M., 1184, Consoling in Fevers: Yerevan, Publishing House of the Academy of Sciences of the Armenian SSR, Ohanessyan, L., tr., 1955, 267 p. [in Armenian]. Hewsen, R.H., 1968, Science in seventh-century Armenia, Ananias of Širak: Isis, v. 59, no. 1, p. 34. Hewsen, R.H., 2001, Armenia, A Historical Atlas: Chicago, Illinois, University of Chicago Press, 342 p. Hinnells, J.R., 1975, Persian Mythology: New York, The Hamlyn Publishing Group Ltd., second edition, 141 p., ISBN: 0-600-03090-3. Holmes, A., 1944, Principles of Physical Geology: London, Thomas Nelson and Sons, 532 p. Holmyard, E.J., and Mandeville, D.C., 1927, Avicennae de Congelatione et Conglutinatione Lapidum, being sections of the Kitāb al-Shifā, The Latin and Arabic texts edited with an English translation of the latter and with critical notes: Paris, Librairie Orientaliste, Paul Geuthner, 86 p. Horton, B.K., Hassanzadeh, J., Stockli, D.F., Axen, G.J., Gillis, R.J., Guest, B., Amini, A., Fakhari, M.D., Zamanzadeh, S.M., and Grove, M., 2008, Detrital zircon provenance of Neoproterozoic to Cenozoic deposits in Iran: Implications for chronostratigraphy and collisional tectonics: Tectonophysics, v. 451, p. 97–122. Huber, H., 1952, Geology of the Western Coastal Makran Area: Tehran, filed at the Iranian Oil Company/National Iranian Oil Company, unpublished Internal Report GR 91B. Huber, H., 1960, Geological Map of Iran, with Explanatory Note: Tehran, National Iranian Oil Company, scale 1:2,500,000. Huber, H., 1978, Geological Map of Iran, in Six Sheets with Explanatory Notes: Tehran, National Iranian Oil Company Exploration and Production, scale 1:1,000,000.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084715/spe525-03.pdf by University of Wisconsin Madison user

79

Huckriede, M., Kürsten, H., and Venzlaff, H., 1962, Zur geologie des gebiets zwischen Kerman und Saghand (Iran): Beihefte zum Geologischen Jahrbuch, v. 51, 197 p. Hunt, P., 2008, Archaemenid Persian Griffin Capital at Persepolis: Stanford, California, Archaeology, Stanford University, http://web.stanford.edu/ dept/archaeology/cgi-bin/archaeolog/?p=225 (21 October 2008). Hushmandzadeh, A., and Berberian, M., 1972, Implication of Plate Tectonics for the Sanandaj-Sirjan Zone (Northeastern Border of Zagros): Tehran, Geological Survey of Iran, Internal Report, 6 p. Hushmandzadeh, A., and Berberian, M., 1973, The Relation of Orogenic Movements to Metamorphic Episodes in Iran—A Short Review: Tehran, Geological Survey of Iran, Internal Report, 3 p. Hushmandzadeh, A., Sabzehei, M., and Berberian, M., 1972, A Brief Note on the Early Kimmerian Orogeny and High Grade Metamorphism in the Sanandaj-Sirjan Zone (Sirjan-Esfandagheh) Iran: Tehran, Geological Survey of Iran, Internal Report, 2 p. Hushmandzadeh, A., Majidi, B., Alavi, N., and Berberian, M., 1973, Study of deformational and metamorphic history of Shahr Kord-Sanandaj zone, in First Iranian Geological Symposium, Proceedings: Tehran, Iran Petroleum Institute, 25 p. Hutt, A., 1984, Iran, in Michel, G., ed., Architecture of the Islamic World: New York, Thames and Hudson, p. 251–258. Hutt, A., and Harrow, L., 1977, Islamic Architecture, Iran 1: London, Scorpion Publications Ltd., 192 p. Hutt, A., and Harrow, L., 1978, Islamic Architecture, Iran 2: London, Scorpion Publications Ltd., 177 p. Ibn Khaldun, ‘Abd ar-Rahmān bin Mohammed, 1377, Muqaddimah [An Introduction to History], Rosenthal, F., tr., 1958: http://www.muslimphilosophy.com/ik/Muqaddimah/ (accessed 2014). Also Moqaddameh-ye Ebn Khaldun, Parvin Gonābādi, M., tr.: Bongāh Tarjomeh va Nashr-e Ketāb, Tehran, 1337/1958 [in Persian]. International Astronomical Union (IAU), 2015a, Gazetteer of Planetary Nomenclature: International Astronomical Union (IAU) Working Group for Planetary Nomenclatures (WGPSN), http://planetarynames.wr.usgs .gov/ (accessed 2015). International Astronomical Union (IAU), 2015b, Minor Planet Center: The International Astronomical Center (IAU), Minor Planet Discoveries: minorplanetcenter.org (accessed 2015). Italconsult, 1959–1960, Plan Organisation for Iran-Baluchestan: Tehran, filed at the Geological Survey of Iran, unpublished Internal Report. Jackson, M.P.A., Cornelius, R.R., Craig, C.H., Gansser, A., Stocklin, J., and Talbot, C.J., 1990, The Salt Diapirs of the Great Kavir, Central Iran: Geological Society of America Memoir 117, 139 p. Jackson, P., and Lockhart, L., 1986, The Cambridge History of Iran: Volume 6. The Timurid and Safavid Periods: Cambridge, UK, Cambridge University Press, 1066 p. Jamali, H., Dilek, Y., Daliran, F., Yaghubpur, A., and Mehrabi, B., 2010, Metallogeny and tectonic evolution of the Cenozoic Ahar-Arasbaran volcanic belt, northern Iran: International Geology Review, v. 53, p. 608–630. Jamali, H., Yaghubpur, A., Mehrabi, B., Dilek, Y., Daliran, F., and Meshkan, A., 2012, Petrogenesis and tectono-magmatic setting of Meso-Cenozoic magmatism in Azerbaijan Province, northwestern Iran, in al-Juboury, A., ed., Petrology—New Perspectives and Applications: InTech, p. 39–55, doi:10.5772/24782, ISBN: 978-953-307-800-7. James, G.A., and Wynd, J.G., 1965, Stratigraphic nomenclature of Iranian Oil Consortium Agreement area: American Association of Petroleum Geologists Bulletin, v. 49, no. 12, p. 2182–2245. Janvier, P., and Martin, M., 1978, Le vertebres de l’Iran central: I. Dipneustes [Devonian vertebrates of central Iran: I. Dipneustes]: Geobios, v. 11, no. 6, p. 819–833. Jaubert, A., 1821, Voyage en Arménie et en Perse, fait dans les années 1805 et 1806, suivi d’un mémoire sur le Ghilan et le Masenderan: Paris, Pélicier, 449 p. Joneidi, F., 2008, Shāhnāmé Ferdowsi: Tehran, Nashr Balkh/Bonyād Neyshābur Publ., 6 volumes, 1387 (original publication date) [in Persian]. Kanga, E.M.F., ed., 2001, Yasht-ba-Manā of Ervad Kavasji Edulji Kanga: Mumbai, India, Jenaz Printers, 336 p. Karaji, Abu Bakr Mohammad ibn al-Hassan al-Hasseb, 1010, Estekhrāj Ābhāyeh Penhāni [The Extraction of Hidden Waters (groundwater) to the Surface, translated by H. Khadiv-Djam, 1345/1966]: Tehran, Enteshārāt Bonyād Farhang Irān [Iran Cultural Foundation] no. 8, Science in Iran no. 2. Also Groundwater Exploration, The Oldest Hydrology Text Book,

80

M. Berberian

by Aboubakr Karaji (translated by Nour Zaghi); online at http://www .uscashagent.com/groundwater-exploration.htm (June 2007). Karpova, K.N., Kon’kova, E.A., Larkin, E.D., and Savel’ev, V.F., 1958, Avicennite, a new mineral: Doklady Akademii Nauk Uzbekistan SSR, v. 2, p. 23–26 [in Russian]. Kavoukjian, M., 1987, Armenia, Subartu, and Sumer; The Indo-European Homeland and Ancient Mesopotamia (translated by N. Ouzounian): Montreal, Canada, M. Kavoukjian, Inc., 243 p. Kazmin, V.G., Sbortshikov, I.M., Ricou, L.-E., Zonenshain, L.P., Boulin, J., and Knipper, A.L., 1986, Volcanic belts as markers of the Mesozoic–Cenozoic active margin of Eurasia: Tectonophysics, v. 123, p. 123–152. Kedar, B.Z., and Wiesner-Hanks, M., 2015, The Cambridge World History, volume 5, Expanding Webs of Exchange and Conflict, 500 CE-–1500 CE: Cambridge, UK, Cambridge University Press, 748 p., ISBN 978-0-52119074-9. Kellens, J., 2006, La Quatrième Naissance de Zarathushtra: Seuil, Paris, La Librairie du XXIe Siècles, 188 p. Kennedy, E.S., 1973, A Commentary upon Binurn’s Kitāb Tahdid Al-Amākin: Beirut, Lebanon, American University of Beirut, Lebanon, 270 p. Khadiv-Djam, 1966. See Karaji, 1010. Khalatbari Jafari, M., Babaie, H.A., and Gani, M., 2013, Geochemical evidence for Late Cretaceous marginal arc-to-backarc transition in the Sabzevar ophiolitic extrusive sequence, northeast Iran: Journal of Asian Earth Sciences, v. 70–71, p. 209–230. Khaldun, I., al-Haārami, Abd al-Rahmān bin Muhammad, 1377, The Muqaddimah [The Introduction: An Introduction to History (Bollingen Series, XLIII), translated by Franz Rosenthal, 1958, Volume 3]: New York, Pantheon Books, 512 p. Also https://asadullahali.files.wordpress.com/2012/10/ibn _khaldun-al_muqaddimah.pdf,http://www.muslimphilosophy.com/ik/ Muqaddimah/. Also Moghaddameh ebn Khaldun [An Introduction to History, translated by M. Parvin Gonabadi]: Tehran, Bongāh Tarjomeh va Nashr Ketāb Publishing, 2 vols., 1337/1958, 1430 p. Khanikoff, N., 1858 (in Abich, H., 1858), Tremblement de terre observé à Tabriz en Septembre 1856, Notices physiques et géographiques de M. Khanykof sur l’Azerbeidjan, communiquées par H. Abich (lu le 17 Janvier, 1857): Bulletin de Class Physico-Mathématique de l’Académie Impériale des Sciences de Saint-Pétersbourg, v. 16, no. 22, p. 337–352. Khanikoff, N., 1861 (in Abivh, H., 1861), Mémoire sur la Partie Méridionale de l’Asie Centrale: Paris, Imprimerie de L. Martinet, 234 p. Kittl, E., 1885, Die fossile saugethier-fauna von Maragha in Persien: Verhandlungen der K.K.: Vienna, Geologischen Reichsanstalt, v. 1185, p. 397–399. Kon’kova, E.A., and Savel’ev, V.F., 1960, A new thallium mineral—Avicennite: Zapiski Vsesesoyuznogo Mineralogicheskoo Obshchchestva, v. 89, p. 316–320 [in Russian]. Kramer, S.N., 1972, The Sumerians, Their History, Culture, and Character: Chicago, Illinois, University of Chicago Press, 372 p. Krenkow, F., 1946, The Drug-Book of Beruni: Islamic Culture, v. 20, p. 109–110. Krenkow, F., 1961, ed. See Biruni, 1025. Kreyenbroek, Ph.G., 1993, Cosmogony and cosmology in Zoroastrianism/ Mazdaism: Encyclopaedia Iranica, v. V, no. 3, p. 303–307 (www .iranicaonline.org, last updated 2011). Kummel, B., 1972, Memorial to John Vernon Harrison 1892–1972: Geological Society of America, p. 87–90, ftp://rock.geosociety.org/pub/Memorials/ v04/Harrison-JV.pdf. Ladame, G., 1945, Les ressources métallifères de l’Iran: Schweizische Mineralogische und Petrographische Mitteilungen, v. 25, no. 1, p. 165–303. Laurent, H., 1967, Prospections Hyrogéologique de la Plaine de Karadj à l’Ouest de Tehran, Iran [Thèse de Doctorat]: Paris, Université de Paris, 187 p. Lees, G.M., and Richardson, F.D.S., 1940, The geology of the oilfield belt of southwest Iran and Ilam: Geological Magazine, v. 77, p. 227–251. Lehner, P., 2009, Pierre Albert Soder obituary: Swiss Bulletin Angewwandte Geologie, v. 14, heft 1–2, p. S.189–190. Lescuyer, J., and Riou, L.R., 1976, Géologie de la Region de Mianeh (Azerbaijan), Contribution à l’Étude du Vulcanisme Tertiare de l’Iran [Thèse 3em cycle]: Grenoble, France, Institut Dolomieu, University of Grenoble, 233 p. Le Strange, G., 1905, The Lands of the Eastern Caliphate; Mesopotamia, Persia, and Central Asia form the Moslem Conquest to the Time of Timur: Cambridge, UK, Cambridge University Press, and London, Frank Gass. Also translated by M. Erfān: Tehran, Publisher Sherkat Enteshārāt

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084715/spe525-03.pdf by University of Wisconsin Madison user

‘Elmi va Farhangi, v. 74, 1337/1958 (1st ed.), 1364/1985 (2nd ed.), 610 p. [in Persian]. Lockhart, L., 1939, Iranian petroleum in ancient and medieval times: Journal of Institute of Petroleum Technology, v. 25, p. 1–18. Loftus, W.K., 1855, On the geology of portions of the Turko-Persian frontier, and the districts adjoining: Quarterly Journal of the Geological Society, London, v. 11, no. 1, p. 247–344. Lorenz, C., 1964, Die Geologie des oberen Karadj-Tales (Zentral-Elburz), Iran [Doctoral thesis]: Mitteilungen Geological Institute ETH, University of Zürich v. 22, 113 p. Lydekker, R., 1886, On the fossil mammalia of Maragha in northwest Persia: Quarterly Journal of the Geological Society, London, v. 42, p. 173–176. Magee, P., 2005, The chronology and environmental background of Iron Age settlement in southeastern Iran and the question of the origin of the Qanat irrigation system: Iranica Antiqua, v. XL, p. 217–231. Majd, Gh., T., 1994, Historical Treaties and Contracts During the Qājār Period: Tehran, Bonyād Moqufāt Dr. Mahmud Afshār Publishing, 520 p. [in Persian]. Majidzadeh, Y., 1976, The Early Prehistoric Cultures of the Central Plateau of Iran; An Archaeological History of Its Development during the Fifth and Fourth Millennium B.C. [Ph.D. dissertation]: Chicago, Illinois, University of Chicago, 500 p. Malandra, W.W., 2009, Zoroaster ii, General Survey: http://www.iranicaonline .org/articles/zoroaster-ii-general-survey (accessed 2014). Malcom, J., 1827, Sketches of Persia, From the Journals of a Traveller in the East, 1927: London, John Murray, v. 1, 278 p., v. 2, 281 p. Also at archive.org. Mansuri, F., 1989, Qanāt and its influence in civilization and culture [Qanāt va Ta’sir-e on dar Tamadon va Farhan], in Afshār, I., and Esfahāniān, K., eds., Memorial of Dr. Mahmud Afshār, Volume 5: Tehran, Enteshārāt Bonyād Mowqufāt Dr. Mahmud Afshār, University of Tehran, p. 2532– 2561, 1368/1989 [in Persian]. McCall, G.J.H., 1985, Area Report, East Iran Project, Area No. 1: Tehran, Geological Survey of Iran, Report no. 57, 634 p. McCall, G.J.H., 1997, The geotectonic history of the Makran and adjacent areas of southern Iran: Journal of Asian Earth Sciences, v. 15, p. 517–531. McCall, G.J.H., 2003, A critique of the analogy between Archaean and Phanerozoic tectonics based on regional mapping of the Mesozoic-Cenozoic plate convergent zone in the Makran, Iran: Precambrian Research, v. 127, p. 5–17. McCall, G.J.H., and Kidd, R.G.W., 1992, The Makran, southeastern Iran, the anatomy of a convergent plate margin active from the Cretaceous to Present, in Leggett, J.K., ed., Trench-Fore-Arc Geology: Geological Society, London, Special Publication 10, p. 387–397. McKenzie, D.P., 1970, Plate tectonics of the Mediterranean region: Nature, v. 226, p. 277–290. McKenzie, D.P., Davies, D., and Molnar, P., 1970, Plate tectonics of the Red Sea and East Africa: Nature, v. 226, p. 243–248. Mecquenem, R. de, 1905, Le gisement de vertebres fossils de Maragha: Compte Rendu Académie des Sciences de Paris, v. 141, p. 285–286. Mecquenem, R. de, 1906, Les vertébrés fossiles de Maragha: Nature, v. 24, p. 10. Mecquenem, R. de, 1908, Contribution à l’étude du gisement de vertébrés de Maragha et de ses environs: Annales d’Histoire Naturelle (Délégation en Perse), Paléontologie, v. 1, p. 27–80. Mecquenem, R. de, 1911, Contribution à l’étude du gisement des vertébrés fossils de Maragha: Nature de Maragha, deuxième partie: Annales d’Histoire Naturelle, v. 1, p. 27–28. Mecquenem, R. de, 1924–1925, Contribution à l’études des fossiles de Maragha I: Annales de Paléontologie, v. 13, p. 133–160. Mehryar, M., and Kabiri, A., 1986, Preliminary report on the Delazian (Cheshmeh Shaikh) archaeological site: Tehran, Āthār, Sāzeman Melli Hefazat Āthār Bāstāni Irān, no. 12–13–14, p. 3–46, Esfand 1365 [in Persian]. Melville, C., 2008, Jame al-Tavarikh: Encyclopedia Iranica, v. XIV, fasc. 5, p. 462–468; www.iranicaonline.org (2012). Meyer, S.P., 1967, Die geologie des gebietes Velian-Kechire (Zentral-Elburz), Iran [Doctoral thesis]: Mitteilungen Geological Institute (ETH), University of Zürich, v. 79, 127 p. Milanovsky, E.E., 2007, Hermann Abich (1806–1886): The father of Caucasus geology and his travels in the Caucasus and Armenian Highlands, in Wyse Jackson, P.N., ed., Four Centuries of Geological Travel, The Search for Knowledge on Foot, Bicycle, Sledge, and Camel: Geological Society, London, Special Publication 287, p. 177–181.

Development of geological perceptions and explorations on the Iranian Plateau Minorsky, V., 1942, Sharaf al-Zamān Tahir Marvazi on China, the Turks, and India: London, Royal Asiatic Society, James D. Forlung Fund, v. XXII, p. 170. Moayyed, M., 2001, Geochemistry and Petrology of Volcano-Plutonic Bodies in Tarum Area [Thesis]: Tabriz, Iran, Tabriz University, 256 p. [in Persian]. Moorey, F.R.S., 1994, Ancient Mesopotamian Material and Industries: Oxford, UK, Clarendon Press, 436 p. Morgan, J. de, 1892, Note sur les gites de naphte de Kend-e Chirin: Annales Mines ou Recueil, n.s., Memoire, v. 1, p. 227–238. Morgan, J. de, 1894, Mission Scientifique en Perse: Volume 1. Études Géographiques: Paris, E. Leroux, 328 p. Morgan, J. de, 1895, Mission Scientifique en Perse: Volume 2. Études Géomorphiques: Paris, E. Lerouz, 428 p. Morgan, J. de, 1900, Notes sur la basse-Mesopotamie, Comblement Progressif du Golfe Persique: Paris, Geographie, v. 2, p. 274–262. Morgan, J. de, 1905a, Mission Scientifique en Perse: Volume 3, Part 1. Geologie Stratigraphique: Paris, E. Leroux, 136 p. Morgan, J. de, 1905b, Note sur la géologie de la Perse et sur les travaux paléontologiques de M. H. Douvillé sur cette region: Bulletin de la Société Géologique de France, ser. 4, v. 5, no. 2, p. 170–189. Morgan, J. de, 1907, Le plateau Iranien pendant l’époque Pléistocène: Paris, Révue de l’École d’Anthropologie v. 17, p. 213–216. Mostofi, B., and Gansser, A., 1957, The story behind the 5 Alborz: Oil & Gas Journal, v. 56, p. 78–84. Murray, C.A., 1859, Notes on some mineral springs near Damavand, and on some copper ore and other minerals from near Tabriz: Quarterly Journal of the Geological Society, London, v. 15, p. 198 and 605. al-Nadim, I., Muhammad b. Eshāq ibn, 988, The Fihrist of al-Nadim, A TenthCentury Survey of Muslim Culture, 377/988 (Dodge, B., trans. ed.): New York, Columbia University Press, p. 568–633, ISBN: 321-02925-X, online at muslimphilosophy.com/books/nad-phil.pdf. Nadji, M., and Voight, R., 1972, Exploration for Hidden Water by Mohammad Karaji—The oldest textbook on hydrology?: Ground Water, v. 10, no. 5, p. 43–48, doi:10.1111/j.1745-6548.1972.tb03592.x. Nasr, S.H., 1963, Nazar Motefakerān Eslāmi dar bārehye Tabi’at [Conception of Nature in Islamic Thought], 1342/1963 (1st ed.); 1359/1980 (3rd ed.): Tehran, Khārazmi Publishers, 470 p. [in Persian]. Nasr, S.H., 1964, Three Muslim Sages; Avicenna-Suhrawardi-Ibn ‘Arabi: New York, Caravan Books, Delamr, 185 p. Nasr, S.H., 1966, Sé Hakim-e Mosalmān [Three Muslim Sages, translated by A. Ārām] 1st ed. 1345/1966; 2nd ed. 1352/1973; 3rd ed. 1354/1975; 4th ed. 1361/1982: Tehran, Sherkat Sahami Ketābhayeh Jibi, 239 p. [in Persian]. Nasr, S.H., 1968, Science and Civilization in Islam: Cambridge, Massachusetts, Harvard University Press, 384 p. Nasr, S.H., 1974, Religion in Safavid Persia: Iranian Studies, v. 7, no. 1/2, p. 271–286. Nasr, S.H., 1976, Islamic Science; An Illustrated Study: Kent, World of Islam Festival Publishing Co., Ltd., 273 p., ISBN: 0-905035-02-X. Nasr, S.H., 1980, ‘Elm va Tamadon dar Eslām (translated by A. Ārām): Tehran, Khārazmi Publishers, 1359/1980 [in Persian]. Nasr, S.H., 1993, An Introduction to Islamic Cosmological Doctrines; Conceptions of Nature and Methods Used for Its Study by the Ikhwan alSafa, al-Biruni, and Ibn Sina: Albany, State University of New York Press, 322 p. Natal’in, B.A., and Şengör, A., 2005, Late Palaeozoic to Triassic evolution of the Turan and Scythian platforms: The pre-history of the Palaeo-Tethyan closure: Tectonophysics, v. 404, no. 3–4, p. 175–202. Nicholson, R.A. (trans. and ed.), 1926, The Mathnawi of Jalālu’ddin Rumi: Cambridge, The E.J.W. Gibb Memorial Trust Series 4.2–4.4, vols., v. 1, 496 p., dar-al-masnavi.org. Nylander, C., 1970, Ionians in Pasargadae; Studies in Old Persian Architecture: Uppsala, Sweden, Acta Universitatis Upsaliensis, Boreas, Uppsala Studies in Ancient Mediterranean and Near Eastern Civilizations 1, 176 p. O’Kane, R., 1987, Timurid Architecture in Khurasan: Costa Mesa, California, Mazdā Publishers, 418 p., ISBN 0-939214-35-0. Ohanessyan, L., tr., 1955. See Herātsi, M., 1184. Omrani, H., Moazzen, M., Oberhänsli, R., Altenberger, U., and Lange, M., 2013, The Sabzevar blueschists of the north-Central Iranian microcontinent as remnants of the Neotethys-related oceanic crust subduction: Geologische Rundschau, v. 102, p. 1491–1512. Omstead, A.T., 1948, History of the Persian Empire: Chicago, Illinois, The University of Chicago Press, 568 p.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084715/spe525-03.pdf by University of Wisconsin Madison user

81

Otte, J.K., 1988, Alfred of Sareshel’s Commentary on the Metheora of Aristotle, Critical Edition, Introduction, and Notes: Leiden, Netherlands, E.J. Brill, 131 p. Ovtsinnikov, A., 1930, Outline of the geology of the Demavend region (north Iran): Moscow, Bulletin de la Société des Naturalistes, Section Géologie, v. 7, no. 3–4, p. 373–400. Pazwash, H., and Mavrigian, G., 1980, A historical jewelpiece—Discovery of the millennium hydrological works of Karaji: Water Resources Bulletin: American Water Resources Association, v. 16, no. 6, p. 1094–1096. Perrot, J., 2013, The Palace of Darius at Suza; The Great Royal Residence of Achaemenid Persia: London, I.B. Tauris & Co Ltd., 544 p., ISBN: 9781848856219. Perry, E., 2010, The rise of Shi’ism in Iran: Cross-sections, v. VI, p. 121–136. Peterson, J.P., 2002, Zand-Akasih, digital version of the Iranian or Greater Bundahishn, translated by Ankelsaria, B.T., 1956, online at http://www .avesta.org/mp/grb.htm (accessed 2014). Pilgrim, G.E., 1908, The Geology of the Persian Gulf and Adjoining Portions of Persia and Arabia: Memoirs of the Geological Survey of India, v. 34, no. 4, 177 p. Pilgrim, G.E., 1922, The sulphur deposits of southern Persia: Records of the Geological Survey of India, v. 53, no. 4, p. 343–358. Pilgrim, G.E., 1924, The Geology of Parts of the Persian Provinces of Fars, Kirman, and Laristan: Memoirs of the Geological Survey of India, v. 48, no. 2, 118 p. Pliny the Elder, ca. 79, Naturalis Historia [The Natural History]; Book XXXVI, The Natural History of Stones; Book XXXVII, The Natural History of Precious Stones (Bostock, J. and Riley, T.H., eds.): London, Taylor and Francis, 1855, 6 volumes; also http://www.perseus.tufts.edu/hopper/ text?doc=Plin.+Nat.+toc. Pohlig, H., 1886, On the Pliocene of Maragha, Persia and its resemblance to that of Pikermi in Greece; on fossil elephant remains of Caucasus and Persia: Quarterly Journal of the Geological Society, London, v. 42, p. 177–182. Polak, J.E., 1865, Persien; das Land und seine Bewohner: Leipzig, F.A. Brodhaus v. II, p. 174–178. Also archive.org/details/persien00polagoog. Pope, A.U., 1938, A Survey of Persian Art from Prehistoric Times to the Present: Oxford, UK, Oxford University Press, 6 volumes; reprinted by Mazdā Publishers, Costa Mesa, California, 16 volumes, 3816 p. Pope, A.U., 1965, Persian Architecture, A Triumph of Form and Color: New York, George Braziller, 288 p. Popov, L., Ghobadipour, M., and Hosseini, M., 2008, Early to Middle Ordovician lingulate brachiopods from the Lashkarak Formation, eastern Alborz Mountain Range: Association of Australasian Palaeontologists: Alcheringa, v. 32, p. 1–32. Potts, D.T., 1989, Seleucid Karmania, in De Meyer, L., and Haerinck, E., eds., Archaeologia Iranica et Orientalis, Miscellanea in Honorem Louis Vanden Berghe, Volume II: Leuven, Belgium, Peeters Press, p. 581–603. Poure Davoud (Purdāvud), E., tr., 1977a, The Yashts, Volume 1 (3rd ed.): Tehran, Tehrān University Publisher, Publ. no. 1637, 626 p. [in Persian]. Poure Davoud (Purdāvud), E., tr., 1977b, The Yashts, Volume 2 (3rd ed.): Tehran, Tehrān University Publisher, Publ. no. 1638, 545 p., originally published 1356 [in Persian]. Rashid al-Din Fazlollāh Hamédāni, 1304, Jame’ al-Tavārikh Rashidi (Karimi, B., ed.): Eqbāl Publishing, 2 volumes, Mongols, 1338/1959, Terhan [in Persian]. Also Quatremére, E., tr. and ed., 1836, Histoire des Mongols de la Perse écrite en Persan par Raschieldin: Paris, Imprimé Royale. Also Boyle, J.A., tr., The Successor of Genghis Khan (The History of Rashid ad-Din at-Tabib), 1971: New York, Columbia University Press. Also Thackston, W.M., tr., 1998, Rashiduddin Fazlullah’s Jami u’t-Tawarikh [Compedium of Chronicles]: Cambridge, Massachusetts, Harvard University Press, Volume 1. Rashidi, K., and Senowbari-Daryan, 2011, Sponges from a section of the Upper Triassic Nayband Formation, northeast of Esfahan, central Iran: Annallen des Naturhistorischen Museums Wien, Serie A, v. 113, p. 309–371. al-Rawi, M.M., 2002, The contribution of Ibn Sina (Avicenna) to the development of earth science: Manchester, UK, Foundation for Science, Technology, and Civilization (FSTC) Ltd., no. 4039, 12 p. Also muslimheritage .com/upload/ibnsina.pdf; muslimphilosophy.com/sina. Razavi, M., 1984, Tansukh-Nāmeh Ilkhāni Khājeh Nasir al-Din Tusi 1256–59: Tehran, Etela’at Publishing, 1363/1884, 358 p. [in Persian]. Reza, ‘A., Kuros, Gh.R., Emam Shustari, M.’A., and Entezami, ‘A.A., 1971, Water Resources and Irrigation Technique in Ancient Iran: Tehran, Ministry of Water and Power, 309 p. [in Persian].

82

M. Berberian

Richardson, R.K., 1926, Die geologie und die salzdome im südwetlichen teile des Persischen Golfes: Verhandlunger der Naturhistorischen.-Medzin Verein Heidelberg, n.s., no. 15, 51 p. Ricou, L.E., 1971, Le croissant ophiolitique péri-Arabe, une ceinture de nappes mises en place au Crétace supérieur: Révue de Géographie Physique et de Géologie Dynamique 2, v. 13, no. 4, p. 327–350. Rieben, H., 1930, Geological description of the district around Tabriz (and Yam), in Gilliath-Smith, B., and Turrill, W.B., eds., A Contribution to Our Knowledge of the Flora of Azerbaidjan, N. Persia: Bulletin of Miscellaneous Information, Royal Botanic Gardens, Kew, v. 7, p. 274–277. Rieben, H., 1935, Contribution à la géologie de l’Azerbeidjan Persan: Bulletin de Société Naturelle, Neuchâtel, v. 59, no. 1934, p. 20–144. Rieben, H., 1942, Notes sur la géologie du nord de l’Iran: Elat à Ebolowa, Cameroon, Halsey Memorial Press, American Presbyterian Mission, 17 p. Rieben, H., 1953a, Note préliminaire sur les terrains alluviaux de Téhéran et particulièrement du territoire de Shemran: Bulletin de la Société Vaudoise des Sciences Naturelles, v. 65, no. 283, p. 377–388. Rieben, H., 1953b, Contribution à la géologie de l’Azerbeidjan Persan: Bulletin de la Société Neuchâteloise des Sciences Naturelles, Neuchâtel, Imprimerie Centrale, 144 p. Rieben, H., 1955, The geology of the Tehran plain: American Journal of Science, v. 253, p. 617–639. Rieben, H., 1966, Geological Observations on Alluvial Deposits in Northern Iran: Geological Survey of Iran, Report no. 9, 40 p. Rigvedā. See Griffith, 1896. Rivière, A., 1930, Sur une coupe observé dans la callée moyenne du Djeredjeroud (Perse): Comptes Rendus des Séances de l’Académie des Sciences de Paris, v. 12, p. 759–760. Rivière, A., 1931a, Sur la région de Demavend et de roudehen (Elbourz, Perse): Compte Rendu Sommaire des Séances de la Société Géologique de France, v. 1931, no. 14, p. 115–116. Rivière, A., 1931b, Sur la chaîne de Zarinkouh: Compte Rendu Sommaire des Séances de la Société Géologique de France, v. 1931, 103 p. Rivière, A., 1931c, Tectonique de la bordure nord basin tertiaire inferieur de l’Elbourz central, de la vallée du Djerdjeroud à Roudehen: Compte Rendu Sommaire des Séances de la Société Géologique de France, v. 1931, no. 7, p. 91–92. Rivière, A., 1931d, De la structure de la montagne de Schimran (Elbourz, Perse): Compte Rendu Sommaire des Séances de la Société Géologique de France, v. 1931, no. 7, 91 p. Rivière, A., 1931e, De deux espèces nouvelles dans l’Elbourz central: Compte Rendu Sommaire des Séances de la Société Géologique de France, v. 1931, no. 7, p. 73–89. Rivière, A., 1931f, De l’existence de deversements vers le nord dans le sud de l’Elbourz central: Compte Rendu Sommaire des Séances de la Société Géologique de France, v. 1931, no. 14–15, p. 133–134. Rivière, A., 1931g, Du basin du haut Djerdjeroud (Elbourz, Perse): Compte Rendu Sommaire des Séances de la Société Géologique de France, v. 1931, p. 214–215. Rivière, A., 1931h, Remarque sur la stratigraphie de la base de l’Eocene dans l’Elbourz central: Comptes Rendus des Séances de l’Académie des Sciences de Paris, v. 192, no. 20, p. 1246–1247. Rivière, A., 1932a, Contribution à l’étude du Jurassique dans l’Elbourz central: Comptes Rendus des Séances de l’Académie des Sciences de Paris, v. 195, p. 540–542. Rivière, A., 1932b, La transgression du Cretace dans le nord de l’Elbourz central: Comptes Rendus des Séances de l’Académie des Sciences de Paris, v. 195, p. 551–552. Rivière, A., 1932c, Contribution à l’étude du Paleozoique de l’Elbourz central: Comptes Rendus des Séances de l’Académie des Sciences de Paris, v. 195, p. 811–813. Rivière, A., 1932d, Sur l’existence d’importants mouvements horizontaux dans le nord de l’Elbourz central: Compte Rendu Sommaire des Séances de la Société Géologique de France, v. 1932, no. 16, p. 226–228. Rivière, A., 1932e, Description d’une espèse nouvelle d’Orbitolinae: Compte Rendu Sommaire des Séances de la Société Géologique de France, v. 1932, 183 p. Also Bulletin Société Géologique de France, ser. 5, v. 2, p. 429–430. Rivière, A., 1933, Sur la tectonique de la region du col de Gudack et de la haute vallée du Talar (Elbourz, Perse): Compte Rendu Sommaire des Séances de la Société Géologique de France, v. 1933, no. 12, p. 169–171.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084715/spe525-03.pdf by University of Wisconsin Madison user

Rivière, A., 1934a, Note preliminaire sur les montagnes de la région de Sepaieh (Elbourz, Perse): Compte Rendu Sommaire des Séances de la Société Géologique de France, v. 1934, no. 16, p. 242–244. Rivière, A., 1934b, Contribution à l’étude géologique de l’Elbourz (Perse): Révue de Géographie Physique et de Géologie Dynamique, v. 7, no. 1–2, 190 p. Rivière, A., 1935a, Details tectoniques nouveaux dans la vallée du Djerdjerud (arc Iranien septentrional): Compte Rendu Sommaire des Séances de la Société Géologique de France, v. 1935, no. 10, 140 p. Rivière, A., 1935b, Observations nouvelles sur le secondaire de l’Anti-Elbourz (Perse): Comptes Rendus des Séances de l’Académie des Sciences de Paris, v. 200, no. 11, p. 951–953. Rivière, A., 1935c, Observations nouvelles sur la structure des montagnes de Schah Abdul Azim (Perse): Compte Rendu Sommaire des Séances de la Société Géologique de France, v. 1935, no. 5, p. 60–62. Rivière, A., 1936, Contribution à l’étude géologique de Anti-Elbourz: Bulletin Société Géologique de France, ser. 5, v. 6, no. 4–5, p. 277–295. Rivière, A., Sobhani, A., and Vernhet, S., 1973, Sédimentologie de la plaine de Gorgan: Comptes Rendus de l’Académie des Sciences, ser. D, v. 276, no. 19, p. 2613–2615. Roaf, M., 1983, Sculptures and Sculptors at Persepolis: London, British Institute of Persian Studies, 164 p. Rodler, A., 1885, Das knochenlager und die fauna von Maragha: Wien, Verhandlunger der Kaiserlich-Königlichen Geologischen Reichsanstalt, p. 333–337. Rodler, A., and Weithofer, K.A., 1890, Die Wiederkauer der Fauna von Maragha: Vienna, Denkschrifen K. Akademie der Wissenschaften: Mathematische-Naturwissenschaftlichen Klasse, v. 62, p. 753–772. Rolland, Y., Galoyan, G., Bosch, D., Sosson, M., Corsini, M., Fornari, M., and Verati, C., 2009, Jurassic back-arc and Cretaceous hot-spot series in the Armenian ophiolites—Implications for the obduction process: Lithos, v. 112, p. 163–187. Rosenthal, F., tr., 1958. See Ibn Khaldun, 1377. Rossetti, F., Nasrabady, M., Vignaroli, G., Theye, T., Gerdes, A., Razavi, M.H., and Moin Vaziri, H., 2010, Early Cretaceous migmatitic mafic granulites from the Sabzevar range (NE Iran): Implications for the closure of the Mesozoic peri-Tethyan ocean in Central Iran: Terra Nova, v. 22, no. 1, p. 26–34, doi:10.1111/j.1365-3121.2009.00912.x. Rumi (sic, Balkhi) Maulānā Jalāl U’ddin, 1263–1273, Mathnavi-ye Ma’navi [The Mathnawi of Jalālu’ddin Rumi]. See Nicholson, R.A. (trans. and ed.) 1926. Persian edition: Ketāb Mathnavi Ma’navi (6th ed.): Tehrān, Maulā Publishing, 1368/1989, 3 volumes, 1522 p. Ruttner, A., ed., 1991, The Triassic of Aghdarband (AqDarband), NE-Iran, and its pre-Triassic frame: Vienna, Geologische Bundesanstalt, v. 38, no. 1991, 252 p. Ruttner, A., Nabavi, M.H., and Hajian, J., 1968, Geology of the Shirgesht Area (Tabas Area, East Iran): Tehran, Geological Survey of Iran, Report no. 4, 133 p. Sabzehei, M., and Berberian, M., 1972, Preliminary note on the structural and metamorphic history of the area between Dowlatabad and Esfandagheh, southern central Iran: Tehran, Geological Survey of Iran, Internal Report, 19 p. Also in First Iranian Geological Symposium Proceedings: Tehran, Iranian Petroleum Institute, 1973. Sachau, E.C., ed., 1910, Albiruni’s India, Philosophy, Geography, Chronology, Astronomy, Customs, Laws, and Astrology of India: London, Keagan Paul, Trench Trübner & Co. Ltd., 2 volumes. See also Biruni, 1030. Sa’id, H.K., ed., 1989, Al-Beruni’s Book on Mineralogy, The Most Comprehensive in Knowledge on Precious Stones [Ketāb al-Jamāher fi Ma’refat al-Javāher], 916–1048: Islamabad, Pakistan Hijra Council, 1410/1989, 355 p. Salam, A., 1984, Islam and Science, in Lai, C.H., ed., 1987, Ideals and Realities, Selected Essays of Abdus Salam (2nd ed.): Singapore, World Scientific, p. 179–213. De Santillana, G., 1968, Preface to Science and Civilization, in Nasr, S.H., 1968, Science and Civilization of Islam: Cambridge, Massachusetts, Harvard University Press, p. vii–xiv. Sarton, G., 1927, Introduction to the History of Science: Volume 1. From Homer to Omar Khayyam: Washington, D.C., Carnegie Institution of Washington, p. 60–783. Sarton, G., 1952, A Guide to the History of Science: Waltham, Massachusetts, Chronica Botanica Company, 316 p. Scheppler, B., 2006, Al-biruni, Master Astronomer and Muslim Scholar of the Eleventh Century: New York, Rosen Publishers Group, Inc., 112 p. Schmadel, L.D., 2012, Dictionary of Minor Planet Names (6th ed.): New York, Springer, 1309 p., ISBN: 978-642-29718-2.

Development of geological perceptions and explorations on the Iranian Plateau Schmidt, E.F., 1937, Excavations at Tepe Hissar, Damaghan, 1931–1933: Philadelphia, Pennsylvania University, University Museum, Publication of Iranian Section, University of Philadelphia Press, 478 p. Schmidt, E.F., 1953, Persepolis I: Structure, Reliefs, Inscriptions: Chicago, Illinois, The University of Chicago Press, Oriental Institute v. LXVII, 160 p. Schmidt, E.F., 1970, Persepolis III: The Royal Tombs and Other Monuments: Chicago, University of Chicago Press, Oriental Institute v. LXX, 378 p. Schröder, J.W., 1944, Essai sur la structure de l’Iran [Basel]: Geneva, Eclogae Geologicae Helvetiae, v. 37, p. 37–81. Schröder, J.W., 1945a, Sur l’âge des couches vertes de l’Albourz (Iran): Geneva, Compte Rendu des Seances de la Société de Physique et d’Histoire Naturelle de Genève, v. 62, no. 1, p. 31–33. Schröder, J.W., 1945b, Quelques aspects de la géologie de l’Iran: Geneva, Bulletin Vereingung Schwizerissher Petroloumgeologen un Petroleumingenieure, band 12, heft 39, p. 11–20. Schröder, J.W., 1946, Géologie de l’île de Larak, Contribution à l’étude des dômes de sel du Golfe Persique (Comparaison avec Salt Range): Geneva, Archives des Sciences, Société de Physique de d’Histoire Naturelle de Genève, v. 28, p. 5–18. Schröder, J.W., 1965, Dictyoconus pachymarginalis n. sp. aus dem Apt des Elburzgebirges (Nordiran), Studien über primitive Orbitolinodae, III: Genève, Eclogae Geologicae Helvetiae, v. 58, no. 2, p. 975–980. Seely, P.H., 1991, The firmament and the water above: Westminster Theological Journal, v. 53, p. 227–240. Seely, P.H., 1997, The geographical meaning of ‘Earth’ and ‘Seas’ in Genesis 1:10: Westminster Theological Journal, v. 59, p. 231–255. Sedlacek, A.M., 1956, Petrographieche ergebnisse der forscgungsreise Dr. A. Gabriels in die südliche Lut und Persisch Baluchistan: Wien, SpringerVerlag, Tschermaks Mineralogische und Petrographische Mitteilungen, pt. 3, v. 5, no. 4, p. 412–443. Şengör, A.M.C., 1979, Mid-Mesozoic closure of Permo-Triassic Tethys and its implications: Nature, v. 279, p. 590–593. Şengör, A.M.C., 1984, The Cimmeride Orogenic System and the Tectonics of Eurasia: Geological Society of America Special Paper 195, 82 p., doi:10.1130/SPE195. Şengör, A.M.C., 1990, A new model for the late Palaeozoic–Mesozoic tectonic evolution of Iran and implications for Oman, in Robertson, A.H., Searle, M.P., and Ries, A.C., eds., The Geology and Tectonics of the Oman Region: Geological Society, London, Special Publication 49, p. 797–831. Şengör, A.M.C., and Hsu, K.J., 1984, The Cimmerides of Eastern Asia, History of the Eastern End of Paleo-Tethys: Memoires de la Société Géologique de France, Nouvelle Serie, v. 174, p. 139–167. Şengör, A.M.C., and Natal’in, B.A., 1996, Paleotectonics of Asia, fragments of a synthesis, in Yin, A., and Harrison, M., eds., The Tectonic Evolution of Asia: Cambridge, UK, Cambridge University Press, p. 486–640. Senowbari-Daryan, B., 2005, Hypercalcified sphinctozoan sponges from Upper Triassic (Norian–Rhaetian) reefs of the Nayband Formation (central and northeast Iran): Jahrbuch der geologischen Bundesanstalt, Vienna, v. 145, no. 2, p. 171–177. Setudehnia, A., 1972, Lexique Stratigraphique International, Volume III, Fascicule 1, Iran du Sud-Ouest: Paris, Centre Nationale de la Recherche Scientifique (CNRS), p. 289–376. Seyed Emami, K., Schairer, G., Aghanabati, S.A., Fuersich, F.T., Senowbari Daryan, B., and Majidifard, M.R., 1998, Cadomites from the Baghamshah Formation, southwestern Tabas: Mitteilungen der Bayerischen Staatssammlung für Palaeontologie und Historische Geologie, v. 38, p. 111–119. Shafaii Moghadam, H., Zaki Khedr, M., Arai, S., Stern, R.J., Ghorbani, G., Tamura, A., and Ottley, C.J., 2013, Arc-related harzburgite-dunitechromitite complexes in the mantle section of the Sabzevar ophiolite, Iran: A model for formation of podiform chromitites: Gondwana Research, v. 27, no. 2, p. 575–593, doi:10.1016/jgr.2013.09.007. Shaked, S., 2005, Zoroastrian origins, Indian and Iranian connections, in Arnason, J.P., Eisenstadt, S.N., and Wittrock, B., eds., Axial Civilizations and World History: Leiden, Netherlands, E.J. Brill, p. 184–200. Shojaat, B., Hassanipak, A.A., Mobasher, K., and Ghazi, A.M., 2003, Petrology, geochemistry and tectonics of the Sabzevar ophiolite, north Central Iran: Journal of Asian Earth Sciences, v. 21, no. 9, p. 1053–1067. Simanek, D.E., 2006, The Flat Earth: lhup.edu/~dsimanek/flat/flateart.thm (accessed 2014). Skjærvø, P.O., 1997, The state of Old Avestan scholarship: Journal of the American Oriental Society, v. 117, p. 103–107.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084715/spe525-03.pdf by University of Wisconsin Madison user

83

Sjögren, H., 1888, Beiträge zur Geologie des Berges Savelan in nordlichen Persien [Contributions to the Geology of the Mountains in Northern Persia]: Verhandlungen der Russisch Kaiserlichen Mineralogischen Gesellschaft zu St. Petersberg, Series 2, v. 24, p. 36–66. Skjærvø, P.O., 2003, Zarathustra; first poet-sacrificer, in Adhami, S., ed., Paitimāna. Essays in Iranian, Indian, and Indo-European Studies in Honor of Hanns-Peter Schmidt, 2 volumes in one: Costa Mesa, Mazda Publishers, p. 157–194. Soder, P.A., 1983, Karl Theodor Goldschmid (1896–1982): Zürich, Bulletin der Vereinigung Schweiz Petroleum-Geologen und Ingenieure, v. 49, no. 117, p. S54–S56, October 1983. Soder, P.A., 1992, Dr. Heinrich Martin Huber, 10.1.1917–27.6.1992: Zürich, Bulletin Vereinigung Schweiz Petroleum-Geologen und Ingenieure, v. 59, no. 135, p. S13–S14, October 1992. Soduqi-Sahā, 1983. See Biruni, 1030. Sorkhabi, R., 2005, Pre-modern history of bitumen, oil and natural gas in Iran (Persia): Oil-Industry History, v. 6, p. 153–177. Sorkhabi, R., 2012, Memorial to Augusto Gansser (1910–2012): Geological Society of America Memorials, v. 41, p. 15–21. Sosson, M., Roland, Y., Müller, C., Danelian, T., Melkonyan, R., Kekelia, S., Adamia, S., Babazdeh, V., Kangarli, T., Avagyan, A., Galoyan, G., and Mosar, J., 2010, Subductions, obduction and collision in the Lesser Caucasus (Armenia, Azerbaijan, Georgia), new insights, in Sosson, M., Kaymakci, N., Stephenson, R.A., Bergerat, F., and Starostenko, V., eds., Sedimentary Basin Tectonics from the Black Sea and Caucasus to the Arabian Platform: Geological Society, London, Special Publication 34, p. 329–352, doi:10.1144/SP340.14. Sparavigna, A.C., 2013, The science of al-Biruni: Torino, Italy, Department of Applied Science and Technology, Politecnico di Torino, Italy, 11 p., http:// arxiv.org/ftp/arxiv/papers/1312/1312.7288.pdf. Spellman, F.R., and Whiting, N.E., 2005, Environmental Engineers Mathematical Handbook: Boca Raton, Florida, CRC Press, 664 p. Stahl, A.F. von, 1893a, Das petroleum von Persien: Chemiker-Zeitung, v. 17, no. 77, p. 1409–1410. Stahl, A.F. von, 1893b, Die eisenerze Persiens: Chemiker-Zeitung, v. 17, no. 83, p. 1910–1911. Stahl, A.F. von, 1893c, Die steinkohlen Persiens: Chemiker-Zeitung, v. 17, no. 87, p. 1596. Stahl, A.F. von, 1894a, Die Kupfererze Persiens: Chemiker-Zeitung, v. 18, no. 1, p. 3–4. Stahl, A.F. von, 1894b, Die bleierze Persiens: Chemiker-Zeitung, v. 18, no. 21, 364 p. Stahl, A.F. von, 1894c, Verschiedene erze und mineralien Persiens: ChemikerZeitung, v. 18, no. 27, p. 487–488. Stahl, A.F. von, 1895, Reisen in nord- und zentral-Persiens: Pettermans Mitteilungen, v. 25, no. 118, p. 1–39. Stahl, A.F. von, 1897, Zur geologie von Persiens; Geognostische Beische Beischreibung des norlichen und zentral Persiens: Pettermans Mitteilungen, v. 37, no. 122, p. 1–72. Stahl, A.F. von, 1900, Karte der Umgegend von Teheran: Gotha, Germany, Justus Perthes, scale 1:210,000. Stahl, A.F. von, 1903a, Petroleum im Persia: Petroleum, v. 3, p. 623. Stahl, A.F. von, 1903b, Von der Kaukasischen grenze nach Tabriz und Kazwin: Pettermans Mitteilungen, v. 49, p. 60–64. Stahl, A.F. von, 1904a, Die erze des Karadag im Persien: Chemiker-Zeitung, v. 17, no. 6, p. 58; v. 17, no. 8, p. 85–86. Stahl, A.F. von, 1904b, Die orographischen und geologischen verhaeltnisse des Karadag in Persien: Pettermans Mitteilungen, v. 50, no. 10, p. 227–235. Stahl, A.F. von, 1905, Reisen in zentral- und west Persien: Pettermans Mitteilungen, v. 51, p. 4–12, p. 31–35. Stahl, A.F. von, 1907, Geologische beobachtungen in zentral- und nord-westPersien: Pettermans Mitteilungen, v. 53, no. 8, p. 169–177, p. 205–214. Stahl, A.F. von, 1909, Geologische beobachtungen im nordweslichen Persien: Pettermans Mitteilungen, v. 55, no. 1, p. 1–10. Stahl, A.F. von, 1911, Persien: Heidelberg, Hansboch der Regionalen Geologie, v. 5, pt. 6, no. 8, p. 1–46. Stahl, A.F. von, 1925, Die brennmaterialien Persiens, Braunkohle und brikettindustrie: Chemiker-Zeitung, v. 18, no. 46. Stahl, A.F. von, 1927a, Die kupfererze vorkommen Persiens: Berg-Technik, v. 20, no. 25, p. 329–330. Stahl, A.F. von, 1927b, Die orographischen und hydographischen Verhältnisse des Elburzgebirges in Persien: Pettermans Mitteilungen, v. 73, p. 211–215.

84

M. Berberian

Stahl, A.F. von, 1928, Persien und seine wirtschaftlichen hilfsquellen: Geographische Zeitschrift, v. 34, no. 4, p. 227–235. Stahl, A.F. von, 1933, Die geologischen verhaeltnisse des suedlichen ufergebietes des Kaspischen Meers in bezug auf erdoelvorkommen: Petroleum, v. 29, no. 5, p. 6–7. Stalder, P., 1971, Magmatisme Tertiare et subrécant entre Taleqan et Alamut, Elbourz Central (Iran): Zürich, Schweizerische mineralogische und petrographische Mitteilungen (Bulletin Suisse de Minéralogie et Pétrologie), v. 51, no. 1, p. 1–138, e-periodica.ch/digbib/view?pid=smp-001:1971:51#9. Stampfli, G., Marcoux, J., and Baud, A., 1991, Tethyan margins in space and time: Palaeogeography, Palaeoclimatology, Palaeoecology, v. 87, p. 373–409. Steiger, R., 1966, Die geologie der west-Firuzkuh area (Zentralelburz/Iran) [Ph.D. thesis]: Zürich, Mitteilungen Geological Institute (ETH), University of Zürich, n.s. 145, 145 p. Steingruber Stöcklin, E., 2011, Biographical Sketch of Jovan Stocklin (1921– 2008), A Self Portrait Conserved by His Widow Elisabeth: Erice, Italy, Ettore Majorana Foundation and Centre for Scientific Culture (EMFCSC), Extended Abstract Book (37th Interdisciplinary Workshop of the International School of Geophysics, 4–9 October 2011, Erice, Italy), p. 213–218, ftp://ftp.ingv.it/pro/web_ingv/37th%20Course/Extended%20 Abstracts%20Book.pdf. Stöcklin, J., 1952, Geology of the Central Coastal Makran Area: Tehran, filed at the Iranian Oil Company/National Iranian Oil Company, unpublished Internal Report GR 91C. Stöcklin, J., 1968a, Structural history and tectonics of Iran, a review: American Association of Petroleum Geologists Bulletin, v. 52, p. 1229–1258. Stöcklin, J., 1968b, Salt deposits of the Middle East, in Mattox, R.B., ed., Saline Deposits: A Symposium Based on Papers from the International Conference on Saline Deposits, Houston, Texas, 1962: Geological Society of America Special Paper 88, p. 157–181. Stöcklin, J., 1972, Lexique Stratigraphique International, Volume III, Fascicule 1, Iran Central, Septentrional et Oriental: Paris, Centre Nationale de la Recherche Scientifique (CNRS), p. 1–283. Stöcklin, J., 1974a, Possible ancient continental margins in Iran, in Burk, C.A., and Drake, C.L., eds., The Geology of Continental Margins: New York, Springer Verlag, p. 873–887. Stöcklin, J., 1974b, Northern Iran; Alborz Mountains, in Spencer, A., ed., Mesozoic–Cenozoic Orogenic Belts, Data for Orogenic Studies: Geological Society, London, Special Publication 4, p. 213–234. Stöcklin, J., 1977, Structural correlation of the Alpine ranges between Iran and central Asia: Paris, Mémoire Hors série Société Géologique de France, v. 8, p. 333–353. Stöcklin, J., 1981a, A brief report on geodynamics of Iran, in Gupta, H.K., and Delany, F.M., eds., Zagros–Hindu Kush–Himalaya Geodynamic Evolution: American Geophysical Union Geodynamics Monograph 3, p. 70–73. Stöcklin, J., 1981b, Orogeny and Tethys evolution in the Middle East; An appraisal of current concepts, in 27th International Geological Congress, Colloquium 05, Reports Volume 5, Tectonics of Asia, p. 65–84. Stöcklin, J., 2006, Persien, Einnerungen Eines Geologen [Persia, Memoirs of a Geologist]: Zurich, Switzerland, Eigenverlag, self-published, 165 p. Manuscript deposited in the ETH Archive, Zurich, Switzerland, and in the Geologists’ Archive, the Albert Ludwig University of Freiburg, Freiburg im Breist, Germany. Stöcklin, J., Ruttner, A., and Nabavi, M., 1964, New Data on the Lower Paleozoic and Pre-Cambrian of North Iran: Tehran, Geological Survey of Iran, Report no. 1, 29 p. Stöklin, J., Eftekhar-nezhad, J., Hushmand-zadeh, A., Nabavi, M., Zahedi, M., Valeh, N., and Tatavussian, S., 1965, Geology of the Shotori Range (Tabas area, East Iran): Tehran, Geological Survey of Iran, Report no. 3, 69 p. Stöcklin, J., Eftekharnezhad, J., and Hushmanzadeh, A., 1972, Geological Reconnaissance Map of Central Lut: Tehran, Geological Survey of Iran, scale 1:500,000. Stöllner, T., 2005, Early mining and metallurgy on the Iranian Plateau, in Yalcin, U., ed., Anatolian Metal III, Der Anschnitt Zeitschrift FürKunst und Kultur IM Bergbau: Deutsches Bergbau-Museum (DBM), Bochum, v. 18, p. 191–207, ISBN: 978-3937203072. Strabo, 63 BCE–24 CE, The Geography (translated by H.L. Jones): Cambridge, Massachusetts, Harvard University Press, 1917–1932, http://penelope .uchicago.edu/Thayer/E/Roman/Texts/Strabo/15B*.html.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084715/spe525-03.pdf by University of Wisconsin Madison user

Stratil-Sauer, G.,1934, Kampf um die Wüste, Ein Bericht ueber unsere Fahrten in die ostpersische Lut: Berlin, R. Hobbing, 174 p. Stratil-Sauer, G., 1937a, Meschhed—Eine Stadt baut am Vaterland Iran: Leipzig, Verlag Ernst Staneck, 167 p. Stratil-Sauer, G., 1937b, Eine Route im Gebiet des Kuh-e-Hezar (Südiran): Berlin, Petermanns Geographische Mitteilungen, v. 11, p. 309–313, and v. 12, p. 353–356. Stratil-Sauer, G., 1941, Beobachtungen zur Sommerwitterung einer südpersischen Höhenstation (Ra’in): Gerlands Beiträge Geophysische, v. 57, p.193–225. Stratil-Sauer, G., 1953, Geographische Forschungen in Ostpersien: Wien, Abhandlungen der Geographischen Gesellschaft, v. 17, no. 2–3, p. 1953–1956. Stratil-Sauer, G., 1957, Die pleistozänen Ablagerungen im Innern der Wüste Lut: Wien, Festschr. 100, Jahrfeier Geographische Gesellschaft, p. 1856– 1956, p. 460–484. Stratil-Sauer, G., and Weise, O.R., 1974, Zur Geomorphologie der südlichen Lut und zur Klimageschichte Irans: Würzburg, Würzburger Geographische Arbeiten, Mitteilungen der Geographischen Gesellschaft Würzburg, no. 41, 29 p. Stronach, D., 1978, Pasargadae; A Report on the Excavations Conducted by the British Institute of Persian Studies from 1961 to 1963: Oxford, UK, Oxford University Press, 326 p. Surhone, L.M., Tennoe, M.T., and Henssonow, S.F., eds., 2011, Pax Mongolica, Gengis Khan, Black Death, Ögedei Khan: Saarbrücken, Germany, Betascript (OmniCriptum) Publishing 108 p., ISBN: 978-613-5-03492-9. Süssli, P.E., 1974, The Geology of the Lower Haraz Valley Area (Central Alborz/Iran) [Ph.D. diss.]: Zürich, Naturwissenschaften, ETH Zürich, no. 5407. Also Tehran, Geological Survey of Iran, no. 38, 116 p., 1976. Tabari, 850, The History of al-Tabari (Ta’rikh al-rusul wa’l-Muluk): Volume IV. The Ancient Kingdoms (translated and annotated by Moshe Perlmann): New York, State University of New York (SUNY) Press, SUNY Series in Near East Studies, 1987, 224 p., ISBN: 978-0-88706-181-3; Volume XIII. The Conquest of Iraq, Southwestern Persia, and Egypt, The Middle Years of ‘Umar’s Caliphate A.D. 636–642/15–21 A.H. (translated by H.A. Gautier Juynboll): New York, State University of New York (SUNY) Press, 1989, 251 p.; Volume XIV. The Conquest of Iran, A.D. 641–643/ A.H. 21–23 (translated by G. Rex Smith): New York, State University of New York (SUNY) Press, 1994, 216 p., ISBN: 978-09791412974. Also Bibliotheca Persica, edited by Ehsan Yar-Shater, State University of New York Press. Tajaddod, M.R., 1987, Al-Fehrest Mohammad b. Eshāq ibn-Nadim [Al-Fehrest of ibn Nadim]: Tehran, Asātir Publishing, 825 p. Takin, M., 1972, Iranian geology and continental drift in the Middle East: Nature, v. 235, no. 5334, p. 147–150. Tavernier, J.B., 1681, Les Six Voyages de Jean-Baptiste Tavernier, qu’il a Fait en Turquie, en Perse, et aux Indes: Paris, G. Clouzier, Volume 3, 698 p., online at gallica.bnf.fr/ark:/12148/bpt6k853250/f744.image. Ter-Davtyan, K.S., and Arevshatyan, S.S., 1962. See Ānāniā Shirākātsi, ca. 660–681 CE. Tietze, E., 1875a, Ueber quellen und quellenbildungen am Damavend und dessen umgebung: Jahresbericht Geologische Bundesanstalt, v. 25, p. 129–140. Tietze, E., 1875b, Ueber die brankohlen von Tabriz: Verlag der Kais Kön Geologischen Reichsanstat, Wien, v. 1875, p. 23–46. Tietze, E., 1875c, Geologische mitteilungen aus Persien, Schreiben an Herrn rergrat Dr. Stur, de dato Tehran 28 February 1875: Wien, Verlag der Kais Kön Geologischen Reichsanstat, 1875, p. 129–133. Tietze, E., 1875d, Ein ausflug nach dem Siah-Kuh (Scgwarzer Berg) in Persien: Mitteilungen der Geographischen Gesellschaft Wien, ser. 8, v. 18, p. 257–267. Tietze, E., 1877a, Zur theorie der entstehung der salzsteppen und der angeblichen enststehung der salzlager aus salzsteppen: Jahresbericht Geologische Bundesanstalt, v. 27, p. 341–374. Tietze, E., 1877b, Bemerkugen under die tektonik des Alburs gebirges: Jahresbericht Geologische Bundesanstalt, v. 27, p. 375–430. Tietze, E., 1877c, Beiträge zur geologie Persiens: Wien, Jahrbuch der Kais Kön Geologischen Reichsanstalt, band 27, heft IV, p. 1877. Tietze, E., 1878, Der Vulcan Demavend in Persien: Jahresbericht Geologische Bundesanstalt, v. 28, p. 169–206. Tietze, E., 1879, Die Mineralreichtümer Persiens: Jahresbericht Geologische Bundesanstalt, v. 29, p. 565–568. Tietze, E., 1881, Ueber bildungen der juengeren epochen in nord-Persien: Jahresbericht Geologische Bundesanstalt, v. 31, p. 67–130.

Development of geological perceptions and explorations on the Iranian Plateau Tietze, E., 1886, Ueber die bodenplastik in der geologischen beschaffenheit Persiens: Wien, Mitteilungen aus dem Geologischen Gesellschaft, v. 29, p. 513–523, p. 561–567. Tirrul, R., Bell, I.R., Griffis, R.J., and Camp, V.E., 1983, The Sistan suture zone of eastern Iran: Geological Society of America Bulletin, v. 94, p. 134– 150, doi:10.1130/0016-7606(1983)942.0.CO;2. Toulmin, E., and Goodfield, J., 1965, The Ancestry of Science, The Discovery of Time: London, Hutchinson and Co., Ltd., and Chicago, Illinois, University of Chicago, ISBN: 0-226-80842-4, 280 p. Tozer, E.T., 1972, Triassic Ammonoids and Daonella of the Nakhlak Group, Anarak Region, Central Iran: Geological Survey of Iran, Report no. 28, p. 29–69. Turner, A.K., 2008, The historical record as a basis for assessing interactions between geology and civil engineering: Quarterly Journal of Engineering and Hydrogeology, v. 41, p. 143–164. University of Tehran Library, n.d., Old Manuscripts: University of Tehran Central Library, Noskheh-ye Khatti. Vallat, F., 2013, The Main Achaemenid inscriptions at Susa, in Perrot, ed., The Palace of Darius at Suza; The Great Royal Residence of Achaemenid Persia: London, I.B. Tauris & Co Ltd., p. 281–294. Vaughan, H.B., 1893, A journey through Persia: Royal Geographical Society Journal, Supplementary Papers 3, p. 89–115. Verdel, C., Wernicke, B.P., Hassanzadeh, J., and Guest, B., 2011, A Paleogene extensional arc flare-up in Iran: Tectonics, v. 30, no. 3, TC3008, doi:10.1029/2010TC002809. Vincent, E.A., 1972, Address read at Memorial Services for Dr. J.V. Harrison in Christ Church Cathedral, Oxford, 18 November 1972, 3 p.: http://www .robertnrussell.com/harrisonJVernon.pdf. Vredenburg, E., 1901, A geological sketch of the Baluchistan desert and part of eastern Persia: Memoirs of the Geological Survey of India, v. 31, no. 2, p. 179–203. Vredenburg, E., 1908, Occurrence of the genus Orbitolina in India and Persia: Records of the Geological Survey of India, v. 26, no. 4, p. 314. Vredenburg, E., 1909, Note on a hippuritic-bearing limestone in Seistan and on the geology of the adjoining region: Records of the Geological Survey of India, v. 38, no. 3, p. 216–229. Watson, R.G., 1862, Notes of a visit to the Elburz Mountains and ascent of Demavand: Royal Geographical Society of London Proceedings, v. 6, p. 103–105. Wertime, Th.A., 1964, Man’s first encounters with metallurgy: Science, v. 146, no. 3649, p. 1257–1267. Wertime, Th.A., 1968, A metallurgical expedition through the Persian desert: Science, v. 159, no. 3818, p. 927–935. Wertime, Th.A., 1973, The beginning of metallurgy; a new look: Science, v. 182, p. 875–886. West, E.W., ed., 1880a, The Bundahis, Pahlavi text, Sacred Books of the East (Max Muller, F., ed.): Delhi, India, Motilal Banarsidass, 1987, v. 5, p. 1–187. West, E.W., ed., 1880b, The Bahman/Zand-e Vahouman Yast, Pahlavi Texts, Sacred Books of the East (Max Muller, F., ed.): Delhi, India, Motilal Banarsidass, 1987, v. 5, p. 189–235. West, E.W., ed., 1882, Dadistan-I Dinik, Pahlavi Texts, Sacred Books of the East (Max Muller, F., ed.): Delhi, India, Motilal Banarsidass, 1987, v. 18, p. 1–276. West, E.W., ed., 1885, The Dinā-I Maing-I Khirad, Pahlavi Texts, Sacred Books of the East (Max Muller, F., ed.): Delhi, India, Motilal Banarsidass, Delhi, 1987, v. 24, part III, p. 1–119. West, E.W., ed., 1897, Marvels of Zoroastrianism, Dinkard and Selections of Zād-Sparam, Pahlavi Texts, Sacred Books of the East (Max Muller, F., ed.): Delhi, India, Motilat Banarsidass, v. 47, p. 3–130, 133–170. Whitaker, E.A., 2003, Mapping and Naming the Moon; A History of Lunar Cartography and Nomenclature: New York, Cambridge University Press, 262 p. Wilber, D.N., 1955, The Architecture of Islamic Iran; the Il-Khanid Period: Princeton Monographs in Art and Archaeology XXIX, Oriental Studies XVII, Department of Art and Archaeology and Oriental Languages and Literatures: Princeton, New Jersey, Princeton University Press, 208 p. + 60 figs., 216 plates.

85

Wilcke, C., 1969, Das Lugalbandaepos: Wiesbaden, Harrassowitz, 236 p. + XII plates. Wilmsen, M., Fürsich, F.T., Seyer-Emami, K., Majidifard, M.R., and Taheri, J., 2009, The Cimmerian orogeny in northern Iran; tectono-stratigraphic evidence from the foreland: Terra Nova, v. 21, no. 3, p. 211–218, doi:10.1111/j.1365-3121.2009.00876.x. Wilson, A.T., 1930, Earthquakes in Persia: Bulletin of the Scholl of Oriental Studies, University of London, v. 6, no. 1, p. 103–131. Witter, W., 1938, Die älteste Erzgewinnung im bordisch-germanischen Lebenskries,Volume 2 Die Kenntnis von Kupfer und Bronze in der Alten Welt: Leipzig, Kabotzsch, 118 p. Woskoboinikow/Voskoboinikov, Lt.-Col., 1846, Voyages en Perse: St. Petersburg, Gorn. Zhurn., 171 p. Wulff, H.E., 1966, The Traditional Crafts of Persia: Cambridge, Massachusetts, Massachusetts Institute of Technology Press, p. 16. Wulff, H.E., 1968, The qanats of Iran: Scientific American, v. 218, no. 4, p. 94–105. Yarshater, E., 1983a, The Cambridge History of Iran: Volume 3. The Seleucid, Parthian and Sasanid Periods, Part 1: Cambridge, UK, Cambridge University Press, 1396 p. Yarshater, E., 1983b, The Cambridge History of Iran, Volume 2. The Seleucid, Parthian and Sasanid Periods, Part 2: Cambridge, UK, Cambridge University Press, 1488 p. Yarshater, E., 2004, Iranian history: An overview: Encyclopaedia Iranica, v. XIII, no. 2, p. 212–224, and v. XIII, no. 3, p. 225; also http://www .iranicaonline.org/articles/iran-ii1-pre-islamic-times#1. Yarshater, E., 2012, Iran ii, Iranian history (3) Chronological table: Encyclopaedia Iranica, v. XXX(3), p. 247–299; also iranicaonline, http://www .iranicaonline.org/articles/iran-ii3-chronological-table. Zaghi, N., 2007, Groundwater Exploration; the Oldest Hydrology Text More than 1000 Years Old Water Supply Technology, by Aboubakr Karaji, a Scientist Who Was Avicenna Contemporary: Tehran, The Iran Cultural Foundation, 122 p., ijee.ias.ac.ir/pdf_2619_174c9cd05 [Persian into English translation]. Also published in 2013, Groundwater Exploration: The World’s Oldest Water Resources Engineering Textbook: Amazon Digital Services, LLC, ASIN:B00HJDS5292. Zaghi, N., and Finnemore, E.J., 1973, Discussion of exploration for hidden water by M. Karaji—The oldest textbook on hydrology?: Ground Water, v. 11, no. 1, p. 44. Zanchetta, S., Zanchi, A., Villa, I., Poli, S., and Muttoni, G., 2009, The Shanderman eclogite, a Late Carboniferous high-pressure event in the NW Talesh Mountains (NW Iran), in Brunet, M.-F., Wilmsen, M., and Granath, J.W., eds., South Caspian to Central Iran Basins: Geological Society, London, Special Publication 312, p. 57–78. Zanchetta, S., Berra, F., Zanchi, A., Bergomi, M., Caridroit, M., Nicora, A., and Heidarzadeh, G., 2013, The record of the late Palaeozoic active margin of the Palaeotethys in NE Iran: Constraints on the Cimmerian orogeny: Gondwana Research, v. 24, p. 1237–1266. Zanchi, A., Zanchetta, S., Berra, F., Mattei, M., Garzanti, E., Molyneaux, S., Nawab, A., and Aabouri, J., 2009, The Eo-Cimmerian (Late? Triassic) orogeny in North Iran, in Brunet, M.-F., Wilmsen, M., and Granath, J.W., eds., South Caspian to Central Iran Basins: Geological Society, London, Special Publication 312, p. 31–55. Zarrinkub, A.H., 1975, The Arab conquest of Iran and its aftermath, in Frye, R.N., ed., The Cambridge History of Iran: Volume 4. The Period from the Arab Invasion to the Saljuqs: Cambridge, UK, Cambridge University Press, p. 1–56. Zāvosh, M., 1969, Mineralogy in Ancient Iran [Kānishenāsi dar Iran-e Qadim], Volume 1: Tehran, Bonyād Farhang Irān, 339 p. [in Persian]. Zāvosh, M., 1976, Mineralogy in Ancient Iran [Kāânishenāsi dar Iran-e Qadim], Volume 2: Tehran, Bonyād Farhang Irān, 402 p. [in Persian].

MANUSCRIPT ACCEPTED BY THE SOCIETY 15 JUNE 2016 MANUSCRIPT PUBLISHED ONLINE 23 NOVEMBER 2016 REVISED VERSION PUBLISHED ONLINE 23 MARCH 2017

Printed in the USA

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084715/spe525-03.pdf by University of Wisconsin Madison user

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084715/spe525-03.pdf by University of Wisconsin Madison user

The Geological Society of America Special Paper 525

Tehran: An earthquake time bomb Manuel Berberian* Fellow, School of Mathematics, Science, and Technology, Department of Science, Ocean County College, Toms River, New Jersey 08754-2001, USA, and Onduni Grung Scientific Enterprise, 1224 Fox Hollow Drive, Toms River, New Jersey 08755-2179, USA Robert S. Yeats College of Earth, Ocean, and Atmospheric Sciences, Oregon State University, 104 Wilkinson Hall, Corvallis, Oregon 97331, USA

ABSTRACT The megacity of Tehran, the political, economic, and military center of Iran, is exposed to a risk of large-magnitude earthquakes originating on several adjacent and inner-city active faults. The city lies at the southern foot of the central Alborz Mountains, which frame the South Caspian Basin and have been the source of damaging historical left-lateral strike-slip and reverse-fault earthquakes. The most recent destructive earthquake in Alborz was the Rudbār left-slip earthquake of Mw 7.3 on 20 June 1990 northwest of Tehran, taking more than 40,000 lives and destroying three cities. This earthquake was in a seismic gap, and its source fault did not show clear geomorphic signs of being active prior to the earthquake. East of Tehran, the 22 December 856 Komesh (Dāmghān) earthquake had a magnitude previously estimated at Ms 7.9, with estimated losses of 40,000–200,000 lives. Our reevaluation of historical, archaeological, and structural evidence reduces estimates of both magnitude and losses, similar to the 1990 Rudbār earthquake. The latest earthquake to affect the present Tehran metropolitan area was the Lavāsānāt earthquake on the central section of the Moshā fault, on 27 March 1830, with its epicenter located ~30 km northeast of the city, which had a magnitude of Mw ~7.0–7.4. Prior to this event, the Ruyān earthquake north of Tehran struck the same section of the fault on 23 February 958, with a magnitude previously estimated as Ms 7.7, although our reevaluation reduces the magnitude to around Mw ≥7.0 (7.0–7.4). Both the 958 and 1830 earthquakes along the central segment of the Moshā fault, with an interval of 872 yr, might have loaded the North Tehran fault system near the cities of Tehran and Karaj, as well as the faults underneath the metropolitan area. The North Tehran fault system west of Tehran might have sustained an earthquake of Mw ~7.0 in May 1177. The earthquake histories of the Niāvarān, Darakeh, Farahzād, and Kan leftlateral strike-slip faults (part of the North Tehran fault system at the mountain front

*E-mails: Berberian—[email protected], [email protected] (corresponding e-mail); Yeats—[email protected]. Berberian, M., and Yeats, R.S., 2016, Tehran: An earthquake time bomb, in Sorkhabi, R., ed., Tectonic Evolution, Collision, and Seismicity of Southwest Asia: In Honor of Manuel Berberian’s Forty-Five Years of Research Contributions: Geological Society of America Special Paper 525, p. 87–170, doi:10.1130/2016.2525(04). © 2016 The Geological Society of America. All rights reserved. For permission to copy, contact [email protected].

87

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

88

Berberian and Yeats north of the city), the inner-city Mahmudieh and Dāvudieh south-dipping reverse faults, the central-eastern section of the North Tehran fault system (now within the metropolitan area) as well as blind thrusts under the city are unknown. Except for the 1830 distant earthquake, no medium- to large-magnitude earthquake (Mw 6.5–7.5) has occurred within the Tehran metropolitan area during the past 839 yr along the faults beneath the metropolitan area or in the immediate vicinity. This may indicate a >839-yr-long period of strain accumulation within a long interseismic period between large-magnitude earthquakes in Tehran. With the active-fault hazard to the rapidly growing population along several faults, it is necessary for the government to: (1) conduct extensive paleoseismic trenching to identify the most hazardous of Tehran’s faults, previous rupture areas, average coseismic slip rates, earthquake magnitudes, and average recurrence periods of earthquakes from at least eight fault systems within the metropolitan area; (2) deal with extensive corruption of the construction and building-inspection industries; and (3) enforce the 1969, 1988, and 1999 Iranian Code for Seismic Resistant Design of Buildings. As with the 12 January 2010, Mw 7.0 Haiti earthquake, losses from the next Tehran earthquake of Mw ≥7.0 could exceed 100,000 people. It is necessary to prepare and implement an earthquake management master plan as a disaster prevention tool, enforce the building code with transparency, and retrofit public structures and infrastructure in order to mitigate earthquake risk in Tehran and protect the lives of ~15 million people (roughly 20% of the country’s population) living in the Tehran and Alborz (Karaj city) Provinces.

Considering the continuous threat of earthquake hazard in Tehran and the suburbs, we should immediately prepare the necessary plans to mitigate the earthquake risk by utilizing the scientific and engineering techniques derived from earth sciences; we should do this before an earthquake strikes. —Professor Setrāk Ābdāliān, 1951, Tehran (translated from Persian)

1. INTRODUCTION AND STATEMENT OF THE PROBLEM1 Tehran, the capital of Iran (35.70°N, 51.40°E), covering an area of 1274 km2 with a population density of 10,327 persons/km2 at the southern foot of the central Alborz Mountains (Fig. 1), had a population of 8,656,506 in the 24 October 2011 census (Statistical Center of Iran [SCI], 2012). The Tehran and

1

All dates have been converted to the Georgian calendar system (new style; BCE and CE) used in the West. To assist with accounts in the post–seventhcentury CE Persian and Arabic sources, Arabic Hijra lunar (pre–1900 CE) calendar years are added throughout the text in a few cases and the tables. Note that the Persian and Arabic names and words in this report are written as correctly pronounced and written originally, with direct and simplified transliteration into English. Diacritical marks and special characters are used to differentiate vowel “a” (short; e.g., ant) from “ā” (long, e.g., Ārmenian), and Arabic “ain” (used also in Persian) as “‘a” (e.g., ‘Ābbās). Iran (pronounced “Irān”) and Tehran (“Tehrān”) are exempted from this rule due to familiarity with their correct pronunciation. The recognition of the Persian possessive (afzudeh; or ezāfé in Arabic), which inaccurately appears variously in English, especially as “-i” (thus: Shahr-i Kurd) is correctly shown as “-e” (cf. French “é”; thus: Shahr-e Kord), as conforms to the correct current usage in the Persian language (Pārsi/Fārsi). Elevations are given in meters above mean sea level. Coordinates of the sites discussed are given in the tables or the text.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

Alborz Provinces, with the two large cities of Tehran and Karaj, respectively, stretching along the southern central Alborz Mountain front, were the home of 14.6 million people, i.e., one-fifth of the 75 million population2 of Iran, in October 2011 (SCI, 2012). The economic, administrative, political, and military center of the central government is located in a high-seismicity zone at the southern foot of the faulted central Alborz Mountains and on the northern faulted edge of the Central Iran seismotectonic province. The megacity is surrounded by numerous active faults in the north, south, east, southeast, and northwest, including inner-city faults, and it has recorded large-magnitude historical (pre-1900) earthquakes. The expansion of Tehran, which has accelerated since the 1980s, has significantly increased the seismic risk. The probability of large-magnitude earthquakes resulting in enormous loss of life, property, and economic damage is very high in the city (Berberian et al., 1985; Berberian, 1994, 2005, 2014; Berberian and Yeats, 1999, 2001; JICA, 2000).

2

Population of the country in October 2011 was 75,149,669 (urban: 53,446,661; rural: 21,436,783). The population grew to 78,609,449 by August 2015. The population of the provinces of Tehran (12,183,391, with the capital city of Tehran) and Alborz (2,412,513, with its provincial capital city of Karaj, ~40 km WNW of Tehran) was 14,595,904 in October 2011 (SCI, 29 August 2015).

Tehran: An earthquake time bomb

Figure 1. Topographic map of the central Alborz Mountains (north of Tehran) and northern Central Iran (south of Tehran) with location of the figures, labeled by numbers, used in this study. The Figure 3 location is near no. 10. The Figure 23 location is near no. 22. Contour lines are in meters above the mean sea level and are shown in half tone. D with a solid contour line—Damāvand Quaternary volcano in central Alborz; F—Firuzkuh; M—Moshā; P—Pishvā; R—Rudbār; T—Tāléqān; V—Varāmin. Inset bottom left: AZ— Āzarbāijān; KP—Kopeh Dāgh; M—Makrān; S—Sistān.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

89

Tehran originated as a village in a semiarid region, established ca. 1500 CE (Fig. 2) on the “High Road” of ancient Iran, later called the “Silk Road,” between Mesopotamia and China, directly north of the ancient Zoroastrian city of Rhagae (modern Ray, now a southern suburb of Tehran), along the Alborz range front to the north. The city owes its location to the availability of water from the southern slopes of the Alborz Mountains and to nearly flat arable land with numerous qanāts (underground water aqueduct) for irrigation. The city has been the capital of Iran since 1795, when its population was estimated ~15,000 (Curzon, 1892; Adle and Hourcade, 1992; Planhol, 2004). Tehran’s selection as Iran’s capital in 1795 by the Qājār dynasty (1779–1925 CE), based on its agreeable climate and its proximity to the Qājār tribal lands southeast of the Caspian Sea, led to an increase of population (Fig. 2): 60,000 by 1811 (Ouseley, 1819) and 90,000 by 1847 (Issawi, 1971). The population of the city was 540,000 in 1939, 1.79 million in 1956, 4.67 million in 1976, 6.7 million in 1996, 7.79 million in 2007, and 8.6 million in 2011, surpassing 14.6 million in the wider metropolitan area (SCI, 1956–2015). The 2011 census shows that the total population of the two provinces of Tehran (with population density of 10,327 persons/km2) and Alborz in the west (Karaj, the third largest city in Iran after Tehran and Mashhad, with a population density of 696 persons/km2) along the southern margin of the Alborz Mountain was 14.6 million (SCI, 2012). This indicates that ~20% of the 75 million population of Iran in October 2011 lived along the active North Tehran fault system, inner-city faults, and in the vicinity of the seismically active Moshā, Pārchin, and Pishvā faults (Fig. 2). The reasons for Tehran’s population increase in recent decades include: (1) a strong central government; (2) an influx of refugees from the devastating Iran-Iraq war of 1980–1988 from both countries; (3) Afghan refugees from wars in Afghanistan, including the 1979–1989 Soviet occupation, the civil war with the Taliban that followed, and 1994–2001 Taliban rule (~1.4 million Afghanis in Iran, mostly in Tehran in 2011; SCI, 2012); and (4) migration from the countryside, where few jobs exist, to Iran’s largest city,3 a rapid growth common to other Asian capital cities. The rural population of Iran in October 2011 constituted 28.5% of the total population, and 71.4% of the population were living in urban areas (SCI, 2012), mostly close to active faults (Berberian, 2014). The increases in population (14.6 million in 2011), industries (>50% of Iran’s industrial centers), and all the government offices and military headquarters in Tehran suggest that a large part of the national economy of Iran, as well as its political and military centers, is vulnerable to a natural disaster. The Alborz

3

During the severe drought in recent years, mismanagement of water resources, excessive groundwater withdrawal, and declining subsurface water levels in the country (as well as in the Varāmin plain south of Tehran) have resulted in aquifer-system compaction, land subsidence, fissuring of agricultural lands, and drying of lakes and wetlands. This has caused abandonment of farming and villages and mass migrations to towns and cities, especially to Tehran.

90

Berberian and Yeats

Figure 2. Active fault map of Tehran at the southern Alborz Mountains piedmont in northern Central Iran (see Fig. 1 for location), with rapid and disproportionate area/population growth since it became the capital city of Iran on 20 March 1785. The old Tehran village is shown as a solid ellipse surrounded by later city limits (1554, 1867, 1957, 1970, and 1996). The ancient city of Rhagae (modern Ray) is southeast of Tehran. The inner and outer Ray city limits are those of the eighth century CE. Reverse faults with teeth on hanging-wall side; strike-slip faults shown with arrows. Faults without teeth or arrows: sense of latest slip uncertain (BF—Bāgh-e Fayz fault; T.T.—Takht-e Tāvus fault). Solid triangles—archaeological sites (see Appendix B). H—Hārun Prison in the southeast; N—Nezāmābād; Q—Qaytarieh in the north with the dates of ancient settlements. The E-W undulating line with + and − and open squares in central Tehran indicates anomalies in groundwater elevation dropping to the aquifer to the south, which could represent a hidden fault (Ministry of Water and Power [MWP], 1970; Knill and Jones, 1968; Tchalenko et al., 1974a). The escarpments south of Tehran (North Ray, South Ray, and Kahrizak) are shown by dashed lines with question marks; they may overlie faults. Radiation signs: TNRC—Tehran Nuclear Research Center (built in 1968); IAP/PHRC—Lavizān-Shiyān (Institute of Applied Physics [IAP]; later, Physics Research Center [PHRC]). Inset top left: AZ—Āzarbāijān; KP—Kopeh Dāgh; M—Makrān; S—Sistān; TP—Turān plate (Eurasia). Figure is modified from Berberian et al. (1985); Berberian and Yeats (1999, 2001); Berberian (2005, 2014).

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

Tehran: An earthquake time bomb Mountains are a source of water for Tehran, but they are also the location of active faults that have the potential for generating large-magnitude earthquakes. Minimizing human casualties in areas built upon and surrounded by numerous active faults is the main goal of seismicrisk mitigation plans, especially in urban areas. In a comprehensive pioneering study of the earthquake-fault hazard and risk assessment for the city of Tehran, detailed maps of the geology, land uses, seismicity, and ground acceleration were prepared, and the vulnerability of old buildings and human life was stressed to the authorities in 1985. Accordingly, setback zones along mapped active faults in the urban area were proposed for the first time. The preliminary ground acceleration for different levels of design was estimated at 0.27g for a 100 yr return period (Berberian et al., 1985). Based on this investigation, a seismic microzonation study for the city was conducted by the Japanese firm JICA (2000). However, neither the recommendations presented in Berberian et al. (1985) and JICA (2000), nor the implementation of the Ira-

91

nian Code for Seismic Resistant Design of Buildings since 1969 (ISIRI, Code No. 519, dated 1969, and 2800 dated 1988; BHRC, 1988, 1999) have been taken seriously by the authorities and the inhabitants at large. Unfortunately, since then, construction of buildings on active faults has covered almost all the inner-city faults, and most parts of the North Tehran fault system in the city are no longer accessible to paleoseismic investigation (Fig. 3). Although an emergency response plan for the first 72 h after an earthquake was prepared by the Japanese firm for Tehran (JICA, 2009), vulnerable structures of the old city have not yet been retrofitted, evacuated, or rebuilt. Disaster management is poor, emergency-response preparation and facilities are insufficient, and the megacity is already paralyzed by its poor traffic and transportation system. Five years after submittal of the Japanese emergency response plan, on 1 November 2014, Mehdi Hasani, chairman of the Development Commission of the Islamic Parliament, stated that the Tehran Disaster Mitigation and Management Center has not been able to address earthquake

Figure 3. Construction of the buildings over the North Tehran fault system at the southern foot of the Alborz Mountains (see Fig. 1 for the location). View to the north-northwest. Fault is at range front. Photographed in August 2013 (courtesy of R.Q.M. Tābān).

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

92

Berberian and Yeats

preparedness and minimize earthquake risk requirements (iranwire.com, 1 November 2014). Six months later, on 4 May 2015, Mohammad Shekarchizādeh, director of the Road, Housing and Urban Development Research Center (BHRC) of the Ministry of Housing and Urban Development, stated that a major earthquake could cause 1 million deaths and injuries in Tehran. On the same day, ‘Ali Baitollāhi (BHRC) mentioned that at least 300,000 out of 1 million housing units in Tehran are poorly constructed and are very vulnerable to a major earthquake. He added that official construction permits on active faults have been issued by the Tehran Municipality (news.gooya.com, 4 May 2014). On 2 June 2015, Amir Hasan Ziyāee, director of the Iranian Red Crescent Organization, stated that the future earthquake in Tehran will be one of the world’s worst crises (farsnews.com, 2 June 2015). Following the two small-magnitude earthquakes of 13 August 2015 (Ml 4.1), south of the city of Tehran, and 25 August 2015 (Ml 4.6), east of Tehran, Mehdi Chamrān, chairman of the Islamic Assembly of the city of Tehran, declared that the next earthquake in Tehran will be a major historic disaster. He added that neither the authorities nor the people are considering this risk, and numerous newly constructed buildings have collapsed during construction. He added that the dedicated emergency centers of Tehran are not prepared for emergency response, and some of them have been converted to storage depots (farsnews.com; roozonline.com, 26 August 2014). Finally, in his address to the fourteenth General Assembly of the Tehran Construction Engineering Organization (Sāzémān Nezām Kārdāni Sākhtémān Ostān Tehrān) held on 5 January 2016, Director Mehdi Mo’azen stated that: “Governance systems overseeing proper and safe construction standards and approvals have been virtual, not effective, and failed to implement earthquake-resistant building codes in the country and Tehran; this failure resulted in construction of structures that cannot withstand earthquakes” (irna.ir/fa/News/8190842, 5 January 2016; t-nezamkardani.ir, 6 January 2016). Based on the 1996 census, the number of residential buildings in Tehran was 1,484,138 units, composed of: (1) steel structures; (2) reinforced concrete buildings (relatively new largesize buildings in the northern affluent quarter of the city); and (3) older traditional buildings with low-quality material and poor workmanship (SCI, 1997; JICA, 2000). Urban facilities included 59 fire departments, 93 police stations, 16 traffic police stations, 180 hospitals, 88 governmental facilities, 1420 primary schools, 846 intermediate schools, 844 high schools, 225 colleges, and 1194 religious facilities in 2000. About 16.5% of the buildings were built prior to 1966; 19.2% between 1966 and 1975; 42.6% between 1976 and 1988; 20.6% between 1988 and 1996; and 1.1% with unknown dates. The city also had 240 road, railroad, and metro bridges (SCI, 1997; JICA, 2000). Some bridges are near inner-city faults, and the underground metro lines cross some of these faults. Despite establishment of the Iranian Code for Seismic Resistant Design of Buildings (ISIRI Code No. 519, 1969, 1999; ISIRI Code No. 2800, 1988; BHRC, 1988, 1999), most buildings in Tehran are not reinforced to withstand

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

a medium-magnitude earthquake of Mw 6.5. On 26 April 2006, Hamidrezā Vosuqifard of the Engineering Department of the Tehran Islamic Open University reported that their study shows only 2% of the buildings in Tehran may survive a big earthquake, and only 54% of the buildings in the city can be retrofitted (irna .ir, 26 April 2006). On 1 November 2014, Mehdi Hasani, chairman of the Development Commission of the Islamic Parliament, declared that despite the fact that 70% of the Iranian residential buildings are in serious danger of collapsing during future earthquakes, especially in the megacity of Tehran, the country is not ready to cope with any natural disaster (iranwire.com, 1 November 2014). The lifeline network in metropolitan Tehran (natural gas, water, electricity, telecommunications, and sewer), which crosses numerous fault lines, is vulnerable to strong ground motion and faulting and would suffer enormous damage in a major earthquake. Almost all of the operational processes of the lifelines (including natural gas pressure reduction stations, main valves, and distribution valves) are manual and are not designed to shut down automatically during an earthquake. Other vulnerable facilities are: (1) fire and police stations, hospitals, schools, governmental offices, and prisons; (2) transportation network, including roads, railroads, bridges, and subways; and (3) 1500 hazardous material facilities, including the Tehran Nuclear Research Reactor (TNRC) and Institute of Applied Physics (IAP)/Physics Research Centre (PHRC) (Fig. 4), Ray petroleum refinery, petrol stations, kerosene distributors in each quarter, military arsenals and ammunition storage sites, and factories utilizing hazardous and dangerous substances, which may cause secondary disasters such as fires, explosions, and releases of hazardous nuclear and chemical material creating an environmental catastrophe. Areas of potential landslides, rock avalanches, and steep slopes are located mostly in northern Tehran at the southern foot of the Alborz Mountains, whereas potential liquefaction becomes important farther south, where the groundwater level is very shallow. Houses built on fill material and soft grounds are another hazard in Tehran (Berberian et al., 1985; JICA, 2000). The TNRC (Fig. 4; 35.73°N, 51.38°E), the experimental nuclear reactor of the Atomic Energy Organization, built outside the city prior to urban expansion at Amirābād on lands of the University of Tehran, is a critical facility now located in the city 8 km south of the North Tehran fault system (Fig. 4). The construction of the plant started on 10 October 1960, and the facility became operational on 1 November 1967, prior to any knowledge or consideration of seismicity and active faulting in the Tehran region (Berberian et al., 1985). Another inner-city nuclear facility site run by the Islamic Revolutionary Guard in northern Tehran is the Lavizān-Shiyān (later, IAP/PHRC in Fig. 4: 36°46′15″N, 51°29′59″E) at the foot of the Mahmudieh thrust. The controversial Pārchin Revolutionary Guard nuclear complex (35.52°N,51.77°E), with alleged steel bomb cylinder (IAEA, 2011; Porter, 2015), is located ~20 km southeast of Tehran on the hanging-wall block of the Pārchin fault (Fig. 4). It is, apparently,

Figure 4. Locations of paleoseismologic trenches (stippled rectangles; see text for individual trench results and references) dug across faults in and around Tehran in the Alborz and north Central Iran in the south (see Fig. 1 for location). Symbols and abbreviations are as in Figures 1 and 2. The 1996 city limits of Tehran and Karaj (to the northwest) in the Tehran and Alborz (Karaj) Provinces are shown with diagonal lines. Faults with historical seismicity records are drawn by thicker line (modified after Berberian et al., 1983, 1985; Berberian and Yeats, 1999, 2001; the eastern tip of the Taléqān fault [Annells et al., 1975a, 1975b] is connected to the western tip of the Garmābdar fault [Assereto, 1966] after Nazari et al., 2009). Locations of (i) the five large dams supplying water to Tehran (for more information, see Table 1 and text), and (ii) nuclear facilities (radiation signs) adjacent to active faults are added to the figure. Dates of earthquakes along the Moshā and Pārchin (not constrained) faults are added (see Appendices A and B). NTF—North Tehran fault.

Tehran: An earthquake time bomb

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

93

94

Berberian and Yeats

an implosion testing site as well as site for testing and manufacturing of explosives (iaea.org; isisnucleariran.org), although the recent nuclear agreement with Western countries may change its use (if any). A 250 psi (1723.75 kPa) natural gas pipeline enters the city from Karaj in the west (Fig. 4), hugging the North Tehran fault system, and extending within Tehran parallel to some major faults and crossing many others. The main pressure reduction station, and distribution and main valves of the city’s natural gas system are adjacent to the North Tehran fault system, east of the city of Karaj (40 km west of Tehran; Fig. 4). The distribution network at the city crosses the Niāvarān, Mahmudieh, Dāvudieh, Bāgh-e Fayz (Fig. 2), and numerous other inner-city faults (Berberian et al., 1985; JICA, 2000). Five large hydroelectric dams (Karaj, Lār, Latyān, Māmlu, and Tāléqān) provided ~72% of the city’s water demands in 2010 (Fig. 4; Table 1). The rest of Tehran’s water is provided from deep wells mostly contaminated by cesspools (Asadilour et al., 2012; thrw.ir, 4 September 2013), oil spills from unprotected gasoline station underground storage tanks, car repair shops, and released chemicals from factories. The main water lines from these dams cross numerous active faults and meizoseismal areas of large-magnitude historical earthquakes and are vulnerable to strong ground motion (Fig. 4). Because of these recent and dramatic cultural changes, especially the increase in population and expansion of the metropolitan area, past disaster statistics deduced from historical (pre-1900) earthquakes cannot be used to assess the impact of the next destructive earthquake on the modern cities of Tehran, Ray, and Karaj (Fig. 4). A largemagnitude earthquake today will affect not only a much larger urban population in the provinces of Tehran and Alborz along the southern foot of the Alborz Mountains, but it will also heavily damage Tehran’s antiquated lifelines and infrastructure, and could paralyze the country. In this work, we consider the earthquake hazard to metropolitan Tehran, beginning with the geological and tectonic background, followed by an in-depth analysis of the historical (pre-1900) and 1900–2016 seismicity of the region, utilizing contemporary historic records, and archaeological, geological, morphological, and social information for correcting the historic

seismic data. In this way, we avoid misconception, delete spurious events, and correct the seismic parameters for a better and clearer assessment of each event and provide a useful seismic data set for hazard assessment and mitigation. With numerous archaeological sites in the area (solid triangles in Fig. 2), we reviewed archaeological publications that were not necessarily looking for evidence of earthquakes. This is followed by a systematic description of the active faults in and near the greater Tehran metropolitan region, including analysis of paleoseismic trench study results as well as the historic and modern earthquakes associated with each fault, thus giving the first comprehensive active fault inventory and a corrected seismic database to be used in future paleoseismologic, seismic risk, and hazard minimization studies. Because of the length of our analysis, the article is divided into two parts. The main text covers the corrected database analysis and commentary on paleoseismologic trench studies used in our inventory of active faulting. Detailed analysis and commentary providing information on correct locations and seismic source parameters, including magnitudes, are added as Appendices A through C. The entries in this study, especially those in the main body of the text, should not be used or cited out of context, particularly without consulting the appendices and the reasoning used. Bearing in mind that some original sources are hard to access, assess, and evaluate by earthquake geologists, the reviews and analyses of numerous sources carried out here form a very important data set for the earthquake geology community, seismologists, risk analysis practitioners, national and local agencies, and stakeholders. The objective of this paper is to synthesize all present knowledge about regional earthquakes and faulting of the past to highlight the seismic and fault hazards from numerous seismic faults threatening ~15 million people, mostly living in unreinforced structures with inappropriate old infrastructure in metropolitan Tehran, the city of Karaj, and nearby towns. During this study, we noted that in all the previously published papers on historical seismicity, active faulting, paleoseismologic trench studies, and probabilistic seismic-hazard assessments (PSHA), unrealistic input data have been utilized. Input data such as: (1) erroneous epicentral locations; (2) spurious historical seismic events;

TABLE 1. DAMS SUPPLYING DRINKING WATER TO THE MEGACITY OF TEHRAN (SEE ALSO FIG. 4) Name

Coordinates (°N, °E)

Construction date

Height (m)

Length (m)

Base width (m)

Reservoir capacity 3 (km )

Impound (River)

Adjacent active faults

Amir Kabir (Karaj)

35.57, 51.05

1957–1961

180

390

38

202

Karaj

North Tehran, Moshā, Dāvud

Lār

35.53, 51.59

1974–1982

105

1150

800

960

Harāz

Moshā

Latyān

35.47, 51.40

1963–1967

107

450

99

95

Jājrud

North Tehran, Moshā, Lār, Garmābdar

Māmlu

35.34, 51.47

1990–2010

86

807

465

250

Jājrud

Pārchin, Palangvāz

Tāléqān

36.11, 50.38

2002–2006

109

1111

638

420

Shāhrud

Tāléqān, Moshā

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

Tehran: An earthquake time bomb (3) uncharacterized seismic sources; (4) overestimated historical earthquake magnitudes (7.5 < M < 8.1) at a very short epicentral distance from the center of the city; and (5) unconstrained radiometric dates with overinterpreted paleoseismic events from dubious and mislocated historical earthquakes have already entered the national and international catalogues. This has overestimated the calculated hazard in the Tehran region. We conclude that, like the 1990 Rudbār earthquake (Berberian et al., 1992; Berberian and Walker, 2010), and in spite of large-magnitude earthquakes 30 km northeast of the city in 958 and 1830 (Fig. 5), Tehran is in a seismic gap that could generate a destructive earthquake at any time along the North Tehran fault system, the inner-city faults, and the Pārchin and Pishvā faults (Figs. 2 and 5). If an earthquake of Mw ≥7.0 scored a direct hit on Iran’s largest city, it would be catastrophic. We refer to Tehran as a city with numerous earthquake time bombs, following Yeats (2016), who first used the term for Port-au-Prince, Haiti, prior to the Mw 7.0 earthquake there on 12 January 2010, which was not a direct hit but struck west of the capital city, killing as many as 316,000 people (Hayes et al., 2010).

95

The 17 August 1999, Mw 7.4 Izmit (Kocaeli, ESE Istanbul), Turkey, earthquake, with losses in poorly constructed housing on the eastern outskirts of Istanbul (Yeats, 2015, p. 185–188), may be an analogue for the future of Tehran, unreinforced buildings and corruption in Iran’s and Turkey’s building industry. Istanbul and Tehran are similar cities with a similar earthquake past and future. Unlike Istanbul and Port-au-Prince, Haiti, Tehran is subject to numerous active faults (Fig. 2). In contrast to Tehran, however, Istanbul is now apparently making major preparations against an earthquake expected in the near future (Erdik et al., 2003). Moreover, Istanbul has an active earthquake insurance industry, but Tehran does not have such protection against earthquakes. 2. GEOLOGICAL AND TECTONIC BACKGROUND 2.1. The Alborz Mountains The Alborz Mountains, located immediately north of the megacity of Tehran (Fig. 3), represent a tectonically active

Figure 5. Digital elevation model (DEM; EROS Data Center, USGS) of northern Iran, showing the southern Caspian Sea (green), the Alborz Mountains (brown elevation isopleth south of the Caspian Sea), divided into a W-NW–trending section on west (left) and an E-NE–trending section on east (right), bounded by northern Central Iran to the south. Tehran (T) is located at southern edge of the western section, where the low-relief part of northern Central Iran extends into the Alborz Mountains along a range front formed by the North Tehran fault system, with reverse and left-lateral strike-slip motion. Meizoseismal areas are added of some large-magnitude earthquakes discussed in the paper: 280 BCE, Mw ~≥7.0 Rhagae (unconstrained); 22 December 856, Mw ~7.1 Komesh/Dāmghān (K/D, moderately constrained); 23 February 958, Mw ~≥7.0 Ruyān (similar meizoseismal area as the 1830 Lavāsānāt earthquake); 1–30 May 1177 Ray-Qazvin? (unconstrained); 15 June–13 July 1665, Mw ~6.5 Damāvand (slightly constrained); 27 March 1830, Mw ~7.0–7.4 Lavāsānāt (moderately constrained); 20 June 1990, Mw 7.3 Rudbār (well constrained). D—Damāvand volcano; K/D—Komesh/Dāmghān; Q—Qazvin; R—Rudbār; T—Tehran.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

96

Berberian and Yeats

arcuate, double-verging intracontinental oroclinal fold, thrust, and strike-slip belt in northern Iran, south of the Caspian Sea (Figs. 5 and 6). It overthrusts the South Caspian Basin in the north along the Khazar reverse fault (Berberian 1981, 1983a, 1983b), a remnant of the Paleo-Tethys Ocean; and northern Central Iran in the south along the North Tehran fault system (Tchalenko et al., 1974a), a remnant of the Sevān–Qaradāgh (Arasbārān)– Sabzévār northern Neo-Tethys sliver, which closed during the Late Cretaceous–Paleocene (Berberian, 1983a, 1983b, 1989, 2016). The mountain belt is ~100 km wide and ~1000 km long and curves immediately south of the relatively aseismic South Caspian Basin with oceanic basement (Berberian, 1983a, 1983b; Jackson et al., 2002; Ballato et al., 2015). The structural trend of the Alborz changes from N110°E in the western Alborz to N80°E in the eastern Alborz, with a marked hinge near longitude 52.5°E (Figs. 1 and 6). The Alborz includes summits mainly from +3600 to +4800 m in altitude (m above mean sea level [amsl]; the High Alborz), culminating in Mount Damāvand, a Quaternary volcano ~+5670 m in altitude, which lies approximately in the center of the mountain belt, ~70 km northeast of Tehran (“D” in Figs. 1 and 4). Beginning in the early Paleozoic, the Alborz evolved during several contractional and extensional/transtensional episodes associate with plate movements. The “early Alborz Mountain” was uplifted during the Late Cretaceous–Paleocene collisional orogeny (Stöcklin, 1968, 1974; Berberian and King, 1981; Berberian, 1983a, 1983b, 1989, 2016; Allen et al., 2003; Guest et al., 2006b; Rezaeian et al., 2012). During this time, the northern Central Iran block underthrusted the southern Alborz front during closure of the northern Sevān-Sabzévār Neo-Tethys sliver (Berberian, 1983a, 1983b, 1989, 2016). During the late Eocene– early Oligocene contractional movements, the southern front of the “early Alborz Mountain” continued overthrusting northern Central Iran (Ballato et al., 2011, 2015; Rezaeian, 2008; Rezaeian et al., 2012). A possible back-arc spreading stage above the Sevān–Qaradāgh (Arasbārān)–Sabzévār subduction zone created the Oligocene–Miocene Lower Red and Qom Formation back-arc/interarc basins (Berberian, 1983a, 1983b, 1989, 2016). During enhanced exhumation, the Moshā fault and the North Tehran transpressional duplex evolved by ca. 18 Ma (Ballato et al., 2013, 2015; Landgraf et al., 2013). At present, the Alborz Mountains show evidence of over 50 km of crustal shortening and 15–20 km of thickening over the last ~36 m.y. (Guest et al., 2006b; Motavalli-Anbaran et al., 2011). Recent crustal deformation in the Alborz Mountains constitutes a complex, transpressional oblique-slip deformation zone of strain partitioning onto range-parallel reverse and left-lateral strike-slip faults along the entire range (Fig. 6; Jackson et al., 2002; Guest et al., 2006a, 2006b; Landgraf et al., 2009). There is abundant evidence for recent uplift along reverse faults and lateral shift along left-lateral strike-slip faults in the Alborz Mountains, as expressed in incised river terraces and coastal marine terraces (Berberian, 1983a; Berberian et al., 1992; Jackson et al., 2002; Ritz et al., 2006; Berberian and Walker, 2010; Ballato et

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

al., 2015). The left-lateral strike-slip faults in the Alborz Mountains trend northeast in the east and northwest in the west (Figs. 5 and 6) and nearly east-west in the northern Tehran region (Figs. 4 and 6). Right-lateral strike-slip faulting is also documented in the Alborz (Axen et al., 2001; Guest et al., 2006a, 2006b). Apparently, during the late Miocene, between ca. 9 and 6 Ma, exhumation rates decreased along the southern Alborz and reached minimum values in the early Pliocene, between ca. 6 and 3 Ma (Ballato et al., 2015), which is also marked by Pliocene– Pleistocene conglomerates unconformably covering the folded Miocene beds (Guest et al., 2007). Despite the existence of large klippes of allochthonous nappes overlying autochthonous rocks (Assereto, 1966; Stöcklin, 1974), coseismic strike-slip surface faulting and instrumental seismic data (29 October 1985, 20 January 1990; 20 June 1990, 28 May 2004, and microseismicity) show that deformation is occurring within the basement in the Alborz Mountains (discussed in the following). Notwithstanding the statement by Allen et al. (2003), no evidence of an active décollement surface above the Precambrian basement in the Alborz has been documented since the advent of instrumental seismicity. 2.2. Neogene Regional Change in Kinematics The fault plane solution of the 19 December 1977, Mw 5.9 Dartangal earthquake along the Kuhbanān range-front fault in southeast Iran showed a well-constrained pure right-lateral strike-slip motion compatible with coseismic surface rupture. This observation was incompatible with geological observation

Figure 6. Map showing geodetic evidence for active deformation of the central Alborz Mountains superimposed on active fault map from Figure 9. Symbols as in Figure 2. (Top) Global positioning system (GPS)–based velocity vectors of deformation relative to stable Eurasia, showing that displacements are oblique to strike, leading to reverseslip and left-oblique displacements (based on 54 GPS sites surveyed three times from 2000 to 2008; Djamour et al., 2010). Bottom-right inset as in Figure 2. B—Bastām; BA—Bālirān Paleolithic site 13 km SE of Āmol (see text for its vertical slip rate); K—Karaj; SH— Shāhrud. Faults: AL—Alamut; DA—Dāmghān; ES—Eshtehārd; GA—Garmsār; IP—Ipak; KA—Kashachāl; KE—Kelishom; MAN— Manjil; ND—North Dāmghān; NTF—North Tehran fault system; NQ—North Qazvin; PA—Pārchin; PI—Pishvā; RU—Rudbār; TA— Tāléqān fault. (Bottom) Fault slip rates in mm/yr–1 deduced from block model of Djamour et al. (2010) superimposed on the fault map. Top numbers (without parentheses) are strike-slip rates (positive being right-lateral; negative being left-lateral). Bottom numbers in parentheses are fault-normal slip rates (negative being closing). Inset lower left: Schematic illustration of the present-day active tectonics of the South Caspian Basin underthrusting in the Tālesh (west), Alborz (south), and Kopeh Dāgh (east). It also shows the left-lateral strike-slip component in the Alborz and right-lateral component in the Kopeh Dāgh. The open arrow shows the approximate motion of the South Caspian Basin relative to Iran, and the black arrow shows its motion relative to Eurasia (based on Jackson et al., 2002). The dip directions of the North Alborz and Shāhkuh fault are questionable (see the text). MA— Maydānak segment of the Moshā fault in the west.

Tehran: An earthquake time bomb

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

97

98

Berberian and Yeats

and the youngest slickensides showing a predominantly reversefault mechanism, indicating a pre-Quaternary change in kinematics (Berberian et al., 1979). Comparisons of the direction of maximum horizontal shortening deduced from Neogene axes of folding with the horizontal component of compressional axes deduced from focalmechanism solutions (Berberian, 1976a), and more recently with global positioning system (GPS) horizontal velocity data (Nilforoushan et al., 2003; Vernant et al., 2004; Djamour et al., 2010), clearly show a post-Neogene pre-Quaternary regional change in kinematic direction from NE-SW to nearly N-S compression. Furthermore, studies of the complex tectono-stratigraphic evolution support the idea of major changes in the kinematics of the Alborz (Axen et al., 2001; Jackson et al., 2002; Allen et al., 2003; Guest et al., 2006a, 2006b; Zanchi et al., 2006; Ritz et al., 2006; Ritz, 2009; Landgraf et al., 2009, 2013; Morley et al., 2009; Djamour et al., 2010; Solaymani Azad et al., 2011; Ballato et al., 2008, 2011, 2013, 2015; Landgraf et al., 2013). This major change may have resulted from oblique compression due to the combination of north-northeast convergence of ArabiaEurasia and the northwest motion of the rigid South Caspian Basin (Jackson et al., 2002). Sections of both the North Tehran fault system as well as the Moshā fault north of Tehran (Fig. 4) show the predominant reverse-fault mechanism in the Neogene (Assereto, 1966; Tchalenko et al., 1974a; Berberian et al., 1983, 1985; Moinabadi and Yassaghi, 2007) changing to the present predominantly strike-slip motion (Allen et al., 2003; Bachmanov et al., 2004; Guest et al., 2006a, 2006b; Ritz et al., 2006; Yassaghi and Madanipour, 2008; Landgraf et al., 2009; Solaymani Azad et al., 2011; Ballato et al., 2013). A succession of three faulting regimes in the southern central Alborz within the North Tehran and Moshā fault systems was interpreted by Landgraf et al. (2009) as: (1) a pre-Pliocene dextral transpressional stage under northwest shortening; (2) a Pliocene to Holocene northeast shortening, during which the North Tehran and Moshā faults formed a transpressional duplex; and (3) a young left-lateral transtensional stage inverting the kinematics of the regional structure observed only locally. Ritz et al. (2006) suggested that the beginning of the northwestward motion of the South Caspian Basin relative to Eurasia and/or its clockwise rotation took place in the Pleistocene. This has resulted in modern-day oblique strain partitioning into longitudinal subparallel thrusts and left-lateral strike-slip faults on separate fault systems in the Alborz (Berberian et al., 1992; Jackson et al., 2002; Berberian and Walker, 2010). Zircon and apatite (U/Th)/He dates indicate two pulses of exhumation associated with southwest-directed thrusting across the frontal ramps of the transpressional duplex at ca. 18–14 Ma and 9.5–7.5 Ma (Landgraf et al., 2009; Ballato et al., 2013). A uniform cooling age of ca. 7–6 Ma along the North Tehran thrust system and across the major frontal ramps suggests a third event of exhumation associated with south-directed thrusting and reactivation of the North Tehran thrust (Ballato et al., 2013). Cifelli et al. (2015) concluded that the onset of the modern kinematic regime must have post-

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

dated oroclinal bending and started ca. 3 to 6.0 ? 864.01.15–02.12 Ray (S. Tehran) – + >VII+ 898.01.28– Deleted – – – – – – – 899.01.16 898.02.24 Deleted – – – – – – – Ray, Dizeh Qasrān; >VIII ≥7.1 ≥7.0 Moshā 958.02.23 Ruy n (NE 35.81, 51.76? ++ Tehran) landslide blocking rivers; change in the Caspian Sea water level (?) Ca. 1000 CE: Invasion by the Saljuq Turk Hordes—Destruction of infrastructure. Major gap in seismic data. 1175.07.22 Deleted – – – – – – 1176 Deleted – – – – – – 1177.05.01–30

Ray-Qazvin?†

36.10, 50.43?

++

Ray, Qazvin, many towns & villages

IX

7.1

7.0

Western North Tehran?

Q

C

S S S S C C S S B

S S C

1220–1221: Invasion by the Mongol Hordes—Destruction of infrastructure and documents. Major gap in seismic data. 1316–1335: Iran lost its cohesion during collapse of the Mongol Ilkh nid dynasty. 1370–1375; 1386–1388; 1399–1405: Invasion by Timur (Tamerlane) Turco-Mongol Hordes—Destruction of infrastructure and documents. Major gap in seismic data. 1384 Deleted – – – – – – – S Damāvand (E. C 1665.06–07 35.75, 52.08? VIII 6.5 6.5 Moshā Tehran) 1786.04.15 Deleted – – – – – – 1830.03.27 Lavāsānāt (NE ++; >500 in Damāvand, Jājrud & B 35.81, 51.76? IX 7.1 7.0 Moshā Tehran) Damāvand, 30 many villages in Tehran C 1830.04.06 Lavāsānāt AFS Jājrud, additional >6.1 >6.1 Moshā 35.76, 51.96? ? >VII+ (NE Tehran) damage Note: Bold text indicates earthquakes recorded by contemporary and near-contemporary sources. Deleted—Spurious and alleged events deleted after scrutinizing the existing catalogues. Not confirmed by any contemporary or near-contemporary sources. Casualties: ++—very high; +—many killed. AFS—strong aftershock. ~I—approximate highest intensity (modified Mercalli intensity [MMI]) estimated within the meizoseismal area based on written accounts. ~Ms—the equivalent surface-wave magnitude of historical (pre-1900) earthquakes mentioned throughout the text are estimates taken from Berberian (1981, 1994, 2014) and Ambraseys and Melville (1982). Pre-instrumental magnitudes were derived from macroseismic information embedded in written accounts calibrated against instrumental Ms values based on regional twentieth-century earthquakes. Therefore, they may represent poorly constrained magnitude for some historical events. Mw*—the equivalent moment magnitude based on global regression relations (Das et al., 2011) and Middle East regression relations (Zare et al., 2014). Macroseismic epicenter— center of affected (meizoseismal) area estimated from written annals. It is not the real location of the epicenter, and in some cases, it is based on personal judgment. They should be corrected by paleoseismic trench studies. Q—quality of macroseismic data: B—earthquake known from its effects at a number of locations and from extent of the effected regions; C—earthquakes known only from effects at one or two locations, where the shock caused damage and the location shown is not an epicenter; D—effects of events only given in general terms with unidentifiable epicentral location; S—spurious event. *Amalgamated with the 22 December 865 Komesh (modern Dāmghān) earthquake (see Appendix B for commentary on this entry). † Could be two separate earthquakes in 1177 at Ray and Qazvin (see Appendix B for commentary on this entry).

Existing macroseismic data suggest that the two events of 958 and 1830 CE (both of Mw ~≥7.0) took place along the central section of the Moshā fault, within a recurrence interval of ~872 yr (with only two earthquakes) for the segment (Figs. 5 and 7); however, this has not yet been confirmed by paleoseismic study. Between these two large-magnitude events, an earthquake of M ~6.5 took place in 1665 CE along the eastern section of the Moshā fault, ~40 km to the southeast of the two men-

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

tioned events (Fig. 5; Table 2). We do not have enough macroseismic intensity data to locate the seismic faults of the 855–856 (M ≥7.0) and 864 (M ~6.0) earthquakes, which destroyed Ray and Tehran (Fig. 7; Table 2). Neither of these two events could have happened along the central segment of the Moshā fault, since the segment was reactivated during the 958 and 1830 M ~≥7.0 earthquakes (Fig. 7; Table 2). We may speculate that at least one of the events took place along the North Tehran fault

Tehran: An earthquake time bomb

101

Figure 7. Disproportionate growth in population and area (lower right), major recorded historical earthquakes (vertical bars), and fault map (inset top left) in Tehran (stippled). Only the population of the city (excluding Shemirān/Tajrish to the north [SH in the figure] and Ray to the south) is given; the total urban population estimate in 2014 was 12–14 million. The pre-1900 population estimate on the graph is shown by open squares, whereas the rest is shown by filled circles. The ancient city of Rhagae (modern Ray) is located southeast of Tehran (inset top left). Rhagae (Ray) and later Tehran have been shaken by at least seven recorded large-magnitude earthquakes since ca. 280 BCE (see Appendices A through C). Symbols and top-right inset are as in Figure 2.

system and the other one along the Pārchin or northern tip of the Pishvā faults (Fig. 4); however, we do not have any data to support these scenarios. We only know that Ray was damaged or destroyed by these two events (Fig. 7). The seismic source of the 1177 M ≥7.0 earthquake, which destroyed the cities of Ray and Qazvin, might have been the western segment of the North Tehran fault system, if these cities were not destroyed or damaged by two separate events in the same year (Fig. 5). Nonetheless, these scenarios and explanations remain speculative and require further confirmation by well-organized paleoseismic

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

trench studies. During the recorded seismic history, no evidence of rupturing along the Niāvarān and Darakeh left-lateral strikeslip faults, the central segment of the North Tehran thrust, or the inner-city thrust faults of Mahmudieh, Dāvudieh, and others has been documented (Fig. 2). All these earthquakes (Fig. 7), some of them outside the city (Moshā fault in 958 and 1830; Fig. 4), occurred when Tehran was a small village (and later a small town) north of the major ancient city of Rhagae/Ray, with a significantly smaller population than today (Fig. 2). The historical earthquake sequence in

102

Berberian and Yeats

Figure 8. Historical (pre-1900: hexagons) and instrumental period (1900–2015; circles) seismicity and fault map in a radius of 50 and 100 km around center of metropolitan Tehran (hatched). Centroid moment tensor (CMT) solutions (combined inversion of broadband and short-period waveform) of three small-magnitude earthquakes with their centroid depth along the Moshā fault are taken from Donner et al. (2014). Symbols and top-right inset are as in Figure 2. See Figure 1 for location. For individual seismic data, see Tables 2 and 3. Events are discussed in the text as well as Appendices A through C. NTF—North Tehran fault system.

the Tehran region occurred over a period longer than two millennia, since ca. 280 BCE (Fig. 7; Table 2). At the time of the last major earthquake of 27 March 1830, along the central segment of the Moshā fault, ~30 km northeast of downtown Tehran, Tehran was home to only 70,000–100,000 people, and its population was confined to an area of ~4 km2 (Fig. 2). Since then, Tehran has grown more than 100 times in population, and more than 217 times in area. The city has expanded in all directions, and isolated villages have become residential neighborhoods of Greater Tehran. Efforts to accommodate this growth have resulted in a pro-

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

liferation of poorly constructed buildings and urban development along seismically active faults and along folds that may overlie blind reverse faults and in areas subject to landslides, liquefaction, subsidence, and flooding by failed dams. 4. INSTRUMENTAL SEISMIC DATA (1900–2015) Instrumental epicenters of the Iranian earthquakes, especially those of the 1900–1970 time period, are incomplete, inhomogeneous, and suffer from large errors of location and focal

Tehran: An earthquake time bomb depth (Ambraseys, 1978; Berberian, 1979a) and are not reliable for seismotectonic and earthquake hazard studies. Nonetheless, no earthquake of M >5.0 has occurred within a radius of 3.0, although, the 1977 and 1988 earthquakes might have happened along the fault (Table 3). Parameters of the recent Alborz earthquakes determined by body-wave modeling indicate shallow thrust and left-lateral strikeslip faulting with centroid depths ranging from 8 to 13 km (Jackson et al., 2002), except for a slightly greater centroid depth of 22 km for the 28 May 2004 Mw 6.2 Firuzābād Kojur (Baladeh) earthquake possibly along the Khazar reverse fault (Berberian, 1981, 1983a, 1983b) in basement (Tatar et al., 2007a, 2007b; Djamour et al., 2010; Donner et al., 2013). Recent seismic evidence in the Alborz such as: (1) the 29 October 1985 (Mw 6.1, centroid depth 13 km; Priestley et., 1994); 20 June 1990 (Mw 7.3, centroid depth 13 km; Berberian et al., 1992; Jackson et al., 2002; Berberian and Walker, 2010), and 28 May 2004 (Mw 6.3, centroid depth 22 km; Tatar et al., 2007a, 2007b) earthquakes; and (2) microseismicity studies from temporary networks around the Tehran and Firuzkuh-Āstāneh area to the east with leftlateral and thrust faulting with depths ranging from near-surface to 20 km (Ashtari et al., 2005; Nemati et al., 2010, 2011, 2013; Tatar et al., 2012), all indicate that the Alborz basement is involved in

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

103

active deformation. The seismic data coupled with active left-lateral strike-slip faulting throughout the Alborz Mountains (Fig. 9) provide further evidence that the Alborz is a thick-skinned orogen. The seismic history of Tehran (Table 2; Fig. 8; Appendices B and C) suggests that the Niāvarān, Darakeh, Farahzād, and Kan left-lateral strike-slip members of the North Tehran fault system, as well as the central-eastern segment of the North Tehran thrust, north, northwest, and northeast of metropolitan Tehran, the Mahmudieh, Dāvudieh, and other inner-city reverse faults (Fig. 4), and the Pārchin and Pishvā faults to the southeast (Fig. 9) have not shown historical activity, suggesting that the city is located in a seismic gap. This relation is used to highlight the potential of active faults in and around Tehran to generate future destructive earthquakes. 5. INTENSITY, MAGNITUDE, MEIZOSEISMAL AREA, AND MCE ESTIMATE DILEMMA FOR THE EARTHQUAKES IN THE ALBORZ AND TEHRAN Assessment of macroseismic data of historical earthquakes documented in old Persian, Armenian, Arabic, Syriac, and European chronicles is difficult and often controversial, which could result in over- and/or underestimation of earthquake parameters (Ambraseys and Melville, 1982; Berberian, 1994, 2014; Ambraseys, 2009). Most catalogues of historical earthquakes in Iran are unreliable due to: (1) inconsistent, incomplete, and heterogeneous data containing erroneous and dubious events; (2) employment of secondary, tertiary, and later unreliable sources; (3) erroneous dates, times, and locations of some events; (4) duplication of the dates of earthquakes; (5) presence of numerous misinterpretations, mislocations, and misspellings; (6) overestimation or underestimation of intensity, magnitude, and meizoseismal areas; (7) failure to cite the sources of information; (8) lack of uncertainty estimates of the seismic parameters; and (9) unrealistically high numbers of reported death tolls (Ambraseys, 1974; Ambraseys and Melville, 1982; Berberian, 1994, 2014; Guidoboni and Traina, 1995; Ambraseys, 2009; Bilham, 2009). Consequently, some spurious earthquakes and exaggerated seismic parameters have entered into global catalogues (U.S. Geological Survey (USGS), 2014; National Oceanic and Atmospheric Administration [NOAA], 2014; Utsu, 2002, 2014; among many) and been analyzed for the quantification of world fatality estimates and seismic risk evaluation studies. We return to these issues in the following sections and in the appendices at the end of the report. Despite an improved catalogue of the Iranian historical earthquakes from archival research (Ambraseys and Melville, 1982; Berberian, 2014), examples still exist of highly exaggerated death tolls, overestimated meizoseismal areas, overrated intensities and magnitudes, and lack of information about the source faults of the large-magnitude earthquakes. There have been controversial statements about the number of historical earthquakes in the Ray-Tehran region. Based on historical chronicles, Ambraseys (1974) reported that the ancient city of Ray was destroyed five times between the eighth and

104

Berberian and Yeats

TABLE 3. TELESEISMIC DATA FOR THE MODERN INSTRUMENTAL ERA (1930 THROUGH 2015) EARTHQUAKES IN A RADIUS OF 100 KM AROUND CENTRAL TEHRAN: 35.70°N, 51.40°E (SEE FIGS. 8, 9, AND 17; AND APPENDIX C FOR DISCUSSION) Date Origin time Geographic Magnitudes F.D. No. RMS Source Seismic fault (yr.mo.d) (GMT, coordinates (km) stations (s) (this study) Ms Mw* Ml mb h:min:s) (lat °N, long °E) 1930.10.02 15:33:12 35.80, 52.10 5.0 5.1 032 ISS (35.75, 51.98) (ME) Moshā 1930.10.07 20:53:06 35.80, 52.10 5.0 5.1 017 ISS (35.75, 51.98) (ME) Moshā 1945.05.11 20:17:28 34.80, 52.10 4.7 4.6 015 ISS (35.32, 52.41) (ME) Garmsār 1951.04.22 06:32:41 34.80, 52.10 5.0 5.0 031 ISS Siāhkuh? 1954.09.02 22:47:00 35.30, 52.00 4.5 4.5 SSK Garmsār 1955.04.08 21:58:00 36.00, 52.00 SSK 1955.11.24 – (35.75, 52.05) 4.0 3.9 (ME) Moshā 1957.05:06 15:06:50 36.45, 51.30 5.5 5.5 072 3.90 ISS (36.13, 52.07) (ME) 1959.05.01 08:24:04.19 36.48, 51.29 5.0 5.4 35f 124 3.09 ISS 1963.07.05 23:49:34 35.80, 51.50 4.1 4.3 BCIS 1966.11.08 03:14:12.8 36.10, 50.74 4.8 5.0 41 055 1.50 ISC Tāléqān 1967.02.16 11:55:32 35.40, 51.90 4.5 4.7 144 009 3.60 ISC 1970.06.27 07:57:58.20 35.12, 50.80 4.8 5.0 52 082 2.40 ISC 1970.10.03 06:57:03.5 36.01, 51.31 4.1 4.3 78 013 1.55 ISC Moshā 1974.01.10 16:36:19.7 35.800, 51.950 4.3 4.5 33N 023 2.00 ISC Moshā 1975.10.17 06:58:54.2 36.153, 51.819 3.0 32 005 0.80 NEIS 1977.05.14 19:15:51.8 35.847, 52.418 1G 006 1.60 NEIS NTF 1979.02.22 05:07:17.79 35.201, 52.099 4.6 4.8 33f 009 1.48 ISC Garmsār 1981.08.04 18:53:59.79 36.445, 51.268 4.7 4.9 0.0f 009 0.84 ISC 1982.10.25 16:54:51.3 35.208, 52.355 4.6 4.8 33N 042 1.00 NEIS Garmsār 1983.03.25 11:57:46.71 36.132, 52.361 4.1 4.9 5.4 10f 246 1.44 ISC (35.98, 52.22) (ME) Bāijān 1983.03.26 04:07:19:86 36.057, 52.279 5.4 4.9 33f 261 1.31 ISC (35.98, 52.22) (ME) Bāijān 1983.05.29 17:15:40.13 35.242, 52.169 4.4 4.6 39 053 1.08 ISC Garmsār 1988.01.14 11:29:20.2 36.010, 50.600 4.6 4.8 33N 006 1.40 NEIC NTF or Moshā? 1988.08.22 21:23:35.80 35.348, 52.378 5.0 5.0 5.4 18 255 1.25 ISC Garmsār?† 1988.08.23 05:30:51.35 35.419, 52.275 5.0 4.8 5.2 34 249 1.14 ISC Garmsār 1988.08.23 10:58:09.91 35.337, 52.344 4.6 4.6 5.1 16 102 1.32 ISC Garmsār?† 1988.08.23 14:56:08.05 35.640, 52.402 4.0 4.3 10f 008 1.27 ISC Moshā? 1988.10.24 17:01:58.89 35.254, 52.295 4.9 4.4 5.0 34 182 1.38 ISC Garmsār 1988.10.26 14:49:20.01 35.119, 52.228 4.7 4.9 06 029 1.28 ISC Garmsār 1988.12.08 18:40:59.80 35.149, 52.204 4.4 4.5 10f 034 1.13 ISC Garmsār 1990.01.01 18:36:21.9 35.120, 50.720 10G 005 1.20 NEIC 1990.01.20 21:21:14.3 35.590, 51.440 10G 005 1.20 NEIC 1990.01.21 13:39:40.7 35.920, 51.380 10G 005 0.30 NEIC 1990.01.21 21:28:01.9 35.330, 51.830 10G 005 0.10 NEIC 1990.01.23 19:42:37.9 35.298, 50.744 5G 005 1.80 NEIC 1990.01.25 11:39:26.0 35.540, 51.720 10G 005 0.80 NEIC 1990.01.28 13:54:28.0 35.540, 51.720 10G 005 0.80 NEIC 1991.01.22 12:04:25.35 35.440, 52.322 4.5 4.7 33f 041 0.94 ISC 1993.03.08 19:13:24.09 36.495, 51.024 4.4 4.6 56 033 0.76 ISC 1993.05.12 09:41:17.0 36.390, 51.980 4.3 4.5 33N 012 1.40 NEIC Pārchin 1993.08.19 10:04:28.8 35.090, 52.090 4.6 4.8 18D 067 1.00 NEIC 1994.11.21 18:55:18.22 36.050, 51.913 4.5 3.7 4.7 43 066 1.21 ISC 1997.11.05 22:42:57.98 34.940, 51.371 4.2 4.4 43 026 0.90 ISC 1998.07.07 11:06:38.48 36.212, 52.079 3.6 3.9 33f 006 1.03 ISC 1998.08.30 15:12:55.05 36.096, 52.124 3.6 3.9 33f 005 0.76 ISC Bāijān 1998.12.03 13:13:34.82 36.063, 50.972 4.3 3.6 4.5 46 077 1.25 ISC 1988.12.03 21:07:18.39 36.302, 51.112 3.9 4.2 100f 008 1.29 ISC 2001.06.24 07:05:18.21 35.838, 52.184 3.6 3.9 04f 007 1.08 ISC 2002.01.12 05:18:22.70 35.608, 50.555 3.2 3.6 33f 005 0.02 ISC 2002.04.08 18:30:59.56 36.365, 52.017 4.8 4.1 4.8 57 185 1.15 ISC 2002.05.21 10:48:34.76 36.292, 51.547 4.1 3.3 4.3 12 032 1.08 ISC 2002.10.10 12:13:40.97 35.823, 52.249 4.5 4.7 15 090 1.21 ISC 2002.10.15 13:59:34.59 35.838, 52.232 3.7 3.4 4.3 33f 014 1.24 ISC 2002.10.15 16:56:08.10 35.820, 52.223 4.0 3.5 4.4 10f 024 0.91 ISC 2004.02.21 12:07:31 35.740, 52.270 3.9 14–16 D+ Moshā 2004.05.28 12:38:43.04 36.321, 51.587 6.2 6.3 6.2 17f 1345 0.99 ISC Khazar 2004.05.28 13:07:07.43 36.418, 51.717 3.8 4.1 10f 028 1.04 ISC Khazar 2004.05.28 13:15:06.37 36.450, 51.514 4.4 4.8 5.2 20 103 1.03 ISC Khazar 2004.05.28 13:28:41.17 36.363, 51.737 3.3 3.7 10f 006 1.40 ISC Khazar 2004.05.28 13:35:52.95 36.459, 51.576 3.9 4.2 3.3 053 1.07 ISC Khazar 2004.05.28 19:47:02.71 36.488, 51.341 4.5 4.1 4.8 20 160 1.11 ISC Khazar 2004.05.28 20:32:53.79 36.568, 51.147 3.4 10f 006 0.90 ISC Khazar 2004.05.29 04:12:33.02 36.442, 51.694 3.6 3.9 10f 027 1.06 ISC Khazar 2004.05.29 04:52:59.28 36.514, 51.418 3.4 10f 018 0.93 ISC Khazar 2004.05.29 09:23:47.19 36.544, 51.388 4.8 4.6 5.1 12 254 1.20 ISC Khazar (Continued)

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

Tehran: An earthquake time bomb

105

TABLE 3. TELESEISMIC DATA FOR THE MODERN INSTRUMENTAL ERA (1930 THROUGH 2015) EARTHQUAKES IN A RADIUS OF 100 KM AROUND CENTRAL TEHRAN: 35.70°N, 51.40°E (SEE FIGS. 8, 9, AND 17; AND APPENDIX C FOR DISCUSSION) (Continued) Date Origin time Geographic Magnitudes F.D. No. RMS Source Seismic fault (yr.mo.d) (GMT, coordinates (km) stations (s) (this study) Ms Mw* Ml mb h:min:s) (lat °N, long °E) 2004.05.29 11:01:28.52 36.532, 51.412 4.0 3.8 4.6 10f 058 0.95 ISC Khazar 2004.05.29 15:40:59.97 36.545, 51.429 3.5 3.8 10f 016 1.03 ISC Khazar 2004.05.29 17:30:23.98 36.528, 51.457 3.7 4.0 13 010 1.18 ISC Khazar 2004.05.29 18:38:04.44 36.524, 51.361 4.3 3.9 10 109 1.11 ISC Khazar 2004.05.29 18:42:42.06 36.487, 51.425 3.7 4.0 11 015 0.93 ISC Khazar 2004.05.29 22:55:15.17 36.541, 51.415 3.8 4.1 10f 036 1.07 ISC Khazar 2004.05.30 01:42:39.50 36.488, 51.558 4.2 3.5 4.4 25 141 1.14 ISC Khazar 2004.05.30 13:09:50.71 36.563, 51.404 4.1 3.5 4.4 10f 062 1.20 ISC Khazar 2004.05.30 19:27:00.83 36.455, 51.556 3.7 4.0 27 188 1.04 ISC Khazar 2004.05.31 22:05:32.26 36.558, 51.311 3.4 10f 004 1.05 ISC Khazar 2004.06.12 06:43:39.46 36.474, 51.542 3.1 10f 009 0.79 ISC Khazar 2004.06.07 04:01:18.78 36.519, 51.385 4.0 3.4 4.3 14f 052 1.41 ISC Khazar 2004.06.12 21:30:58.79 36.511, 51.571 3.5 3.8 10f 007 0.63 ISC Khazar 2004.09.24 01:42:48 36.657, 52.405 4.1 4–6 D+ Moshā 2004.07.11 13:16:47.05 36.532, 51.616 3.4 3.7 33f 009 0.56 ISC Khazar 2006.02.22 23:19:25.73 36.085, 50.420 3.1 14f 045 1.39 ISC 2006.12.20 04:39:21.1 35.790, 51.926 4.2 3.6 10 0.3 TEH Moshā 04:39.20 35.802, 51.915 3.8 14 D+ Moshā 2007.06.04 08:04:15.76 36.445, 51.374 3.6 3.9 3.5 061 1.20 ISC Khazar 2007.08.25 10:07:36.46 34.926, 51.952 4.3 3.5 4.4 21 203 1.01 ISC 2009.08.13 13:57:43.19 36.371, 52.003 4.1 3.3 4.2 09 130 1.53 ISC 2009.08.15 00:28:37.89 36.437, 51.994 3.9 4.2 03 140 1.37 ISC 2009.10.17 10:53:56.5 35.561, 51.498 4.0 4.3 12 0.40 TEH Pārchin 2010.03.09 03:22:14.09 36.471, 51.862 3.5 3.8 07f 099 1.93 ISC Khazar 2011.02.20 11:22:16.26 35.400, 51.847 4.0 10f 108 1.56 ISC Pārchin 2011.10.07 00:39:12.17 35.233, 51.909 15f 053 2.32 ISC Garmsār 2011.11.20 13:19:08.11 35.223, 51.906 3.4 6.8 044 0.93 ISC Garmsār 2012.02.10 08:59:41.91 35.542, 52.426 4.3 3.4 4.3 15 295 1.80 ISC 2012.07.25 11:52:17.76 36.031, 52.148 3.7 4.0 11f 085 1.37 ISC Bāijān 2012.08.17 20:31:27.18 35.833, 51.874 3.7 4.0 8.7 084 1.31 ISC 2015.04.06 18:19:14.4 35.197, 52.253 3.0 8.5 0.40 TEH Garmsār 2015.08.13 18:41:49.2 35.152, 51.925 3.4 14.2 0.50 TEH Pishvā 2015.08.13 18:42:13.0 35.145, 51.895 4.1 4.3 14.0 0.60 TEH Pishvā 2015.08.13 18:57:37.8 35.150, 51.940 3.6 3.9 11.3 0.50 TEH Pishvā 2015.08.14 06:28:30.0 35.140, 51.942 3.4 8.8 0.50 TEH Pishvā 2015.08.25 17:36:33.7 35.577, 52.628 4.6 4.8 10 TEH 2015.08.25 14:40:22.2 35.570, 52.613 3.2 10 TEH Note: MI—local magnitude; mb—body-wave magnitude; Ms—surface-wave magnitude; Mw—moment magnitude. F.D.—focal depth; f— depth fixed; G—depth was constrained by a geophysicist; N—default depth was restrained at 33 km for earthquakes for which the character on seismograms indicated a shallow focus but for which the depth was not satisfactorily determined; D—depth was restrained by the computer program based on two or more compatible pp phases, and/or unidentified secondary arrivals used as pp. RMS—root mean square. Sources: D+—Donner et al. (2014); BCIS—Bureau Central International de Seismologie, Strasbourg, France (1952 onward with delta ~50° of central Europe); ISS—International Seismological Summary, Berkshire, UK (1913–1963; now ISC); ME—macroseismic epicenter, i.e., center of the meizoseismal area with highest recorded epicentral intensity; MOS—Moscow, Institute of Physics of the Earth, USSR; NEIC—National Earthquake Information Center, U.S. Geological Survey (USGS), Denver, Colorado (1985–onward); NEIS—National Earthquake Information Service, USGS, Denver, Colorado (1964–1985). No. stations—number of stations used in location program. NTF—North Tehran fault. SSK— Savarensky et al. (1962). TEH—Institute of Geophysics, University of Tehran. *Moment magnitude based on global regression relations (Das et al., 2011) and Middle East regression relations (Zare et al., 2014). † The epicenters are located near the junction of the Garmsār and Pārchin faults (Fig. 9).

tenth centuries CE. After scrutinizing the existing parametric and descriptive catalogues and their sources (discussed in the appendices), we were able to document only three genuine earthquakes during this short time period (Table 2; Fig. 7). Nazari et al. (2005) located 12 destructive historical earthquakes in the Tehran-Ray region: three earthquakes (898, 7 July 1847, 30 August 1868) in eastern Tehran, and nine earthquakes (third century, 743, 793, 853, 855–856, January 864, May 1177, 1384, and 1895) at the city of Ray in southern Tehran. Intensity estimates and the equivalent surface-wave magnitude derived from the reported historical macroseismic data are generally calibrated against the instrumental Ms values of the twentieth century (see, for example, Ambraseys and Melville, 1982; Ambraseys, 2009). Such conclusions should be taken

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

cautiously and reexamined in light of tectonic, demographic, and archaeological data (discussed in the text and the appendices). The meizoseismal areas of such events are commonly drawn with a simple elliptical shape of the highest estimated MMI isoseismal. Simultaneity of historical earthquakes being felt or resulting in damage in various places over a short period of time in the chronicles has been misleading. Later accounts may add more localities with damage estimates to those addressed earlier. For example, the magnitudes of some pre-instrumental earthquakes in the study area (ca. 280 BCE [Ms 7.6+], 856 [Ms 7.9], 958 [Ms 7.7], 1608 [Ms 7.6]) appear to have been exaggerated by Ambraseys and Melville (1982; though this is otherwise a valuable and carefully researched source), and these have then been used

106

Berberian and Yeats

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

Tehran: An earthquake time bomb uncritically in later reports and international catalogues. Generally, most authors do not assess the seismological and geological aspects of the reported historical earthquakes in the parametric and descriptive catalogues or in seismic hazard evaluations. New authors, who do not have access to the historic sources or the knowledge of languages used in the sources, accept much of the data introduced in previous catalogues without further inquiry. The assessment of meizoseismal areas, grossly exaggerated magnitudes, and spurious events of early earthquakes have been used in numerous reports, and national and international databases at face value by scientists and engineers in their seismic risk evaluation and design parameters. These issues are discussed herein for each event affecting Tehran to derive a better estimation of the maximum credible earthquake (MCE) for the Alborz that would be applicable for metropolitan Tehran (see discussions in the appendices). Mega-earthquakes of Mw 8.05–8.29 (Nazari et al., 2014), Ms 8.0 (Gorshkov et al., 2009), Mw 7.7–7.9 (Nazari et al., 2009; Asadi and Zare, 2014), and Mw 7.6 and 7.7 (Nazari, 2006; Vasheghani Farahani et al., 2014) have been recently proposed for the Alborz earthquakes.

Figure 9. (Top) Fault map and documented major historical earthquakes of central Alborz Mountains north of Tehran (hatched) and northern Central Iran (south of Tehran), with faults shown in irregular, solid lines. Faults with arrows are predominantly strike-slip; faults with solid triangles on upthrown side are predominantly reverse faults. F— Firuzkuh. Abbreviations of faults: AL—Alamutrud; DA—Dāmghān; ES—Eshtehārd; GA—Garmsār; IP—Ipak; KA—Kashachāl; KAN— Kandévān; KE—Kelishom; MA—Maydānak segment of Moshā; MAN—Manjil; ND—North Dāmghān; NQ—North Qazvin; NTF— North Tehran fault system; PA—Pārchin; PI—Pishvā; RU—Rudbār; SH—Shāhrud; TA—Tāléqān. Fault map extends from the 1990 Mw 7.3 Rudbār earthquake on west (far left) to 856 Dāmghān (Komesh) earthquake on east (far right). Dashed line ellipses with dot pattern mark meizoseismals of earthquakes with year of earthquake given. Meizoseismal areas of the 280 BCE Rhagae (Ray), 1119 Qazvin, and 1177 Ray-Qazvin are not constrained. Beach balls with dates show fault-plane solutions of more recent earthquakes (1957.07.02 Mw 7.1 Band-e Pay—Shirokova, 1962, McKenzie, 1972; 1962.09.01 Mw 7.0 Bu’in—Priestley et al., 1994; 1983.03.25 Mw 5.5 Bāijān—Harvard CMT, 2016; 1983.07.22 Mw 5.5 Charazeh—Priestley et al., 1994; 1985.10.29 Mw 6.1 Nomal—Priestley et al., 1994; 1990.01.20 Mw 6.0 Gaduk—Harvard CMT, 2016; 1990.06.20 Mw 7.3 Rudbār—Campos et al., 1994; 1992.09.22 Mw 5.0—Harvard CMT, 2016; 2002.06.22 Mw 6.4 Changureh—Walker et al., 2005; 2004.05.28 Mw 6.2 Firuzābād Kojur-Baladeh—Tatar et al., 2007a, 2007b). Diagonal line pattern shows urban areas of modern Tehran and city of Karaj (“K,” west of Tehran). Solid triangle with letter D—Damāvand Quaternary volcano in central Alborz. VA (15 km SW of Chālus coastal town)—Valasht Lake. (Bottom) Space-time diagram of seismicity (zone of extensive damage shown by horizontal lines) of the central Alborz and northern Central Iran with dates and magnitudes (~Ms for pre-instrumental earthquakes); queried where uncertain. Where date of the earthquake is shown, it is given by year.month.day. Distances are along strike with respect to scale of map. For some individual events, see the text and Appendices A through C. The dip directions of the North Alborz and Shāhkuh fault are questionable (see the text).

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

107

We went through a lengthy and painstaking analysis of almost all of the source materials. We scrutinized the database, used first-hand contemporary and near-contemporary sources, demography, and regional tectonics, as well as archaeologic data. We then eradicated spurious, dubious, and duplicate events and stated the reasons for doubting the validity of some data. We, therefore, graded the historical seismic data according to the quality and quantity of the macroseismic data (see Table 2). Earlier catalogue entries containing insufficient and incorrect information to assign intensity, magnitude, date, and geographic location were deleted from our catalogue (see the commentaries in the appendices). Despite all the efforts, some events still contain insufficient information to assign reliable intensity, magnitude, meizoseismal area, and causative fault data (Table 2; Fig. 9). 6. ACTIVE FAULTING Several active faults located in the southern Alborz and the northern Central Iranian Plateau have recorded large-magnitude historical earthquakes and pose a major seismic hazard to the megacity of Tehran. These faults show both reverse and strikeslip mechanisms (Fig. 9). The left-lateral strike-slip motion along the North Tehran fault system (the Niāvarān, Darakeh, Farahzād, and Kan segments; Fig. 2), and Moshā, Tāléqān, Firuzkuh, Āstāneh, and Dāmghān faults (Fig. 9) is a component of transpressional deformation along the Alborz Mountains (Jackson et al., 2002; Ritz et al., 2003, 2006; Ritz, 2009 [transtension]; Landgraf et al., 2009; Solaymani Azad et al., 2011). Slip partitioning in and around Tehran can be seen based on subparallel active reverse and strike-slip faults (Figs. 2, 8, and 9). The slip rates and mechanisms along these faults are important in understanding how tectonic strain is accommodated within the ArabiaEurasia collision zone (Fig. 6) and in characterizing the recurrence periods of large-magnitude earthquakes around Tehran. Except for the 515-km-long range-front south-dipping Khazar reverse fault (Berberian, 1981, 1983a, 1983b), with a sharp bend in the middle (240-km-long NW-trending, and 275-km-long NE-trending segments), extending along the northern foothills of the Alborz Mountains (Fig. 9), no major continuous long fault has been mapped in the Alborz. Notwithstanding the presence of active neighboring faults in the Alborz, such as the Moshā, North Alborz, and Shāhkuh faults (Fig. 9), no evidence of rupturing of these faults along their entire length in single large-magnitude earthquakes (i.e., Mw ≥7.5) has been documented (Tables 2 and 3; Fig. 9). The Khazar reverse fault (Berberian, 1981, 1983a, 1983b), along which South Caspian basement underthrusts the northern Alborz front, has been associated with some medium-magnitude historical earthquakes, such as the 874, 1436, and 1498 (in Jorjān/ Gorgān; modern Gonbad Kāvus, SE of the Caspian Sea) and the 1809 (Mw ~6.5 Āmol, south of the Caspian Sea) earthquakes (Fig. 9). Most Iranian continental-crust earthquakes are shallow, with centroid depths around 10 km (Jackson et al., 2002). P and SH waveform analysis of the 28 May 2004 Mw 6.2 Firuzābād-e Kojur

108

Berberian and Yeats

(Baladeh) earthquake in the western Alborz (Fig. 9) showed a centroid depth of 22 km with a thrust mechanism and a damage zone located ~25 km south of the Khazar fault line (Tatar et al., 2007a, 2007b; Donner et al., 2013). The centroid depth, together with the aftershock distribution (Tatar et al., 2007a, 2007b), revealed a south-dipping thrust plane projected to the surface on the hanging wall of the Khazar reverse fault (Fig. 9) and casts doubt on the active reverse faults in the Alborz Mountains being thin-skinned features. Despite the societal importance of studying the major active faults in the central Alborz Mountains and northern Central Iran threatening major urban areas such as the megacity of Tehran (Fig. 9), little has been achieved in evaluating the seismic hazard of faults in these areas, and our seismotectonic understanding of them is very limited. For example, a recently published 1:100,000 geological map covering a portion of the North Alborz reverse fault in the eastern Alborz (Fig. 9) shows the dip of this major reverse fault suddenly reversing along the longitudinal line from north-dipping (east of longitude 53°15′E) to south-dipping to the west, mapped by two different geologists who did not reach an agreement about the dip direction of one of the fundamental reverse faults of the Alborz Mountains (Saidi and Vahdati Daneshmand, 2005); paradoxically, the map was approved for publication (GSI, 2005). The North Alborz reverse fault was shown as a: (1) north-dipping reverse fault in Zanchetta et al. (2009; and Stefano Zanchetta, 2 April 2013, personal commun.); (2) a south-dipping fault in Allen et al. (2003), Nazari et al. (2009), and Ballato et al. (2015); and (3) both north- and southdipping along longitude 53°15′E (GSI, 2005), and west of the Harāz River gorge, 53°20′E (Nazari et al., 2005). The Shāhkuh reverse fault in the eastern Alborz (south of the North Alborz fault; Fig. 9) is mapped as a north-dipping reverse fault in GSI (1989), whereas it is shown as a south-dipping reverse fault in Zanchetta et al. (2009). The 80-km-long coseismic surface faulting of the 1990 Mw 7.3 Rudbār earthquake in the western Alborz (Fig. 9; Berberian et al., 1992; Berberian and Walker, 2010) was not mapped on the 1:250,000 geological map of the area published prior to the event (Stöcklin and Eftekhārnezhād, 1969; Annells et al., 1985). The western portion of the coseismic surface fault (from the east of the Sefidrud deep gorge to the west, with a length of 43 km and 100 cm horizontal and 120 cm vertical coseismic surface displacements) was also not shown on the 1:100,000 scale postearthquake geological map of the area prepared 8 yr after the earthquake (Nazari and Salamati, 1998), and only ~15 km (out of 37 km length) of the eastern end of the 1990 coseismic surface fault was marked on the neighboring map (Ghalamghāsh et al., 2002). In an attempt to describe in advance the future largemagnitude earthquakes in the Alborz Mountains, Gorshkov et al. (2009) stated that the pre–1900–2004 earthquakes with magnitude Ms >6.0 in the Alborz nucleated at 135 identified major “morphotectonic nodes” located at intersections of two or several “first, second (both longitudinal), and third (transverse) lin-

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

eaments” crisscrossing each other. They added that since some historical earthquakes, such as the 958 earthquake in northern Tehran (Figs. 5 and 9), do not fall on the assumed “nodes,” then the locations of these events should be revised! Furthermore, the authors claimed that several “seismotectonic nodes” where the largest earthquakes have not yet happened are potential sources for future earthquakes. This antiquated approach and these findings are not followed in this study. As discussed earlier, some historical earthquakes in the Alborz cannot be assigned to specific active faults due to uncertain meizoseismal areas, inadequate numbers of paleoseismic trench studies, and lack of proper knowledge of the deeper structure of the Alborz and its faults (Table 2). Figure 4 shows that the megacity of Tehran is considered at high risk from future earthquakes on the nearby North Tehran, Niāvarān, Moshā, Pārchin, and Pishvā faults, as well as inner-city surface faults (Fig. 2), and possible blind reverse faults underneath the young folds in alluvial deposits (Berberian et al., 1985). These faults have ruptured during historic times, and any of them may rupture in the near future with a large-magnitude earthquake. Only a few paleoseismic trench studies have been carried out on the North Tehran (off the main fault line), Moshā, Pishvā, Firuzkuh, Āstāneh, and Tāléqān faults, as well as the Kahrizak, North Ray, and South Ray escarpments south of the Tehran metropolitan region (Figs. 4, 6, 7, 9; references cited at each fault entry). Due to presence of large Islamic Revolutionary Guard military bases in the east, west, and southeast Tehran, access to parts of the North Tehran, Moshā, Mahmudieh, Pārchin, and some other faults is prohibited since 1979. Some further paleoseismic study results (along the eastern Moshā, Firuzkuh, and Āstāneh faults east of Tehran, and the Tāléqān faults west of Tehran) are also included in this study because some RayTehran medium- to large-magnitude earthquakes were erroneously assigned to these distant faults, and some paleoseismic events on these distant faults were allocated as the source fault ruptures of the historical Ray/Tehran earthquakes (discussed in the following sections). We have summarized the paleoseismic results and recorded historical earthquakes in Tables 4–14 and discuss the results under each fault entry. The reported dates of paleoseismic events are those reported in the original, and no conversion was made in this work. Almost all the estimated paleoearthquake magnitudes derived from the paleoseismic trench studies, or from the surface fault length based on the empirical relations between surface fault rupture and earthquake moment magnitude (Wells and Coppersmith, 1994), were derived without proper knowledge of the thickness of the seismogenic layer, centroid depths of each event, the ratio of dip slip to surface slip, interseismic near-surface fault creep, fault segmentation, and near-surface geology (Manighetti et al., 2007; Dolan et al., 1995, 2007; Dolan and Haravith, 2014), and the results should be read with caution (e.g., Mw 7.86–8.29 in Nazari et al., 2014). Although we do not place much weight on magnitudes derived from Wells and Coppersmith (1994) and other empirical relations based on global events, we report them

Tehran: An earthquake time bomb here as provisional estimates and add comments based on fault lengths. As will be discussed in the following sections, a majority of the estimated magnitudes are over exaggerated, and magnitudes in the range of 7.6–8.29 have been proposed for some paleo-events that cannot be confirmed. Unfortunately, as we discuss later herein, Iranian paleoseismologic data are fragmentary, and the event dates are poorly constrained, with broad error ranges, possibly because of the semiarid nature of the area with high-energy deposits and poor occurrence and preservation of datable material. Therefore, it seems that most of the obtained chronological resolution is too coarse to define a specific prehistoric or historic event. Some paleoseismic trenches were not ideal sites to provide a Holocene chronology of events with datable horizons or comparable isochronous units in the footwall and hanging wall of the faults. Events in adjacent trenches were not satisfactorily correlated. In some cases, the authors overinterpreted the correlation between the observed unconstrained rupture dates and far distant historical earthquakes and forced the paleoseismologic data into matching data from a secondary or tertiary literary source of historical earthquakes, and in a few cases, the rupture event was correlated with a spurious earthquake, which could not be real, or the authors could not decide about the exact number of rupturing events in the studied trenches. In several cases, the alluvial/fluvial layer dates were estimated by guessing mean sedimentary rates over a long time. Therefore, detailed chronologies and characterizations of interval recurrence periods of paleoearthquakes and their estimated magnitudes are still not possible. One of the early paleoseismologic trench studies across the South Eshtehārd fault, 90 km west of Tehran and 60 km southwest of Karaj (Fig. 9), reported an earthquake ~800– 900 yr ago (Bolourchi, 1997). The regional history of the area is well documented for the period, and no earthquake was reported in that region. The author linked this fault to the 1 September 1962 Mw 7.0 Bu’in earthquake, which was not associated with motion on the South Eshtehārd fault (Ambraseys, 1963; Berberian and Yeats, 1999; Berberian, 2014). Without radiometric dating, Bolourchi (1997) speculated three events ca. 9000, 4000, and 900 yr B.P., with 25–30 cm vertical slip for each event. The author further claimed a recurrence interval of 1000 yr, a sliprate of 0.5 mm/yr, and characteristic earthquakes of M 5.5–6.0 for the South Eshtehārd fault with 3 m vertical and 2 m leftlateral displacement. Based on recurrence periods obtained from the fragmentary paleoseismic events with poorly constrained dates, Nazari (2015) forecasted that a probable earthquake of magnitude 6.5–7.2 will occur in Tehran within the next few decades (V II I IX

7.1

7.0

7.1

>6.0 7 .0

+

– – – – 6.4 ± 4.3 k a (M 6.9) –

E5: 7.9–26.7 ka ( M 6.6) E3: 7.9 ka (M 7.1 ) – E 4: 12 k a (M6.5 ) E6: ? – – E7: 16.7–29 ka ( M 6.5) – – – E 5 : 33 k a ( 6. 6 ) – E8: 28.5 ka (M 6. 6) – – – – E6: 50 k a ( M 7.0 ) Note: MMI—modified Mercalli intensity; I—epicentral intensity; Ms—surface-wave magnitude; Mw—moment magnitude. *The Vardāvard trench (35°45′06.2″N, 51°05′05.06.5″E) was located south of the North Tehran fault line and a landslide mass. † Based on alluvial sedimentation rate; no radiometric age was detected across a 6-km-long normal fault east of the Vardāvard trench and south of the North Tehran thrust trace (39°55′N, 51°75′E). § 35°49′49″N, 51°38′10″E.

northwest, and the event was not misallocated at Rhagae (Ray) city (see Appendix B for the contemporary/near contemporary facts about the earthquake). Nazari et al. (2009) also stated that colluvial unit 20, covering the first paleoseismic event 1 (dated younger than 80 CE) along the Tāléqān fault, could correspond to a deposit following one of the strong historical Ray earthquakes of 855, 958, 1177, 1665, or 1830 that occurred in the region (covering a 975 yr interval). As discussed in Appendix B, the 855–856, 958, and 1177 earthquakes destroyed and damaged the city of Ray and its numerous villages far from Tāléqān and Qazvin. Furthermore, as we show in this paper (Figs. 4, 5, 8, and 9; Table 2; Appendix B), the 958 Ruyān and 1830 Lavāsānāt earthquakes, damaging Ray and Tehran, took place along the central segment of the Moshā fault, ~30 km north and northeast of Ray and Tehran, and not at Tāléqān or along the Tāléqān fault. Nazari et al. (2009) also admitted that their youngest event 1 (dated younger than 80 CE in trench T1) was not observed in the nearby trench T2, and that they might have overinterpreted the fractures in trench T1, and they were not certain of a rupture cutting through unit 40 and sealed by unit 20.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

8. THE INNER-CITY RECENT FAULTS AND FOLDS The inner-city surface faults cutting the alluvial deposits are now beneath the Tehran metropolitan area. These including (1) the Mahmudieh, Dāvudieh, ‘Abbāsābād, Takht-e Tāvus, E-W, Nārmak, Shiyān, Telo, and Sorkheh Hessār reverse faults (Berberian et al., 1985); (2) the Bāgh-e Fayz and Television strikeslip faults (Engalenc, 1968; Tchalenko et al., 1974a; Berberian, 1976c); and (3) numerous minor inner-city faults (Berberian et al., 1985), which were mapped in the city prior to expansion of the metropolitan area covering the faults (Fig. 2). Some of the reverse faults, such as the Mahmudieh and Dāvudieh, show pronounced escarpments up to 15 m high (Berberian et al., 1985). Furthermore, several folds have been developed in the Pliocene– Pleistocene alluvial deposits of the city (Rieben, 1955, 1966; Engalenc, 1968), some of which might have concealed blind reverse faults. No paleoseismologic trench study has been conducted along these faults, and no recorded seismic history has been documented along them. Except for a few scattered microearthquakes that were removed during an analysis of instrumental seismicity, the

Tehran: An earthquake time bomb 1999–2000 microearthquake survey based on temporary networks did not detect any seismicity along the inner-city recent faults for the short time period of recording (Ashtari et al., 2005). 9. THE MOSHĀ FAULT SYSTEM The Moshā fault system is one of the major active faults in the central Alborz Mountains 30 km north of Tehran (Figs. 4, 8, and 9). The fault system can be followed for more than 220 km from the Firuzkuh area in the east to the south Tāléqān area in the west. The trace of the Moshā fault system is sinuous, consisting of three segments: (1) an approximately E-W western Maydānak segment (5 km south of the Tāléqān fault), with a length of ~70 km, (2) the central Moshā WNW-ESE–striking segment, 130 km long; and (3) the ENE-WSW eastern segment of more than 26 km length in the Firuzkuh area. The closest distance of the Moshā fault to Tehran is 16 km to northern Tehran and 28 km to downtown (Fig. 4). The fault is a north-dipping, left-lateral, oblique reverse fault with dips varying from 35° to 70° that has juxtaposed the Upper Neoproterozoic– Lower Cambrian to Mesozoic sediments of the southern Paleozoic–Mesozoic zone of the High-Alborz in the hanging wall with intensely folded rocks of the border folds of the southern Tertiary zone (the Eocene Karaj Formation) in the footwall (Assereto, 1966; Tchalenko et al., 1974a; Berberian, 1981; Berberian et al., 1983, 1985, 1996; Guest et al., 2006a, 2006b; Ehteshami Moinabadi and Yassaghi, 2007; Yassaghi and Madanipour, 2008). At Moshā village in the east (Fig. 4), the Lower Cambrian Lālun Formation is thrust over Quaternary alluvial deposits in the south. A 10 m highly sheared zone at the west side of Moshā village with dips ranging from 36° to 80° was reported by Berberian et al. (1985). The cumulative vertical displacements amount to at least 4 km along the Moshā fault (Assereto, 1966). Similar to the Kuhbanān fault in southeast Iran (Berberian et al., 1979; Berberian, 2014) and the North Tehran fault system (Berberian et al., 1985; Landgraf et al., 2009; Solaymani Azad et al., 2011), the kinematics of the eastern segment of the Moshā fault have changed during the post-Neogene from reverse to predominantly left-lateral strike-slip within the transpressional system of the Alborz (Jackson et al., 2002; Allen et al., 2003; Guest et al., 2006a; Ritz et al., 2006; Landgraf et al., 2009; Solaymani Azad et al., 2011; Ballato et al., 2013, 2015). Based on piercing points on the geological map of the central segment of the Moshā fault (GSI, 1987), Allen et al. (2003) measured a cumulative left-lateral offset of ~30–35 km in ~5 m.y., which resulted in a slip rate of 6–7 mm/yr. In the area south of Damāvand volcano (Figs. 4 and 9), the Moshā fault shows a subvertical fault zone, with entrained, fault-bounded lozenges of various lithologies, including gypsum. Both steeply dipping and subhorizontal slickensides are present on fault surfaces. West of Tār Lake (Fig. 12), the Moshā fault with a prominent linear topographic expression has left-laterally disrupted and beheaded the drainage. The fault yielded an estimated average left-lateral slip rate of 2.2 ± 0.5 mm/yr in the east (Ritz et

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

115

al., 2003, 2006; Solaymani Azad et al., 2003, 2011; Solaymani Azad, 2009; Bachmanov et al., 2004), whereas to the west, reverse motion predominates. A digital elevation model indicates the ratio between the horizontal (H = 100 m) and the vertical (V = 20 m) components is 5:1 (Ritz et al., 2006). This result differs from the previous H/V ratios of ~2:1 and ~1:1 (Trifonov et al., 1996; Bachmanov et al., 2004), and ≥3:1 in the Irā valley (Fig. 4; Solaymani Azad et al., 2011). Based on 2–3 km of left-lateral drainage displacement along the Moshā fault, Ritz et al. (2006) suggested that inception of strike-slip faulting in the Alborz took place in the middle Pleistocene. Hollingsworth et al. (2008), by referring to longer leftlateral displacements of geological key beds along the Āstāneh fault, suggested that commencement of strike-slip motion took place ca. 10 Ma. Later, Ghassemi et al. (2014) proposed that strike-slip motion along the central Moshā fault (Fig. 4) initiated between 3.2–4.7 Ma. This is consistent with the westward motion of the South Caspian Basin postdating the thrusting phase of the North Tehran fault system ca. 7–6 Ma (Ballato et al., 2013) or 5.5 Ma (Allen et al., 2002). 9.1. Seismicity (Moshā Fault) The Moshā fault was associated with recorded medium- to large-magnitude historical earthquakes of 958 Mw ≥7.0 Ruyān, 1665 Mw 6.5 Damāvand, and 1830 Mw 7.0 Lavāsānāt (Table 7), the latter damaging the city of Tehran with some casualties (Table 2; Figs. 5, 8, and 9; Appendix B); as well as small- to medium-magnitude earthquakes in 1930 (Mw 5.1, two events), 1947 (VI+), 1955 (Mw 3.9), 1970 (Mw 4.3 and 4.5), 1988 (Mw 4.8), 2004 (Mw 4.1 and 3.9), and 2006 (Mw 3.8; Tables 3 and 8; Figs. 8 and 9; Appendix C; Berberian, 1981, 2005, 2014; Berberian et al., 1983, 1985, 1996; Berberian and Yeats, 1999, 2001; Donner et al., 2014). The 1999–2000 microearthquake survey showed activity related to the eastern portion of the fault, with left-lateral mechanisms along the Moshā fault and a clear northward deepening of seismicity associated with the fault (Ashtari et al., 2005), consistent with geological observations on the dip of the fault. The five month microearthquake survey along the fault in 2006 showed that most earthquakes were located along the fault at a depth range of 0–20 km with a maximum concentration between 10 and 15 km. Cross sections revealed that the central section of the fault dips 70o to the north with focal mechanisms showing left-lateral strike-slip motion (Tatar et al., 2012). The 2004– 2010 microearthquake survey showed some earthquakes along the central and eastern segments of the Moshā fault (Vasheghani Farahani et al., 2014). Inversion of broadband and short-period waveforms of three small-magnitude earthquakes along the Moshā fault zone showed E-W left-lateral mechanisms and NW-SE reverse mechanisms with a slight left-lateral component, with the regional seismicity distributed to a depth of 20 km (Donner et al., 2014). The events were: 21 February 2004 (12:02 UTC; Mw 3.9, centroid depth 14–16 km)

116

Berberian and Yeats TABLE 7. SUMMARY OF PALEO- AND HISTORICAL EARTHQUAKES ALONG THE MOSHĀ FAULT (FIGS. 4, 8, 9, and 12; SEE ALSO APPENDIX B) Paleoearthquake age (ka)

Irā (Ghassemi et al., 2014)

0.083 0.183 1.055 1.105 – (6.9–5.2)* – –

Ab-e ‘Ali (Solaymani Azad et al., 2011)

– – – – 1.21 ± 0.15 – – –

Documented historical earthquake (yr.mo.d)

Location/ this study

2 0 0 6. 1 2 .2 0 2004.09.24 2004.02.21 1988.08.23 1988.01.14

Moshā Moshā Moshā Moshā Moshā North Tehran? Rud-e Hen Rudbār Qasrān Moshā Lavāsānāt Āh AFS Āh AFS Āh Lavāsānāt AFS Lavāsānāt Damāvand Damāvand Damāvand Ruyān – – – – – – –

~I (MMI)

~Ms

mb

~Mw

Tār (Ritz et al., 2003)

– – – – – –– 7 . 51 4– 7.569 –

1974.01.10 19 70 . 1 0. 0 3 1955.11.24 19 47 . 09 . 0 5 1 93 0. 10 .0 7 1 93 0. 10 .0 6 1930.10.02 1 83 0. 04 .0 6 18 30 . 0 3. 2 7 18 15 . 0 6. 0 0 18 11 . 0 6. 2 0 1 66 5. 0 6. 15– 07 . 1 3 958.02.23 – – – – – – – Po st 24 –14 k a?

+

V VI VI VI V V + VI + >VII IX F F + VIII X – – – – – – –

4.0 4.6

3.8 4.1 3.9 4.3 4.8

4.3 4.1 4.0

4.5 4.3 3.9

5.0

5.1

– – – – – – –

5.2 >6.0 7.1

5.1 >6.0 7.0

6.5 >7.1 – – – – –

6.9 >7.0 – – – – – – –



Lāsem – – – Seismite ? – – – – – – – 120.5–89.9 – – – – – – – ? – – – – – – 150.5–120.5 – – – – – – – 150.5 ± 23 – – – – – – – Note: MMI—modified Mercalli intensity; ~ indicates approximate value; ? indicates unknown with questionable source. AFS—aftershock; I—epicentral intensity; F—shock felt; mb—body-wave magnitude; Ms—surface-wave magnitude; Mw—moment magnitude. *Unpublished data, Angela Landgraf (22 April 2014, personal commun.).

with NW-SE reverse mechanism and minor left-lateral slip component; 24 September 2004 (01:42 UTC, Mw 4.1, centroid depth 4– 6 km), with E-W left-lateral mechanism steeply dipping north); and 20 December 2006 (04:39 UTC, Mw 3.8, centroid depth 14 km), with E-W left-lateral mechanism steeply dipping south. The first two events were located near the Moshā village, and the third event was recorded just to the east of the junction of the North Tehran fault system with the Moshā fault (Figs. 8 and 12).

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

– – – – – –

9.2. Paleoseismologic Trench Studies (Moshā Fault) Three trench studies have been carried out across the Moshā fault (Figs. 4 and 12; Table 7); these are described next. The Tār Lake Valley Trench (Moshā Fault) A trench study (T3 at Tār Lake valley; Fig. 12; Table 7) showed a fine-grained unit trapped behind the fault possibly at

Tehran: An earthquake time bomb

117

TABLE 8. SUMMARY OF PALEOEARTHQUAKES OBTAINED FROM THE IRĀ TRENCH (35°48′33″N, 51°40′43″E) ACROSS THE CENTRAL MOSHĀ FAULT (MODIFIED FROM GHASSEMI ET AL., 2014; ANGELA LANDGRAF, 22 APRIL 2014, PERSONAL COMMUN.; SEE ALSO FIGS. 4, 9, AND 12; APPENDICES B AND C) Event type

Age (ka)

Documented historical earthquake





1930.10.02 Ms 5.2 Āh





1830.03.27 Ms ~7.1 Lavāsānāt





1665.06.15–07.13 Ms ~6.5 Damāvand





958.02.23 Ms ~7.1 Ruyān



0.083 ±?



Faulting 24

0.183 ±?



Debris flow (units 29 & 30)

1. 055 (? )



(yr.mo.d)

De bris flow (unit 27)

1.105 ± 30



(Paleoearthq uakes 1 to 5)*

(6.9–5.2)*



?



Second major rock avalanche (units 8, 9, 15, 16)

120.5 ± 9.8–89.9 ± 15.8



Strike-slip faulting (lower part of unit 12 is offset; soils in the upper part dated 120.5 ka OSL-XII) associated with the first major rock avalanche (units 1, 7, 10, 11)

150.5 ± 23–120.5 ± 9.8



150.5 ± 23



?



Reverse faulting & scarp development (F13 faulting, sealed by unit 29 w/conflicting OSL-VI age)

Strike-slip faulting (F.17, 19) & Ira rock avalanche (unit 33, development of sag pond) Last reverse fault i ng (F.23 )

Note: (?) See discussion in the text. OSL—optically stimulated luminescence. *Unpublished data with five events from 6938 to 5229 ka; Angela Landgraf (22 April 2014, personal commun.).

the beginning of the formation of the fault scarp at a depth of 2–3 m. Radiocarbon and OSL dating resulted in an age of 7514–7569 calibrated yr B.P. (Ritz et al., 2003). Investigation at the Lake Tār valley along the Moshā fault showed 15 ± 1 m of left-lateral slip; considering this displacement within 7514–7569 yr, this resulted in an estimated minimum left-lateral slip rate of 2 ± 0.1 mm/yr (15 ± 1 m/7514–7569 yr). Using empirical relations, the authors claimed a mean recurrence interval between 160 and 620 yr for earthquakes with Mw 6.5 and 7.1 along the Moshā fault (Ritz et al., 2003). The Āb-e ‘Ali Ski Resort Trench (Moshā Fault) OSL dating of the most recent colluvial unit (unit 4 as the last event-horizon) cut by the Moshā fault at the Āb-e ‘Ali Ski Resort, located ~24 km west of the Tār Lake trench (35°46′14.8″N, 51°59′09.6″E; +2462 m; sample T7.OSL-2 Ab Ali; Figs. 4 and 12; Table 7), yielded an age of 1.21 ± 0.15 ka as the maximum age of the last paleoseismic event horizon, with 45 cm vertical slip measured at the trench wall and 1.35 m of horizontal displacement (Solaymani Azad et al., 2011; Solaymani Azad, 12 January 2014 and 10 November 2015, personal commun.). The authors argued that the last paleoseismic event could correspond either to the 1665 or the 1830 earthquake, with the 1830 event

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

preferred (Solaymani Azad et al., 2011; Solaymani Azad 12 January 2014, personal commun.). The OSL age of the “event” is not constrained. The Irā Road-Cut Trench (Moshā Fault) Ghassemi et al. (2014) studied a road-cut trench at Irā (35°48′33″N, 51°40′43″E) across the Moshā fault, located east of the junction with the North Tehran fault system and ~15 km west of the Āb-e ‘Ali trench (Fig. 4; Tables 7 and 8). The trench exposed a view onto an exhumed fault zone showing evidence of prolonged Quaternary faulting. Stratigraphic units span ages between 150,000 and 1100 yr B.P., with a hiatus of ~74,000 yr in the late Pleistocene. Four OSL 14C ages bracketed a sequence of four to five paleoearthquakes defined by their colluvial wedges (Ghassemi et al., 2014; Angela Landgraf, 22 April 2014, personal commun.). The reverse faulting at F13 (Table 8) is sealed by unit 29, indicating the event happened earlier. However, the OSL age for this unit conflicted with the stratigraphic results. The authors found strike-slip faulting roughly between 150 and 120 ka (the lower part of unit 12 has been offset, and soil in the upper part of this unit was dated to ca. 120 ka; Angela Landgraf, 22 April 2014, personal commun.).

118

Berberian and Yeats

9.3. Summary of the Trench Studies across the Moshā Fault Table 7 summarizes the findings in the three trenches across the Moshā fault. Documented historical earthquakes have been added to the table. The distance between the Irā and Āb-e ‘Ali trenches is 16 km, and the distance between the latter and the Tār Lake trench is 25 km (Figs. 4 and 12). Despite their proximity, no single event is recognized in all three trenches. No evidence for the 1930 Mw 5.1 and 1830 Mw ~7.0 earthquakes or independent dating was found in the Irā trench (Tables 7 and 8). However, Ghassemi et al. (2014) proposed possible fracturing associated with a landslide that might be correlated to the 1830 earthquake. Furthermore, no evidence for surface rupturing associated with the 958 Mw ~≥7.0 earthquake was observed in the trench, but a landslide deposit with roughly this date was detected (Angela Landgraf, 22 April 2014, personal commun.). 10. THE KAHRIZAK, SOUTH RAY, AND NORTH RAY ESCARPMENTS Based on groundwater table measurements (MWP, 1970; Knill and Jones, 1968; Tchalenko et al., 1974a; Berberian et al., 1985), groundwater cascades of unidentified origin were reported in a single aquifer along: (1) the North Tehran fault system; (2) in the central Tehran area (along the western continuation of the Sorkheh Hessār fault and eastern continuation of the Shahrdād fault); and (3) in the Ray area (along the North Ray escarpment; Fig. 4). Based on the surface expression of the topographic scarps, incised streams (especially at Kahrizak, though continuing farther north), and the groundwater cascade mentioned earlier, the North Ray, South Ray, and the Kahrizak escarpments facing south (Fig. 13) were speculated to be reverse faults displacing recent alluvial deposits as well as the groundwater level (for details, see Berberian et al., 1985; Feghhi, 1999). Nazari et al. (2005) drew two ellipses as the meizoseismal areas of the 743 and 1895 CE earthquakes along the North Ray escarpment/fault. Later, Nazari et al. (2010) reported that the three topographic escarpments (the North Ray, South Ray, and Kahrizak, spread across a N-S distance of 13 km at elevations of 1088–1097 m, 1055–1070 m, and 1000–1020 m, respectively; Fig. 13) followed contour lines matching with ancient shorelines southwest of the Pārchin fault at an elevation of 1055 m (Fig. 4). They suggested that the three escarpments were most likely to be terrace risers and not fault scarps. A “problematic shoreline” was described by Rieben (1955, figure 1, p. 619; 1966, p. 38–39) in the area southeast of Ray to the Jājrud River farther southeast, southwest of the Pārchin fault (Fig. 4; see also figure 1 in Rieben, 1955; and aerial photograph on p. 103, plate 4.29 in Berberian et al., 1985). Accordingly, light-colored or reddish horizontal deposits occur in long horizontal strips, crossing old or sub-Holocene alluvial fans. Three or four of these closely spaced strips occur between elevations +1050 m and +1040 m. If a lake existed at an altitude of

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

~1040 m, Rieben (1966) added, similar remnants of shorelines should exist elsewhere around the Kavir (desert) basin of north Central Iran. On figure 1 in Rieben (1955), the “problematic shoreline” starts from elevation +1036 m (+3400 ft) in the northwest and then rises up to elevation +1100 m (+3612 ft) in the southeast near the Jājrud River. The elevation of the upper band near the Jājrud River is at +1076 m (+3532 ft), lowering to +1072 m (+3519 ft) to the northwest, whereas the lower band is at an elevation of +1041 m (+3418 ft) near the Jājrud River in the southeast, lowering to +1036 m (+3399 ft) to the northwest, based on a Google Earth map. Rieben (1966) speculated that the bands might have been formed by silting up of old canals built to carry water from the Jājrud River to the ancient city of Rhagae (Ray) to the northwest (?; Fig. 4). The local farmers near Hesāramir attributed the vestiges to their forefathers (Rieben, 1966). If the latter case is correct, it may explain the lowering of the band from the Jājrud River in the southeast (+3612 ft/+1100 m) to the northwest approaching the city of Ray (+3400 ft/+1036 m). However, this interpretation has not been confirmed. In any event, the elevation between +1050 m and +1040 m in the southeast only corresponds to the elevation of the eastern tip of the South Ray escarpment. The main portion of the South Ray escarpment stands at an elevation of +1070 m (Fig. 13). Furthermore, the North Ray escarpment in the north is at elevation of about +1086 m (in the east) and +1097 m (in the west), which is above the elevation of the “problematic shoreline” of Rieben (1955, 1966). The elevation of the Kahrizak escarpment to the south ranges from +1000 m (in the east) to +1020 m (in the middle) and +1040 m (in the west; Fig. 13). Therefore, the North Ray and the Kahrizak escarpments with identical deposits are not at the level of the “problematic shoreline” altitude of Rieben (1955, 1966) in the east. These two escarpments are composed of silty clay deposits of the “Kahrizak Formation” (Rieben, 1955, 1966) or the “q2cc” (“croute calcaire” or “Ray et Kahrizak-B” of Engalenc, 1968), and the “Bs” (the Kahrizak clayey silt of Berberian et al., 1985). Reported thicknesses up to 200 and 300 ft (60.9–91.4 m) were viewed in deeply dissected ravines, and a thickness of 250 ft (76.2 m) has been measured in a well log (Rieben, 1955). Apparently, three boreholes were drilled in the formation for 200 m in depth, but the base of the formation was not reached (JICA, 2000). However, the logs of the boreholes are not available to confirm the reported thickness of the formation. 10.1. The Kahrizak Escarpment The Kahrizak escarpment forms an impressive E-W–trending, 35-km-long south-facing scarp up to 15 m high in the Holocene (?) alluvial deposits ~22 km south of central Tehran. The eastern part of the scarp follows the +1000 m contour line; the elevation increases to +1020 m in the middle and +1040 m in the west (Fig. 13). As a direct consequence of the relative difference in altitude of the two sides of the escarpment, incipient drainage with

Figure 13. Topographic map of the south Tehran region (see Fig. 1 for location) showing the North Ray, South Ray, and Kahrizak escarpments (nearly E-W broken lines) with locations of paleoseismologic trenches (stippled rectangles; see text for trench results and references). Contours are in meters above mean sea level. Symbols and bottom-left inset are as in Figure 2. Triangles—archaeological sites (see Appendix B). The location of the ancient city of Rhagae (modern Ray, south of Tehran) and the inner and outer areas of Ray city are added.

Tehran: An earthquake time bomb

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

119

120

Berberian and Yeats

several deeply incised small creeks (though continuing further north into the plain) characterizes the northern block, whereas only a few large rivers exist on the southern lower block. De Martini et al. (1998) and De Martini (2007) reported locally small creeks showing evidence for right-lateral slip up to 50 m, and some qanāt lines suggest abandonment following an event of lateral offset. The authors attributed the abrupt double bend of ~90° of the Jājrud River to the east to be consistent with right-lateral slip along the eastern end of the Kahrizak fault. The 1999–2000 microearthquake survey based on temporary networks did not detect any seismicity along the Kahrizak, South Ray, or North Ray escarpments for the short time period of recording (Ashtari et al., 2005). Two controversial paleoseismic trench studies were conducted in 1998 (De Martini et al., 1998; De Martini, 2007) and 2010 (Nazari et al., 2010) across the Kahrizak escarpment, the results of which contradict each other (Figs. 4 and 13). Kahrizak Paleoseismologic Trench Study of 1998 Two nearby paleoseismic trenches across the Kahrizak escarpment (K1 and K2; 35.50°N, 51.47°E; Fig. 13), with 8 m vertical throw, exposed an intensely deformed, 30-m-wide fault zone composed of north-dipping, high- and low-angle reverse faults striking N70°–80°E, dipping 70°–80°N, with horizontal slickenside striations showing a 10°–15°W pitch (165°– 170° rake) in trench K2 (see figure 6 in De Martini et al., 1998; De Martini, 2007). The authors added that the main fault zone corresponds to the location of calcrete deposits, suggesting that faulting, facilitating fluid circulation, might have controlled the formation of calcrete in the silty clay. Horizontal movement was reported as predominantly right lateral, with the northern side of the fault up. No organic-rich deposit was found in the two trenches. Radiocarbon dating of the total organic component contained in unit 3 yielded a mean residence time age of 10,170 ± 150 yr B.P. in trench K2. This unit is underthrust beneath the older unit 8a. A 230Th-234U date of the younger faulted unit 5c (resting on unit 8a) obtained the best constraint in the range of 7000 yr B.P. to present. The oldest unit 10 in trench K1 is younger than 11,500 yr B.P. Furthermore, unit 8a, with faulting and block rotation, is covered by the base of unit 4b-4a in trench K1 (De Martini et al., 1998; De Martini, 2007). Some evidence for individual Holocene earthquakes was also reported by De Martini et al. (1998) and De Martini (2007); however, the authors were not able to reconstruct the seismic history of the fault or to evaluate the extent of deformation produced by each event. A 10 m right-lateral offset of a qanāt line led De Martini et al. (1998) and De Martini (2007) to hypothesize that the most recent event may have occurred in historical times (more recent than 5000 yr B.P.). The authors estimated an elapsed time between 5000 and 800 yr and a maximum slip per event of ~10 m. A minimum Holocene vertical slip rate of ~1 mm/yr, a horizontal slip rate of ~3.5 mm/yr, a maximum of ~3000 yr average recurrence interval, and an

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

expected Mw of 7.0–7.4 were estimated (De Martini et al., 1998; De Martini, 2007). The study was not conclusive enough to reconstruct the sedimentary history of the Kahrizak Formation and the seismic history of the Kahrizak-Ray-Tehran area and did not resolve the seismic issues of the region. Offset qanāts could have been used to determine a slip rate, but De Martini et al. (1998) and De Martini (2007) did not present any archaeological or radiometric evidence for the age of the qanāts. The results of the analysis of the Kahrizak paleoseismic trench study and the nature of the Kahrizak escarpment remain uncertain. Kahrizak Paleoseismologic Trench Study of 2010 Three adjacent trenches were dug across the Kahrizak escarpment, to the west of the aforementioned trench, with an 11-m-high scarp at 35°30′N, 51°56′E (Fig. 13). About 8 m of the horizontal beds of the Kahrizak silty clay rest on the folded and faulted (E-W reverse fault dipping 60° to the north) late Miocene–Pliocene Upper Red Formation molasse. No evidence of faulting was found in the late Quaternary Kahrizak Formation (Nazari et al., 2010). The authors concluded that the scarp was related to a shoreline (discussed earlier herein) of a lake in the northern Central Iranian desert during the warming period of the Holocene, covering an area of ~70,000 km2, and they stated that is not likely to correspond to an active fault. This conclusion is in sharp contrast with those of De Martini et al. (1998) and De Martini (2007), who reported an intensely deformed, 30-m-wide, north-dipping, high- and lowangle fault zone along the Kahrizak escarpment. 10.2. The South Ray Escarpment The South Ray escarpment is a nearly E-W–trending escarpment of 18.5 km length with ~3 m height located ~13 km south of central Tehran (Fig. 13). The eastern part of the scarp follows a contour line of +1050 m; the elevation increases to +1060 m in the middle and +1065 m in the west (Fig. 13). Apparently, a natural cross section at a river bank at 35°33.537′N 51°19.019′E showed no faulting (Nazari et al., 2010). The authors concluded that as with the Kahrizak escarpment in the south, the South Ray scarp corresponds to another shoreline (discussed earlier) of a lake in Central Iranian desert during the warming period of the Holocene covering an area of ~70,000 km2, and added that is not likely to correspond to an active fault. 10.3. The North Ray Escarpment The North Ray Escarpment is a nearly E-W–trending escarpment ~16.5 km long and 2–4 m high located ~9 km south of central Tehran. The eastern part of the escarpment has an altitude of +1080 m; it rises to +1095 m in the middle and +1097 m in the west (Fig. 13). A paleoseismic trench dug at 35°36′21.2″N, 51°23′09.5″E (Fig. 13) showed horizontal alluvial layers with no faulting in the trench (Nazari et al., 2010). The authors concluded that, as with the Kahrizak and South Ray scarps, the North Ray

Tehran: An earthquake time bomb scarp also corresponds to another shoreline of a lake in Central Iranian desert covering an area of ~70,000 km2, and it is not likely to be related to an active fault.

121

Caspian Gates (Sardarreh Defile), at the southeastern tip of the Pārchin fault and on the hanging wall of the Garmsār thrust (Figs. 8 and 9; Table 9). The 2006 microearthquake survey showed some seismic activity along the Pārchin faults (Tatar et al., 2015).

11. THE PĀRCHIN FAULT 12. THE PISHVĀ FAULT The Pārchin fault (Berberian et al., 1985; Berberian and Yeats, 1999, 2001; the Ayvān-e Kay fault in Berberian, 1981) is an ~70-km-long fault with a NW-SE strike that extends from the city of Ray in the northwest to Ayvān-e Kay near Sardarreh Defile (Caspian Gates) in the southeast (Figs. 4, 8, 9, 17, 18; Caspian Gates discussed later Fig. 18 caption). The fault clearly cuts Quaternary alluvial deposits of the plain in the area east of the Jājrud Valley mouth (aerial photographs 4827 and 4828, Worldwide Aerial Surveys, Inc., project 158, scale 1:55,000; see photograph 4.29, p. 103 in Berberian et al., 1985), not far from the controversial Pārchin military nuclear complex (35.52°N, 51.77°E; Fig. 4). Unfortunately, parts of the Pārchin fault near to the military base have not been accessible for field study. Djamour et al. (2012) stated that the 743 CE earthquake (discussed in Appendix B) took place on the Pārchin fault or the Garmsār fault. This dubious event is deleted from our corrected catalogue (Table 2; see Appendix B for discussion). 11.1. Seismicity (Pārchin Fault) The ca. 280 BCE Rhagae (modern Ray) earthquake (discussed in detail in Appendix B; Table 2) might have possibly been generated along the Pārchin fault (Figs. 4, 6, 9). A paleoseismic trench study is needed to evaluate this possibility. Microseismicity recorded by a temporary seismological network in 1999 and 2000 showed seismic activity along the Pārchin fault (Ashtari et al., 2005). The epicenters of the 12 May 1993 mb 4.3, 17 October 2009 Ml 4.0, and 20 February 2011 Ml 4.0 earthquakes were located near the western tip of the fault; however, the location error of the small-magnitude earthquakes is not known (Tables 3 and 9; Fig. 8). The epicenters of the 22 August 1988 Mw 5.3 and the 23 August 1988 Mw 5.1 earthquakes were located near the

The Pishvā reverse fault (Berberian et al., 1985; Berberian and Yeats, 2001) is an ~55-km-long, NW-SE (N130°E)–trending, northeast-dipping reverse fault with a lateral component, located southeast of Ray and Tehran (Figs. 4 and 8). To the southeast, the Garmsār fault bifurcates from the Pishvā fault toward the east. It clearly cuts the Quaternary alluvial deposits, and the Miocene Qom Formation, Neogene molasse deposits, and Pliocene–Pleistocene Hezārdarreh Formation in the northeast (+1113 m amsl) are thrust over Quaternary alluvial deposits of the Varāmin plain (+850 m) in the southwest (plates 4.28 and 4.28a, p. 101 and 102 in Berberian et al., 1985). The fault has a major reverse component with possible lateral motion; the asymmetric hanging-wall anticlinal axis in the northeast is very close to the fault trace with steep to overturned limbs. Majidi Niri et al. (2010, 2011) reported that the fault is located between the Alborz and Central Iran structural zone, whereas in fact it is in the Central Iran zone, sensu stricto. The city of Varāmin (35°19′N, 51°38′E; with population of 218,991 in 2011) and the town of Pishvā (35°18′N, 51°43′E; with population of 47,253 in 2011; SCI, 2013), 6.5 km apart, are located along the Pishvā fault (Fig. 4). The town of Pishvā has expanded in recent years, with buildings covering the fault near the nose of the asymmetric faulted anticline. 12.1. Seismicity (Pishvā Fault) No historical seismic data are recorded along the Pishvāfault. Epicenters of 13 August (three events, mb 3.6, 4.1, and 3.4) and 14 September 2015 took place along the Pishvā fault (Figs. 4, 8, 9; Table 10). However, Ritz et al. (2012, p. 14/B6305) without presenting any supporting data suggested that Pishvā fault might have been the source of the 855 Ray earthquake.

TABLE 9. RECORDED SEISMICITY ALONG THE PĀRCHIN FAULT (SEE FIGS. 8 AND 9; TABLE 3) Date (yr.mo.d) Area Source I Ml mb Ms Mw 2011.02.20 4.0 ISC 2009.10.17 4.0 4.3 TEH 1993.05.12 4.3 4.5 NEIS 1988.08.23* 5.0 4.8 5.3 ISC 1988.08.22* 5.0 5.0 5.1 ISC + Rhagae (Ray) >VIII Ca. 280 BCE – – – >7.0 See text & Appendix B (312–280 BCE) Note: I—epicentral intensity; MI—local magnitude; mb—body-wave magnitude; Ms—surfacewave magnitude; Mw—moment magnitude. Sources: ISC—International Seismological Summary, Berkshire, UK (1913–1963; now ISC); TEH—Institute of Geophysics, University of Tehran; NEIS— National Earthquake Information Service, U.S. Geological Survey, Denver, Colorado (1964–1985). *Epicenters located near the junction of the Garmsār and Pārchin faults (Fig. 9).

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

122

Berberian and Yeats

TABLE 10. SUMMARY OF THE PISHVĀ TRENCH STUDY (MAJIDI NIRI ET AL., 2011; TĀHEREH MAJIDI NIRI, 5 APRIL 2014, PERSONAL COMMUN.)* AND RECORDED EARTHQUAKES ALONG THE PISHVĀ FAULT (FIGS. 4 AND 8) Trench T1

Event

Disp. (m)

Mw

Date (ka)

Trench T2

Corresponding earthquake/area

Event

Disp. (m)

Mw

Date (ka)

Historical earthquakes (this study) (yr.mo.d) Date Ml Mw

2015.08.14 3.4 2015.08.13 3.6 2015.08.13 4.1 2015.08.13 3.4 1 0.87–1.113 1384/Ray – – – – – – 2 0.265 6.40 7.4–11.67 – – – – – – – – – – – – 1 12.1–22.4 – – 3 0.86 6.73 17.2–20.11 – – – – – – – 4 2.90 7.08 23.5–25.4 – 2 0.13 6.20 22.4–25.4 – – 5 – – 25.4–32.6 – 3 0.36 6.49 25.4–32.6 – – Note: Disp—displacement. 1384/Ray—alleged earthquake not supported by any contemporary or near-contemporary sources (see the analysis in Appendix B). Ml—local magnitude; Mw—moment magnitude. *No radiometric dating was implemented. Dates are based on estimated alluvial sedimentation rate from northern Tehran.

12.2. Pishvā Paleoseismologic Trench Study Two neighboring trenches at a construction site in a southeastern suburb of Pishvā across the Pishvā fault at 35°18′N, 51°44′E were studied by Majidi Niri et al. (2010, 2011); see also Figure 4. By comparing alluvial deposits along the Pishvā fault with those of the North Tehran thrust at Chitgar (Kaveh Firouz et al., 2012) at the foot of the Alborz Mountains (a long stretch of 70 km to the northwest!), and estimating an average rate of alluvial sedimentation of 0.16 mm/yr (from the North Tehran fault region at the foot of the Alborz Mountains), and without any radiometric dating, Majidi Niri et al. (2010, 2011) reported the youngest and oldest events at 0.87 ka and 29 ka, respectively (Table 10). Three to five seismic events with magnitudes ranging from 5.90 to 7.08, with displacements ranging from 0.13 to 2.9 m; estimated left-lateral slip rates of 0.12–0.14 mm/yr, and N-S horizontal shortening of 0.10– 0.12 mm/yr; and recurrence time of 3265 yr were estimated (Majidi Niri et al., 2010). Majidi Niri et al. (2010, 2011) and Nazari (2015) assigned the 1384 CE dubious earthquake, allegedly destroying the city of Ray, to the reactivation of the Pishvā fault in their trench study (Table 10). The 1384 Ray earthquake is of dubious authenticity (discussed in Appendix B) and is not confirmed by any contemporary or near-contemporary sources. It is, therefore, removed from our corrected catalogue (Table 2). The history of this period is well documented by contemporary Iranian scholars, including Yazdi (1419–1425), who does not mention destruction of Ray by an earthquake ca. 1384 (Appendix B). 12.3. Archaeoseismic Data The only historical monument in the town of Pishvā is the mausoleum of Ja’far (Sāménāt) with an inscribed date of 1549

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

3.9 4.3 – – – – – –

CE and a repair date of 1812 (Meshkati, 1970). The intact structure shows that there was not any destructive earthquake along the Pishvā fault at least since 1549. The intact 12-m-high ‘Alā’ ad-Din Tomb Tower (built in 1289 CE), mausoleum of Yahyā (1307–1308), Varāmin congregational mosque (1322), and the mausoleum of Shāhzādeh Hossein (1330) in the city of Varāmin (Wilber, 1955; Meshkati, 1970; Michell, 1978; Blair and Bloom, 1994) are proof of the lack of any destructive earthquake along the Pishvā fault at least since 1289; however, the structures were repaired previously (Figs. 4; Table 2; Appendix B). Therefore, the statement of Majidi Niri et al. (2010) and Nazari (2015) about the Pishvā fault being the source of the dubious 1384 earthquake in their trench study (or any other earthquake between 0.87 and 1.113 ka; Table 8) cannot be warranted. We know that since construction of the Varāmin congregational mosque in 1322, the western part of the structure was damaged by a flooding event ca. 1634 (Monshi, 1634). The flooding scars are still visible on the Varāmin congregational mosques and the ‘Alā’ ad-Din Tomb Tower, indicating previous flooding phases in the Varāmin Plain, west of the Pishvā fault. 13. THE GARMSĀR FAULT The Garmsār reverse fault (Berberian, 1981; Berberian et al., 1985; Berberian and Yeats, 1999, 2001) is located ~65 km southeast of Tehran, north of the town of Garmsār (35°13′N, 52°20′; with a population of 40,985 in 2011; SCI, 2013) in northern Central Iran. The fault bifurcated from the southern section of the Pishvā fault (Figs. 8 and 9). Neogene molasse deposits from the north are thrust over alluvial deposits along this E-W– trending, 70-km-long fault.

Tehran: An earthquake time bomb 13.1. Seismicity (Garmsār Fault) No historical seismic data are recorded along the Garmsār fault. However, Tatar et al. (2015) considered that the third century BCE (the 312–280 BCE Rhagae earthquake in this study) took place along the Garmsār fault. The 11 May 1945 M 4.7 Bonkuh/Garmsār, 25 October 1982 Ms 5.4 Garmsār, 22 August 1988 Mw 5.3, and 23 August 1988 Mw 5.2 earthquakes, which may have taken place along the Garmsār or adjacent faults (Fig. 9), were strongly felt at Tehran. The best centroid-moment tensor solution of the 22 August 1988 Mw 5.3 earthquake (Figs. 8 and 9; discussed later in Fig. 20 caption) corresponds to NW-SE right-lateral and NE-SW left-lateral motion along nodal planes (Dziewonski et al., 1989) and does not match the strike of the Garmsār fault. The NW-SE plane corresponds to the trend of the Pishvā and Pārchin faults, and the NEIC (National Earthquake Information Center, Colorado, USA), ISC (International Seismological Centre, Berkshire, UK), and EBH (Engdahl et al., 1998: EBH, in ISC) bulletin epicenters (location errors unknown) are located near the Caspian Gates (Sardarreh Defile), at the southeastern tip of the Pārchin fault and on the hanging wall of the Garmsār thrust (Fig. 8, and discussed in Fig. 20 caption; Table 11). The best centroid-moment tensor solution of the 23 August 1988 Mw 5.1 earthquake shows NW-SE–trending normal faulting (Dziewonski et al., 1989). The normal faulting mechanism is clearly unusual for this region (Fig. 9 and discussed in Fig. 20 caption), and it is not possible to assess the reliability of this solution. The waveforms for this event are too small to confirm its unusual mechanism by waveform modeling (Jackson et al., 2002). However, normal faulting could accompany flexural-slip folding, producing normal faulting on the convex surface of

TABLE 11. RECORDED SEISMICITY ALONG THE GARMSĀR FAULT (SEE FIGS. 8 AND 9) Date (yr.mo.d) Area Ms Ml mb 2015.04.06 3.0 2011.11.20 3.4 2011.10.07 1988.12.08 Garmsār 4.4 1988.10.26 Garmsār 4.7 1988.10.24 Garmsār 4.9 4.4 1988.10.23 Garmsār 4.6 4.6 1988.08.23* Bonkuh 5.0 4.8 1988.08.22* Bonkuh 5.0 5.0 1983.05.29 Garmsār 4.4 1982.10.25 Garmsār 4.6 1954.09.02 Garmsār 4.5 1945.05.11 Bonkuh 4.7 Note: See Table 3 for additional data and sources. MI—local magnitude; mb—body-wave magnitude; Ms—surface-wave magnitude; Mw—moment magnitude. *Epicenters are located near the junction of the Garmsār and Pārchin faults (Fig.9).

Mw

4.5 4.9 5.0 5.1 5.3 5.1 4.6 4.8 4.5 4.6

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

123

folded strata, but its trend does not match with the nearly E-W of the folds on the hanging wall of the Garmsār fault; it matches with the fold axis on the Pishvā fault. Ashoori (2004) reported that the 23 July 1927 (Ms 5.9), 28 October 1845 (Ms 5.5), 2 June 1957 (Ms 6.8), and 7 June 1957 (Ms 6.7) earthquakes took place along the Garmsār fault, which is not confirmed here. Ashoori (2004) also speculated that the 22 September 1953 Ms ~4.0 earthquake was the result of motion along the Garmsār and Ejdehā faults, salt tectonic, and salt domes northwest of Garmsār. No trench study has been conducted on the Garmsār reverse fault. The 1999–2000 microearthquake survey located by temporary networks showed northward deepening of the seismicity associated with the Garmsār thrust (Ashtari et al., 2005). The 2006 microearthquake survey revealed high seismic activity on the central and western parts of the Garmsār fault, with three focal mechanisms showing compressional movements of the central section of the fault (Tatar et al., 2015). 14. THE FIRUZKUH FAULT The Firuzkuh active fault (Berberian et al., 1996; Jackson et al., 2002) is a 38-km-long (in two segments of 26 km in the northeast and 12 km in the southwest), NE-SW–striking, leftlateral strike-slip fault between the Moshā fault on the west and the Āstāneh fault on the east (Berberian et al., 1996). It is located ~128 km east of Tehran (Fig. 9). Its left step along the northeastern continuation of the Moshā fault marks it as part of the Moshā fault system in the Firuzkuh Plain. Left-lateral displacement along the Firuzkuh fault is transferred with a right-step jog to the Āstāneh fault farther east (Fig. 9). Leftlateral deformation east of the bend in the Alborz appears to continue from the eastern Moshā fault to the Firuzkuh and Āstāneh faults farther east (Berberian et al., 1996; Jackson et al., 2002). The fault, with clear active geomorphologic indicators visible on aerial photographs and satellite imagery, dips gently southeast, and the northwestern block (the Firuzkuh plain) has subsided. It is parallel to regional reverse and strike-slip faults in the central-eastern Alborz (Fig. 9), supporting a spatial partitioning of strike-slip and reverse components on the subparallel faults with orthogonal slip vectors (Jackson et al., 2002). The GPS data indicate 2.5 ± 1.5 mm/yr leftlateral shear and 0.5 mm/yr fault-perpendicular shortening across the east-central Alborz Mountains at the longitude of the Firuzkuh fault (Fig. 6; Djamour et al., 2010; Mousavi et al., 2013). A slip rate of 2.3 mm/yr has been estimated for this fault (Ritz et al., 2006; Nazari et al., 2007). A digital elevation model indicated the ratio H/V between the horizontal (H = 15 m) and the vertical (V = 25 m) components to be 0.6:1 (Ritz et al., 2006). The ratio of horizontal to vertical displacement across the fault, calculated from the displacement of landscape features, is 7.6:1 (Nazari et al., 2014).

124

Berberian and Yeats

14.1. Seismicity (Firuzkuh Fault) Despite its active fault features, no recorded historical earthquake has been found along the Firuzkuh fault in historic annals. The constrained meizoseismal area of the 20 January 1990 Mw 5.9 Gaduk, NE Firuzkuh, earthquake (Fig. 14) is located along the northeastern tip of the Firuzkuh fault (Berberian et al., 1996). The Harvard centroid moment tensor (CMT) solution with leftlateral strike-slip motion along an E-W fault has a fixed default depth of 33 km, and the strike is not well constrained. Insufficient clear waveforms occur for a full inversion of this event. The fit of synthetic P waves to the observed P seismograms is poor, mainly because, at the 33 km default depth, the centroid is too deep. A better fit to P waves is achieved when the centroid is moved to 13 km, but the fit of the SH waveforms remains poor. The SH waveforms fit better when the strike is rotated from 090° to 075°, corresponding better to the strike of the Firuzkuh fault (Jackson et al., 2002; Fig. 14). Field investigation of the 20 January 1990 Mw 5.9 Gaduk earthquake did not show any coseismic surface rupture along the Firuzkuh fault (Berberian et al., 1996). Buildings at the town of Firuzkuh were slightly damaged (VI), and two people were injured. The earthquake was followed by numerous teleseismically recorded aftershocks, the largest of which took place 48 min after the main shock with mb 4.8. This “aftershock” damaged Chāshm village along the Chāshm thrust bifurcating near the northeastern tip of the Firuzkuh fault (Fig. 14), located ~32 km ENE of Gaduk, even though Chāshm was not damaged during the main shock (Berberian et al., 1996). The epicenter of the 23 November 1995 mb 4.0 earthquake (ISC) was located on the Chāshm thrust. 14.2. Firuzkuh Paleoseismologic Trench Study A paleoseismic trench study across the Firuzkuh fault was conducted in two trenches, TF1 at 35°47′20″N, 52°51′05″E and TF2 at 35°47′12″N, 52°50′43″E (Nazari, 2006 2015; Nazari et al., 2014). Nazari (2006) presented two events (with displacements Ev1 = 1.2 m and Ev2 = 1.0 m), whereas Nazari et al. (2014) observed five events in the two trenches (Fig. 14; Table 12): (1) Event 0: with 4 cm vertical displacement and estimated 30 cm strike slip, cutting sedimentary unit 5 in trench TF1. No rupture dating was available in trench TF1; however, the authors interpreted the timing of the last surface rupture on the fault in TF1 to be close to the age of human skeletal remains in the second trench (TF2), where human bones yielded a radiocarbon age of 1159 ± 28 cal yr B.P. (763–819 CE calendar yr; Nazari et al., 2014). This “indirectly dated event” was not recorded in the literary sources. (2) Event 1: with 2.40–4.40 m vertical displacement, corresponding to an estimated 18.2–33.4 m strike-slip displacement. The average horizontal slip rate was calculated to be 1.1–2.2 mm/yr. The authors assumed that the

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

displacement occurred during several earthquakes with an estimated strike-slip Mw 8.05–8.29 event occurring since 15.9 ± 0.9 ka (Nazari et al., 2014). The magnitude of Mw 8.05–8.29 is highly exaggerated for the Alborz Mountains and the 38-km-long Firuzkuh fault would not be expected to have such seismic capability. (3) Event 2: with vertical displacement of 0.95 m and estimated strike-slip Mw 7.76 occurring in the interval between 27.1 ± 1.7 ka and 15.9 ± 0.9 ka (Nazari et al., 2014). Similar to the previous event, the 38-km-long Firuzkuh fault does not have the seismic capability of generating an earthquake of Mw 7.76. This magnitude is exaggerated for the 38-km-long Firuzkuh fault. (4) Event 3: with indirect estimated vertical displacement of 0.60 m and estimated strike-slip Mw 7.51 occurring before 27.1 ± 1.7 ka (Nazari et al., 2014), which also seems to be an exaggerated magnitude for the 38-kmlong Firuzkuh fault. (5) Event 4: with unknown parameters (Nazari et al., 2014). Nazari et al. (2014) assumed a 55 km fault length in each paleo-event with a maximum magnitude of 7.1. With an estimated late Quaternary left-lateral slip rate of ~1.1–2.2 mm/yr, magnitude of Mw 7.1, and ~1.2 m average displacement, the authors estimated a recurrence period of ~1150–540 yr along the Firuzkuh fault. The fact is that the Firuzkuh fault is a 38-km- long fault, composed of two segments of 26 km in the northeast and 12 km in the southwest. Estimating an Mw 8.05–8.29 for the strike-slip motion of event 1 (Nazari et al., 2014; Table 12) does not sound reasonable considering the 38 km length of the Firuzkuh fault and the seismotectonic history of the Alborz Mountains reviewed in this study. Furthermore, no evidence exists of an earthquake with Mw >8.0 during two millennia of historical earthquakes on the Iranian Plateau (Appendix B). The largest post-1900 earthquake in Iran was the 23 January 1909 Mw 7.4 Silākhor earthquake with coseismic surface strike-slip faulting along the Zāgros Main Recent fault (Berberian, 2014). Nazari et al. (2014, p. 133, 134) and Nazari (2015, p. 264) argued that “since the 856 Qumis [Komesh] earthquake with a death toll of 200,000 and complete destruction in a region of ~150 km length” took place ~50 yr after their assumed

Figure 14. Firuzkuh fault map (northeastern continuation of the eastern segment of the Moshā fault) with epicentral area of the 20 January 1990 Mw 5.9 Gaduk earthquake and location of a trench dug across the Firuzkuh fault (stippled rectangles; see text for trench result and reference). For location, see Figure 1. Symbols are as in Figure 1. Fault-plane solution is from Jackson et al. (2002). Inset top left: eastern segment of the Moshā fault with left-step continuation as the Firuzkuh fault and location of the 1935 earthquake showing a southwest migration of seismicity. Intensities are in modified Mercalli intensity (MMI). Figure is modified from Berberian et al. (1996). Bottom-right inset is as in Figure 2.

Tehran: An earthquake time bomb

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

125

126

Berberian and Yeats TABLE 12. SUMMARY OF PALEOSEISMIC EVENTS IN TRENCHES ACROSS THE FIRUZKUH FAULT (NAZARI ET AL., 2014) AND TELESEISMIC DATA WITH UNKNOWN LOCATION ERRORS (FIG. 14)

Event



Vertical/horizontal displacement (m)

Average displacement (m)

Mw general fault

Mw normal fault

Mw strikeslip fault

Date













Earthquakes (yr.mo.d)

Time (h:min)

mb

1996.08.25 1990.01.20 1990.01.20 1973.10.27 –

14:17 02:15 01:27

4.2 4.2 5.5

Ms

Ref.

ISC ISC ISC ISC

5.8 – 4.3 Post–1159 ± 28 – – – cal yr B.P. (763–819 CE)* 1 2.40–4.40/18.2– 13.82–25.34 7.86– 7.52– 8.05– Post–15.9 ± 0.9 33.4 8.08 7.69 8.29 ka 2 0.95/– 5.47 7.53 7.26 7.76 Between 27.1 ± 1.7 ka and 15.9 ± 0.9 ka 3 0.60/? 3.40 7.36 7.13 7.51 Pre–27.1 ka 4? – – – – – Pre–27.1 ka Note: Bold: The very large magnitudes (see Table 1 in Nazari et al., 2014) seem to be unreal for the 38-km-long Firuzkuh fault, as well as considering the seismotectonic and seismic history of the Alborz Mountains reviewed in this study. Mw—moment magnitude; mb—body-wave magnitude; Ms—surface-wave magnitude. ISC—International Seismological Summary, Berkshire, UK (1913–1963; now ISC). *No rupture dating was available in trench TF1; the dating was obtained from human skeletal remains in the second trench (TF2) where no rupturing was observed. – 0

– 0.04/0.3

– –

– –

– –

763–819 paleo-event, the Firuzkuh fault either ruptured during the 856 Komesh earthquake, or it was triggered by the 856 event. It should be mentioned that the distance between the Firuzkuh trench (where the proposed 763–819 paleo-event is indirectly dated and reported) and the center of destruction of the 856 earthquake (on the Dāmghān fault to the northeast) is 160 km (Figs. 9 and 15); therefore, the Firuzkuh fault did not rupture during the 856 earthquake, which destroyed the city of Dāmghān. As discussed in Appendix B, no basis exists for the “200,000 death toll” along the highly exaggerated unconstrained meizoseismal region of 150 km length associated with the 856 Komesh earthquake (Berberian et al., 1996; Berberian, 2014; see Appendix B for the analysis of this event). Therefore, the first scenario as the Firuzkuh fault being the western continuation of the 856 surface rupture, as presented by Nazari et al. (2014) and Nazari (2015), is definitely not supported by the historical seismic data or the contemporary and near-contemporary literary sources (see Appendix B). Furthermore, Nazari (2015) stated that the old Ray city was located in the realm of the 856 large earthquake, which is not the case. The distance between the city of Ray and the destroyed city of Dāmghān during the 856 earthquake is 270 km (Fig. 9). Regarding the second proposed scenario by Nazari et al. (2014) and Nazari (2015), the indirectly dated 763–819 paleo-event has not been documented in the literary sources. We do not question the rupturing events in the trenches, but the poorly constrained dates, fragmentary single trench data, and overinterpretation of the paleo-events, especially speculative connections with the historical earthquakes that took place 160 and 270 km away, deserve careful scrutiny.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

– –

15. THE ĀSTĀNEH FAULT The Āstāneh (Cheshmeh ‘Ali in Berberian, 1976c) fault is a left-lateral, nearly vertical strike-slip fault in the Āstāneh valley west of the city of Dāmghān (Figs. 9 and 15). The NE-SW–trending Āstāneh fault has an 80 km length; to the northeast, the fault sharply bends to E-W and continues for 60 km as the Dāmghān fault (Berberian, 1976c; Berberian et al., 1996; Jackson et al., 2002; Berberian, 2014; Fig. 15). Slip rates of 2.5 mm/yr (Nazari et al., 2007), 1.7–2.2 mm/yr (Hollingsworth et al., 2010), and 1.3–2.5 mm/yr (Rizza et al., 2011) have been estimated for the Āsāneh fault. Since the Āstāneh fault has been considered in the literature as: (1) the source fault of the 856 Komesh earthquake; (2) the largest earthquake in the Alborz and Iran by some authors, with an estimated magnitude of Ms 7.9 (Hollingsworth et al., 2010; Nazari et al., 2014; Nazari, 2015; see analysis in Appendix B); (3) allegedly rupturing, together with the Firuzkuh fault, during the same earthquake as one that affected the city of Ray (Nazari et al., 2014; Nazari, 2015); and (4) a major hazard to Tehran (Nazari et al., 2014; Nazari, 2015) at a distance of >250 km, we review the paleoseismic trench study along the fault in this section. 15.1. Seismicity (Āstāneh Fault) The 22 December 856 Komesh (modern Dāmghān) earthquake, with its overestimated magnitude and meizoseismal area (see Appendix B for analysis), took place along the Dāmghān fault (Berberian et al., 1996; Berberian, 2014), which is the northeastern continuation of the Āstāneh fault east of its bend

Figure 15. Proposed meizoseismal area of the 22 December 856 Komesh (modern Dāmghān) earthquakes in eastern central Alborz Mountains (for location, see Fig. 1). Symbols as in Figure 1. Filled squares—sites destroyed by the 856 earthquakes (triangle with letter C—destroyed caravanserai near the Dāmghān fault destroyed by the 856 earthquake; see Appendix B for discussion). The speculative meizoseismal areas proposed by Ambraseys and Melville (1982) (A+M, 1982), Hollingsworth et al. (2010) (H.et.al. 2010), and Bune and Gorshkov (1980) (B & G, 1980), crossing several active fault features, are shown by broken line ellipses. Small rectangle across the Āstāneh fault (NW of Dāmghān) indicates location of paleoseismologic trench (Hollingsworth et al., 2010; see the text for discussion). Triangles—archaeological sites with dates (see Appendix B). FM (circle near the Āstāneh fault)—Fulād Mahalleh. K—Kiyāsar; T—Tazareh. Figure is modified from Berberian et al. (1996) and Berberian (2014). Inset lower right: sites that reported strong ground shaking during the 856 Komesh earthquake. Top-left inset is as in Figure 2.

Tehran: An earthquake time bomb

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

127

128

Berberian and Yeats

from NE-SW to E-W strike (Fig. 15). Contemporary chronicles show that the earthquake had a magnitude of ~Mw 7.1 (Appendix B). Destruction of the Hecatompylos (Sad-Darvāzeh; i.e., Hundred Gates) city by this earthquake cited by Ambraseys and Melville (1982), Hollingsworth et al. (2010), Nazari et al. (2014), and Nazari (2015) cannot be substantiated by archaeological data (see Appendix B for discussion). A less well-documented earthquake took place at the Gerdkuh Fortress (Dezh-e Gonbadān; Fig. 15), 18 km west of Dāmghān, in ca. 1102 (Sani’ al-Dauleh, 1880–1882; Ambraseys and Melville, 1982). However, no contemporary or near-contemporary source or record of damage to Dāmghān was reported. The epicenters of the 10 November and 10 December 1967 earthquakes (both of Ms 4.9; ISC) were located near the Āstāneh fault; the epicentral errors of these small-magnitude earthquakes in the 1960s are large (Ambraseys, 1978; Berberian, 1979a; Berberian et al., 1996, p. 207–208), and no macroseismic data are available to constrain the locations. The 2007–2008 microearthquake survey located by temporary networks showed clusters of microseismicity around the Āstāneh (Fig. 15) and Chāshm (Fig. 14) faults (Nemati et al., 2011). A single event was recorded along the Āsāneh fault with a left-lateral focal mechanism (event 14 in Nemati et al., 2011). 15.2. Āstāneh Paleoseismologic Trench Study A paleoseismologic trench study across the Āstāneh fault ~30 km WNW of Dāmghān (36.25°N, 54.00°E; Fig. 15) indicated three events between 11.62 ± 0.62 ka and 645–402 ka (1306–1363 cal. CE) with very large error bars and with an esti-

mated average return period of 3.7 ± 4.2 k.y. between R1 and R2 (Hollingsworth et al., 2010; Table 13): (1) R3: 600 BCE–1300 CE (between 1306 and 1363 and 772–506 cal BCE); (2) R2: 4600–600 BCE (between 772 and 506 and 4729– 4546 cal BCE); and (3) R1: 9600–4600 BCE (between 4729 and 4546 BCE and 11.62 ± 0.62 ka). The exact amount of displacement by each paleo-event, their relative magnitudes, and total surface rupture length are not discussed. The authors concluded that their last paleo-event (R3: 600 BCE–1300 CE) is consistent with the 856 Komesh (modern Dāmghān) earthquake; and the second event (R2: 4600– 600 BCE) may correspond to a known earthquake at the Hessār mound (SE Dāmghān, Fig. 15) reported by Houtum Schindler (1877) and Ambraseys and Melville (1982). Hollingsworth et al. (2010), without offering supporting data, assumed that the destruction of the Dāmghān region during the 856 earthquake may have resulted from multiple slip events or clusters of earthquakes along the Āstāneh, Dāmghān, and North Dāmghān faults over a short period of time (Fig. 15). This speculation cannot be confirmed, since almost all the chronicles refer to a single earthquake on 22 December 856 during the night destroying Dāmghān (discussed in Appendix B). The scenario of rupturing of more than one fault in a single earthquake in 856 (Hollingsworth et al., 2010; Nazari et al., 2014; Nazari, 2015), which is also presented for the Tāléqān and Mosā faults for the 958 earthquake in the west (Nazari et al., 2009; Ghassemi et al., 2014), lacks any supporting evidence (see Appendix B for discussion).

TABLE 13. SUMMARY OF PALEOSEISMIC EVENTS IN TRENCHES ACROSS THE ĀSTĀNEH FAULT (HOLLINGSWORTH ET AL., 2010) AND HISTORICAL AND TELESEISMIC DATA WITH UNKNOWN LOCATION ERRORS (FIG. 15) Event

Date

Average displacement (m)

M

Rupture length

Earthquakes

mb

Ms

Mw

Sources

(yr.mo.d)











1 9 8 6 . 0 3 . 26

4. 6











1 9 6 7 . 1 2 . 10

4. 9

I SC











1 9 6 7 . 1 1 . 10

4. 9

I SC











ca . 1102



I SC

SD









856.12.22

R3

600 BCE–1300 CE (between 1306–1363 and 772– 506 cal BCE)

?

?

?





7.2

R2

4600–600 BCE (between 772–506 and 4729– 4546 cal BCE)

?

?

?





R1

9600–4600 BCE (between 4729–4546 and 11.62 ± 0.62 ka)

?

?

?





Appendix B –















Note: M—magnitude; mb—body-wave magnitude; Ms—surface-wave magnitude; Mw—moment magnitude; ISC—International Seismological Summary, Berkshire, UK (1913–1963; now ISC); SD—Sani’ al-Dauleh (1880–1882, I.181). For sources and additional teleseismic parameters, see Table 3.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

Tehran: An earthquake time bomb Regarding the reported earthquake at the Hessār archaeological mound (Houtum Schindler, 1877; Ambraseys and Melville, 1982), we know that the cause of violent destruction of the building in Hessār IIIB (NE Phase B: ca. 2420–2290 BCE) was an attack followed by a severe fire at the end of Hessār IIIB around 2000 BCE (Dyson and Remsen, 1989; Dyson and Howard, 1989; see also Appendix B). 16. THE TĀLÉQĀN FAULT The Tāléqān fault (Figs. 4 and 16) is located south of the Tāléqān intermontane valley in the central Alborz, more than 50 km northwest of Tehran (Berberian et al., 1983). We review this fault since some authors have speculated that it was the source of the 312–280 BCE Ray and 958 Ruyān earthquakes (Nazari et al., 2009; Nazari, 2015) that destroyed the city of Ray in southern Tehran (Fig. 9). The fault was first mapped as an E-W–trending, north-dipping normal fault (Dedual, 1967; Meyer, 1967; Sieber, 1970). Annells et al. (1975a, 1975b) mapped the fault as a southdipping high-angle reverse fault bringing Upper Proterozoic– Lower Cambrian rocks onto the Eocene Karaj Formation. The fault was mapped as a 60-km-long, south-dipping reverse fault (GSI, 1975, 1985, 1991, 2001; Berberian et al., 1983). Nazari et al. (2009) connected the eastern end of the mapped Tāléqān fault (Annells et al., 1975a, 1975b; GSI, 1975) to the western end of the Garmābdar fault (Assereto, 1966) as an 80-km-long, south-dipping, left-lateral strike-slip fault with a normal component and change of strike from E-W (in the west) to NW-SE (in the east) (Fig. 8). A digital elevation model indicates the ratio H/V between the horizontal (H = 13 m) and the vertical (V = 17 m) is 0.76:1 (Ritz et al., 2006). A left-lateral slip rate of 1 mm/yr and an extension rate of 0.5 mm/yr are estimated for the fault (Nazari et al., 2007, 2009). Numerous paleo–rock avalanches and landslides have occurred in the vicinity of the Tāléqān fault (Fig. 16). An important superficial rock failure occurred in the Tāléqān fault zone, and landslip deposits resulting from it are preserved extensively on the southern side of the Tāléqān valley immediately north of the Tāléqān fault, covering an area of 32 km2 (Fig. 16). These deposits rest on folded Neogene sediments and are themselves locally overlain by undated terrace gravels from the early or late Pleistocene. They include many coherent masses, mostly of Permian Ruteh Limestone, but also Mesozoic limestone, which is now no longer represented by exposures upslope (Annells et al., 1975a, 1975b, 1985; Berberian, 1994). 16.1. Seismicity (Tāléqān Fault) The well-constrained meizoseismal area of the 8 November 1966 mb 5.0 Samghābād (Tāléqān) earthquake suggested that this event took place along the Tāléqān fault (Fig. 16; Berberian et al., 1983). The connection of the 958 Ruyān earthquake of NE Tehran to the Tāléqān fault, as well as assigning the poorly known 1428 earthquake to the Tāléqān fault (Nazari

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

129

et al., 2009), which unfortunately entered the literature, are discussed in Appendix B. 16.2. Tāléqān Paleoseismologic Trench Study Paleoseismic trench study (T1: 36°07′25.7″N, 51°12′38.1″E, and T2: 36°07′26.0″N, 51°12′37.4″E; +2995 m; Figs. 16, 17) shows that fault displacement is left-lateral strike slip with a normal component (Ritz et al., 2006; Nazari et al., 2007, 2008, 2009). Originally four paleoseismic events were identified in trench Ta1 (T1) and two in trench Ta2 (T2), with estimated magnitudes of 6.87–7.27 (Nazari, 2006, p. 184–188). The results were later modified to two or three events in T1, and two events in T2 with modified parameters (Nazari et al., 2009). The latter showed two earthquakes in T1 with magnitudes ranging from Mw 7.2 to 7.7 during the past 5300 yr. Considering a three-event scenario, an average recurrence period of ~2000 yr was obtained, whereas a two-event scenario gave an interval of 3295 ± 1465 yr (Table 14). Their estimated minimum horizontal and vertical slip rates were 0.6–1.6 mm/yr and ~0.5 mm/yr, respectively. Nazari et al. (2009) stated that the kinematic change from reverse to strikeslip faulting occurred ca. 1 Ma. The youngest paleo-event (Ev. 1 in trench T1) is considered to be younger than 80 CE, with estimated magnitude Mw 7.2–7.6. Nazari et al. (2009) identified this event from their colluvial unit 20, which covered and eroded underlying units (30, 31, 32, 33, and 40) without evidence for a clear faulting episode, but with evidence for offset at its immediate base (Table 14). Because no displacement could be measured for this event, Nazari et al. (2009) concluded that colluvial unit 20 could have been a deposit triggered by a strong historical earthquake in 855, 958, 1177, 1665, or 1830 on the Tāléqān fault. As discussed earlier and in Appendix B, the 855(–856) earthquake was felt strongly at the city of Ray, ~70–80 km southeast of the Tāléqān fault. The macroseismic data from the 958 Ruyān earthquake in NE Tehran (Appendix B, Figs. 18–20), with overestimated magnitude of M 7.7, which was thought to have originated on the Tāléqān fault (Nazari et al., 2009; Nazari, 2015), suggest that it took place along the central segment of the Moshā fault at Ruyān, north of Ray (Figs. 4, 8, 9, 16, 17, and will be discussed in reference to the 958 and 1830 earthquakes and Fig. 20; Appendix B). The 1177 CE earthquake, which caused damage to the cities of Ray and Qazvin, could be attributed to the western segment of the North Tehran thrust (Figs. 8 and 9); however, it requires further paleoseismic study. The 1665 event took place along the central segment of the Moshā fault, destroying the town of Damāvand, ~145 km southeast of the Tāléqān fault (Figs. 8 and 9; Appendix B). Finally, the meizoseismal area of the 1830 Lavāsānāt earthquake along the central segment of the Moshā fault was located ~40 km northeast of Tehran and 80 km southeast of the Tāléqān fault (Fig. 16; Table 14). The second event (Ev. 2 in trench T1) proposed by Nazari et al. (2009) is based on warping/drag of their units 40–41 along the fault, and it was dated sometime between 3470 and 1540

Figure 16. Fault map of the Tāléqān Valley 80 km NW of Tehran (for location, see Fig. 1) and location of the paleoseismic trench on the Tāléqān fault (stippled rectangle; see text for results and references). Symbols as in Figure 1. The paleo-landslides and rock avalanches, most between the Tāléqān River and the Tāléqān fault, are stippled with arrows showing the slip direction. Filled triangles—archaeological sites (see Appendix B). Figure is modified from Annells et al. (1975a, 1975b) and Berberian et al. (1983). Bottom-left inset is as in Figure 2.

130 Berberian and Yeats

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

Tehran: An earthquake time bomb

131

Figure 17. (Top) Meizoseismal areas of medium- to large-magnitude earthquakes near Rhagae (modern Ray) in the southern Tehran metropolitan area (diagonal lines). For location, see Figure 1. Symbols and bottom-left inset of top panel are as in Figure 2. AK—Ayvān Kay; F—Firuzkuh; M—Moshā; MF—Moshā fault; NQT—North Qazvin fault; NTF—North Tehran fault system. (Bottom) Time-space diagram of earthquakes in the Tehran region. Queried where meizoseismal areas are not well defined and constrained. Distances are along strike with respect to the capital city of Tehran. Meizoseismal areas of the ca. 280 BCE Ms ~≥7.0, 1119 Ms ~6.5 Qazvin, and 1177 Ms ~7.0 Ray-Qazvin earthquakes are not constrained (see text and Appendix B for the interpretation). Where date of the earthquake is shown, it is given by year. month.day. Figure is modified from Berberian and Yeats (1999) and Berberian (2014).

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user







1.51– 1.63

1.12

7.22





7.20– 7.22

7.10

6.87



Mw





T1/E3

T1/E2

T1/E1



Event

Trench T1 (Ta1)





4800– 5300 cal CE?

3470– 1540 cal yr B.P.

1870 cal yr B.P. (post–80 CE)



Date







?

1.73–5.11

1.46–1.86/

1.12–3.70

0.94–1.34/



V/H offset (m)

(2)







?

7.4–7.7

7.2–7.6



Mw



T2/E2

T2/E1









Event



Post–3830 cal yr B.P.

Post–3830 cal yr B.P.









Date (cal. yr B.P.)

(2)



?

? –

7.2–7.5









Mw

1.0/1.2– 2.75









V/H offset* (m)

Trench T2 (Ta2)

Note: References: 1—Nazari (2006); 2—Nazari et al. (2009). Mw—moment magnitude; mb—body-wave magnitude. *Horizontal offset is inferred from the vertical displacement.

5282–4974 to 4224–4415 cal yr B.P.





T1/E4



3687–3588 cal yr B.P.

T1/E3



Pre–134–339 cal CE (1780 ± 60 yr B.P.)

T1/E2

– 0.65



1600 cal CE (125 ± 40 yr B.P.) to 134– 339 cal CE (1780 ± 60 yr B.P.)

Offset (m)

Date

T1/E1



Event

(1)

TABLE 14. SUMMARY OF PALEOSEISMIC EVENTS IN TRENCHES T1 AND T2 ACROSS THE TĀ LÉQĀN FAULT AND HISTORICAL EARTHQUAKES (FIGS. 9 AND 16)













1966.11.08

Date (yr.mo.d)













4.8

mb













5.0

Mw

Historical earthquakes (this study)

132 Berberian and Yeats

Figure 18. Fault map and major archaeological sites (filled triangles) in Rhagae (modern Ray) and Tehran (see Appendix B); for location, see Figure 1. Caspian Gates (Caspiae Portae; Sardarreh Defile) are located in the lower-right corner along the Sardarreh River between the southeastern Pārchin and Garmsār faults (see Appendix B for discussion). NTF—North Tehran fault system. Symbols and bottom-left inset are as in Figure 2. Contour lines are in meters above mean sea level. Radiation symbols in urban Tehran-Ray region: TNRC— Tehran Nuclear Research Center; IAP/PHRC—Lavizān-Shiyān (Institute of Applied Physics [IAP]; later, Physics Research Center [PHRC]) in eastern Tehran; and Pārchin military nuclear complex in SE Ray. Archaeological sites: A—‘Abbāsābād; D—Darrus; F—Fakhrābād; GH—Ghār; H—Zendān Hārun; M—Mehdikhāni; MO—Mortezāgerd; Q—Qolhak; QA—Qaytarieh; S—Sādeghābādi.

Tehran: An earthquake time bomb

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

133

Figure 19. The boundary of the Ruyān (Rudān) Mountains (thick dashed line) north of Ray, between the Tehran-Ray plain in the south and the northern old Tabarestān (modern Māzandarān) province located to the south of the Caspian Sea, used to locate the source area of the 958 CE earthquake that destroyed Ray, south of Tehran (see Fig. 8; Table 2; Appendix B for discussion); for location, see Figure 1. The Ruyān Mountains (Inner Qasrān/Kuhsārān) are between the Damāvand volcano in the east and the Karaj River in the west. Symbols and bottom-left inset are as in Figure 2. Triangles (stippled)—location of paleo-landslides possibly triggered or reactivated by earthquakes. Where date of the earthquake is shown, it is given by year.month.day. NTF—North Tehran fault system; GAL—Galukān, south of Dizeh, WNW of the North Tehran and Moshā fault junction; DQ—Dizeh Qasrān.

134 Berberian and Yeats

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

Figure 20. Seismicity and faults of central Alborz. Symbols and bottom-right inset are as in Figure 2 (for location, see Fig. 1). The location of the Ruyān (Rudān) Mountains (stippled), the epicentral area of the 958 Ruyān earthquake, is added along the Moshā fault north of Ray and Tehran, west of the Damāvand volcano. Two filled squares mark sites damaged during the 1830 Lavāsānāt earthquake (see Table 2 and Appendix B for discussion). The city of Tehran in 1867 is shown by a filled scaled hexagon; it was damaged during the 1830 earthquake (~VII+ MMI). Ray (filled circle) also underwent intensity of ~VII (MMI) during this earthquake. Estimated modified Mercalli intensity (MMI) values for other sites to the north and east are for the 1830 event. Fault-plane solutions (with magnitude in parentheses) are given for: 1983.02.25 and 26, and 1988.08.22 and 23 (Harvard CMT, 2016); and 1990.01.20 with 13 km centroid depth (Jackson et al., 2002) earthquakes, where date of the earthquake is given by year.month.day. Three smaller centroid moment tensor (CMT) solutions (combined inversion of broadband and short-period waveform) of three small-magnitude earthquakes along the Moshā fault were taken from Donner et al. (2014): western CMT: 2006.12.20 (Mw 3.8, centroid depth 14–18 km), center CMT: 2004.02.21 (Mw 3.9, centroid depth 14–16 km), and eastern CMT: 2004.09.24 (Mw 4.1, centroid depth 4–6 km).

Tehran: An earthquake time bomb

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

135

136

Berberian and Yeats

cal yr B.P., with estimated Mw 7.4–7.7 (Table 14; Fig. 16). The authors stated that their second event could be associated with the 312–280 BCE Ray earthquake with Ms 7.6, and they estimated a maximum coseismic offset between 2.26 and 5.10 m. The 312–280 BCE event destroyed the ancient city of Rhagae (Ray) and villages to its east and might have taken place on the Pārchin fault, with its meizoseismal area ~110 km southeast of the Tāléqān fault (Figs. 8 and 9; Appendix B). Nazari et al. (2009, p. 1033) claimed that if the 312–280 BCE earthquake occurred on the Tāléqān fault, then the event was “misallocated in the historical catalogue”! Historical records of this event make no reference to destruction of the Tāléqān region during this time, and the event was not misallocated at Rhagae (Ray) (see Appendix B for discussion). Ironically, Nazari (2006), Nazari et al. (2007, 2008, 2011), and Ritz et al. (2012) claimed that their penultimate event in the Vardāvard trench in the North Tehran fault zone (E2, Mw 6.6–7.1) could correspond to the 312–280 BCE earthquake at Ray—a single earthquake with two different sources. Based on cumulative displacement at the Tāléqān fault, Nazari et al. (2009) concluded that more than two events could have been recorded in a sag pond behind the fault scarp. Considering a three-event scenario, an average recurrence interval was calculated as ~2000 yr, whereas considering a two-event scenario, the time interval between the two events was 2295 ± 1465 yr, and the elapsed time since the last event ranges between 3529 and 1599 yr (Nazari et al., 2009). Although we do not place much weight on the empirical relations between surface fault rupture and earthquake moment magnitude (Wells and Coppersmith, 1994), we cannot find any evidence for an earthquake of magnitude 7.7 in the Alborz. The most recent event in trench T2 was dated post-3830 cal yr B.P., with estimated Mw 7.2–7.5, and the second event was dated between 3470 and 8110 cal yr B.P. (Table 14; Fig. 16). 16.3. Archaeoseismicity (Tāléqān) There are numerous archaeological mounds from the fifth and fourth millennium BCE in the Tāléqān area. Those located south of the Tāléqān fault are Ganj Tappeh, Siyāh Tappeh, Chandār, Mushélān/Esmā’ilābād (35°52′50.62″N, 50°41′22.59″E), Khorvin (Fig. 16), and many more (Navai, 1976; Hakemi, 1959; Talai, 1983, 2000; Mousavi, 2001). The closest mound is Ganj Tappeh, located ~12 km south of the Tāléqān fault and 3.5 km south of the eastern segment of the Moshā (Maydānak segment) fault (Fig. 16). Archaeoseismologic investigation of these mounds together with additional paleoseismologic trench studies may help in understanding the seismicity along these faults. 17. A TEHRAN SEISMIC GAP The historical seismicity data (Fig. 8), the fault map of the central Alborz and northern Central Iran (Fig. 9), the seismicity–fault map of the Tehran region (Figs. 5, 17–20), and limited paleoseismic data suggest that the megacity of Tehran is located

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

in a seismic gap in the southern central Alborz. Our study shows that the fault segments in and near Tehran have not ruptured in historical or modern earthquakes, and those fault sections have a greater chance of rupturing in the near future because of the buildup of strain (Fig. 17). Furthermore, the 958, 1665, and 1830 earthquakes along the central segment of the Moshā fault (Figs. 4, 8 and 9) might have loaded the North Tehran fault system, branching off the Moshā fault to the west toward Tehran, and the western segment of the Moshā fault, where no mediumto large-magnitude earthquake has happened at least since the 1830 earthquake (Fig. 17). Since the controversial nature of the Kahrizak and North and South Ray escarpments has not yet been resolved (discussed earlier herein), we do not include them in this discussion. In general, our interpretation of the meizoseismal areas of the ca. 280 BCE Rhagae/Ray (Mw ~≥7.0; though neither the meizoseismal area nor the magnitude is constrained); 958 (Mw ~≥7.0) Ruyān; 1177 (Mw ~7.0? not constrained) Ray-Qazvin; 1666 (Mw ~6.5) Damāvand; and 1830 (Mw ~7.0–7.3) Lavāsānāt earthquakes suggests that the faults in the immediate vicinity of Tehran have not sustained a major earthquake in more than 2000 yr. The 1830 earthquake on the Moshā fault was located ~30 km northeast of the city (Figs. 4, 17–20). Our seismic gap analysis can be summarized as follows: (1) The North Tehran fault system (Fig. 2): There is no evidence of seismic activity of the North Tehran fault system from its junction with the Moshā fault in the east (including the Niāvarān, Darakeh, Farahzād, and Kan left-lateral strike-slip faults and the associated thrust system) to north and northwest of Tehran (the Karaj segment might have ruptured during the 1177 event, but the meizoseismal area is not constrained; Figs. 8 and 9) for the past 2000 yr. So far, the preliminary paleoseismicity results, though ages are unconstrained, also support the historical seismicity records. A properly designed paleoseismic trench study, especially along the mentioned strike-slip faults, will help in understanding the last time these faults ruptured and their recurrence period. (2) The inner-city Dāvudieh, Mahmudieh, Bāgh-e Fayz, and Takht-e Tāvus faults, and blind thrusts (Fig. 2): These reverse and strike-slip faults, as well as other possible unmapped blind thrusts responsible for folding the Pliocene–Pleistocene deposits of the Hezārdarreh and younger formations, have not yet been shown to be sources of major historic earthquakes. Paleoseismic trench study is required to document the seismic history of these faults. (3) The northwestern section of the Pārchin fault (Fig. 4): This section, the possible source of the ca. 280 BCE earthquake (?), has not shown any seismic activity for the last 2000 yr. (4) The northwestern section of the Pishvā fault (Figs. 4 and 8): This fault seems to lack historic seismic activity. (5) The western segment of the Moshā fault (Figs. 4 and 8): Although there have been three documented medium- to

Tehran: An earthquake time bomb large-magnitude earthquakes on the central segment of the Moshā fault (958, 1665, and 1830; Figs. 4, 8, and 17), no comparable seismic activity has been recorded along its western segment northwest of Tehran. A possible analogue earthquake for Tehran is the 20 June 1990 Mw 7.3 Rudbār earthquake (Figs. 5 and 9), which was itself located in a seismic gap (Berberian et al., 1992; Berberian and Walker, 2010). This was the first large-magnitude earthquake with 80 km left-lateral strike-slip coseismic surface faulting documented in the western “High-Alborz.” It was one of the largest and most destructive earthquakes during the instrumental period to have affected both urban and rural regions, killing ~40,000 people and making 500,000 people homeless. The earthquake destroyed four towns and damaged the provincial capital city of Rasht further north. We conclude that the 1990 earthquake is analogous to the 1830 Mw ~7.0–7.4 earthquake along the central Moshā fault segment, which caused some damage in Tehran (Fig. 17). A future earthquake could strike the unruptured western and eastern end of the Moshā fault as well as the already loaded North Tehran fault system and the inner-city faults (Fig. 17). Considering the entirety of the Alborz as similar in response to oblique-slip convergence because of uniform slip rate along strike (Figs. 6 and 9), the absence of earthquakes along the Rudbār fault is evidence that the 1990 Rudbār earthquake filled a seismic gap. This conclusion is strengthened by the poor activetectonic expression on the Rudbār source fault, as seen on pre1990 aerial photographs. Tehran, with its North Tehran fault system with poor geomorphic expression at the surface (especially along the thrust segments) and its inner-city faults, is similar to the pre-1990 Rudbār earthquake setting. However, unlike Rudbār, Istanbul, or Port-au-Prince, but as in Los Angeles, there is a web of several complicated interacting seismic faults in and around Tehran, adding significantly to the seismic hazard like ticking time bombs. Considering the Alborz earthquakes as analogous to the 1939– 1999 earthquake sequence on the North Anatolian plate boundary fault (Barka, 1996; although the westward progression was not as pronounced), we see from east to west, the 985, 1665, and 1830 CE earthquakes on the Moshā fault in the east, the 1177 earthquake possibly on the western segment of the North Tehran thrust, and the 1119 earthquake on the North Qazvin fault farther west (Fig. 17). Seismic gaps, therefore, would appear west and east of the 958– 1665–1830 earthquake cluster and the central and eastern section of the North Tehran fault system north and northeast of Tehran. 18. ALONG-STRIKE SEISMICITY MIGRATION AND INTERACTION OF FAULTS IN THE ALBORZ The long-term distribution of historical large earthquakes along the North Anatolian and East Anatolian plate-boundary faults where the slip rates are known (Ambraseys, 1970, 1971, 2009; Barka, 1996; Stein et al., 1997), as well as in Iran (Berberian, 2014), showed along-strike strain migration of large crustal earthquakes and interaction of active faults. Large sets of range-

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

137

parallel faults in the Alborz Mountains show local appearances of earthquakes along strike, though not as prominent as the pattern observed along the plate-boundary faults. The western Meydānak segment of the Moshā fault and the Tāléqān fault to its northwest and the Firuzkuh fault in the east (Fig. 9) are directly on strike with the Moshā and North Tehran left-lateral strike-slip fault system, and it would be reasonable to conclude that, at depth, these faults could represent the continuation of the Moshā fault system, and an earthquake on one segment or fault could load the other segments, including triggering of the North Tehran fault system bifurcating the Moshā fault (Figs. 4, 9, and 17). This would imply that the 958, 1666, and 1830 earthquakes on the Moshā fault (Table 2) are part of the same earthquake sequence that has loaded the North Tehran fault system branching from the Moshā fault. Continuing west, the 1608 Mw 7.3 earthquake on the Alamutrud fault and the 1119 Mw ~6.5 earthquake on the North Qazvin fault (Fig. 9) could represent the westward continuation of this earthquake sequence. This migration pattern is comparable to the westward migration of earthquakes on the North Anatolian fault (Barka, 1996). However, the Alamutrud fault dips south, opposite to the north dips on the Moshā, Maydānak (West Moshā), and North Tehran faults. Other patterns are also observed in the Alborz. For example, the 856 CE earthquake on the Dāmghān/Āstāneh fault system in the eastern Alborz might have triggered the 958 CE earthquake a decade later on the central Moshā fault in central Alborz to its west (Fig. 9). The 1119 CE earthquake on the North Qazvin fault in the western central Alborz was followed by the 1177 CE earthquake possibly on the western North Tehran fault with a 58 yr time interval (Fig. 9). After the 1405 CE earthquake on the Kelishom fault in the western Alborz, the seismicity migrated to its immediate west with the 1990 Rudār earthquake (Fig. 9). Similarly, a western migration of seismicity was detected along the Larzaneh fault from the 1301–1957 earthquakes in the northern central Alborz (Fig. 9). Other local examples are the 1127 and 1935 earthquakes on the North Alborz fault, the 1809 and 2004 earthquakes on the Khazar fault south of the Caspian Sea, and the 1896 earthquake on the Eshtehārd fault to the 1962 event on the Ipak fault in western Central Iran (Fig. 9). Unfortunately, the paleoseismological trench dates are not constrained enough to add the events to the historical seismicity records of the Tehran region. We might interpret intersections of two or three major active faults loading one another, which should be resolved by properly designed paleoseismicity study. We discuss these issues, since they have major implications for the earthquake hazard to the megacity of Tehran. 19. COSEISMIC NEARLY SIMULTANEOUS RUPTURING OF CROSS-STRIKE SUBPARALLEL FAULTS Our study shows that the Tehran metropolitan area is built on and surrounded by a web of densely populated, interacting,

138

Berberian and Yeats

subparallel seismically active faults, including the North Tehran fault system (thrust, with the Niāvarān, Darakeh, Farahzād, and Kan left-lateral strike-slip faults); the Moshā fault; the innercity reverse and strike-slip faults beneath the city (Mahmudieh, Dāvudieh, ‘Abbāsābād, Takht-e Tāvus, Bāgh-e Fayz, E-W, Nārmak, Shiyān, Telo, Sorkheh Hessār, and many more); and the Pārchin, Pishvā, and Palangvāz faults, among others (Figs. 2 and 4). All these faults are capable of generating medium- to largemagnitude earthquakes, as some fault did in the past. Despite the fact that the actual means of strain release in the Tehran metropolitan area are complex, there is the probability of a worst-case scenario assuming all strain across the metropolitan area is released during more than one earthquake on adjacent cross-faults. In this scenario, an earthquake on one fault loads the nearby cross-strike fault with nearly simultaneous rupturing of side-by-side parallel faults. During the three earthquakes of 14 August 1958 (two events of Mw 5.7 and 5.5) and 16 August 1958 (Mw 6.6), two parallel segments of the Zāgros Main Recent fault, 8 km apart, ruptured at the surface (Berberian, 2014). Detailed seismological and interferometric synthetic aperture radar (InSAR) study of the August 2014 Murmuri earthquake sequence in the Zāgros also showed cross-strike energy release during the 18 August 2014 Mw 6.2 (with its two aftershocks of Mw 5.7 and 5.4) and the Mw 6.0 (with its two aftershocks of Mw 5.6 and 5.8) on two different parallel cross-strike faults 15 km apart (Copley et al., 2015). Nearly simultaneous rupturing of side-by-side faults has also been documented elsewhere during the 1972 Managua, Nicaragua, earthquake (Sultan, 1931; Brown et al., 1973), and the 1987 Elmore Ranch, California (Kahle et al., 1988; Sharp et al., 1989; Hudnut et al., 1989), earthquake. 20. TEMPORAL EARTHQUAKE CLUSTERING We know that active faults do not show unique forms of seismic behavior, and, therefore, recurrence intervals of largemagnitude earthquakes along continental faults are complicated and highly variable in different structural provinces. The earthquake sequence of 958, 1665, and 1930 seems to have clustered in space and time along the central segment of the Moshā fault northeast of Tehran (Table 2; Fig. 9). Although three out of four recorded earthquakes that destroyed the city of Rhagae/Ray in the Tehran basin in ca. 280 BCE, 855–856 (?) 864 CE, and 1177 CE seem clustered in time, the lack of macroseismic data and the fact that they might have happened on different faults prevent us from defining the clustering in space. The 1890 and 1985 earthquakes in eastern Alborz, southeast of the Caspian Sea, show clustering in space and time (Fig. 9). Unlike the North Anatolian or Wasatch faults, the almost two millennia historic seismic record does not show evidence of clustering in the form of complete rupture of long multisegmented faults along the long faults in the Alborz Mountains (like the Khazar fault, or the Moshā fault north of Tehran; Fig. 9). It looks like a long period of seismic quiescence is followed

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

by an active cycle of earthquake clustering as observed along the Moshā fault. 21. DISCUSSION Unlike the simple plate-boundary faulting cases of Istanbul or Haiti, and possibly more similar to the Los Angeles, California, metropolitan area (Dolan et al., 1995, 2007; where the largest threat is still from the southern San Andreas fault, with threats from medium- to large-magnitude earthquakes on other reverse faults), at least eight different active faults in and adjacent to the sprawling Tehran metropolitan area (Fig. 4), acting like ticking time bombs, add significantly to the seismic hazard of greater Tehran in a complicated manner. Another analog for Tehran may be the Ordos Basin of Shanxi and Shaanxi Provinces, China, which showed clustering over three centuries (Yeats, 2012). Unfortunately, we do not have a good knowledge of faulting pattern, fault interaction, strain migration, seismic clustering, strain accumulation, and recurrence intervals along these faults. Study of Tehran faults (Fig. 4) and seismicity of the region (Figs. 5, 8, and 9; Appendices A through C) suggests that the city, with a rapidly growing population of ~15 million, is in a seismic gap. Faults to the north, northeast, northwest, west, and southeast have been activated during historical large-magnitude earthquakes (Fig. 9). The only earthquake in modern times in Tehran struck the Moshā fault located ~30 km north and northeast of the city on 27 March 1830 with an earthquake of Mw ~7.0–7.4. Other active faults include the North Tehran fault system and inner-city faults. No large historical earthquakes have yet been assigned to the Niāvarān, Darakeh, Farahzād, and Kan left-lateral strikeslip faults, or the inner city Mahmudieh, Dāvudieh, ‘Abbāsābād, and Bāgh-e Fayz reverse faults underneath the metropolitan Tehran area (Figs. 2, 4, and 9). The inner-city active Pliocene– Pleistocene and younger alluvial deposits, with evidence for asymmetric folding and escarpments, may also indicate folding above blind thrusts dipping north, which have not yet been investigated properly. Major cities south and north of the Alborz Mountains are located on or adjacent to active faults and have suffered major earthquakes during historical times (Fig. 9; Appendices B and C). Future earthquakes on these faults could cause horrific loss of life and destruction far exceeding losses in earlier earthquakes because of the explosive growth of population in recent years. Unfortunately, earthquake preparedness, including enforcement of building codes, retrofitting of major public buildings and infrastructure, and earthquake insurance has not kept up with the recognition of the hazard since 1951 (Abdalian, 1951) and stressed in 1974 and 1985 (Tchalenko et al., 1974a; Berberian et al., 1985). Tehran adopted a national earthquake resistant building code in 1969 (the First Iranian Code for Seismic Resistant Design of Building; ISIRI Code No. 519), and increased seismic standards, incorporating seismic and fault data, in 1988 and 1999 (the Second Iranian Code and its revision for Seismic Resistant

Tehran: An earthquake time bomb Design of Building; ISIRI Code No. 2800). The Iranian building seismic code is applied only to new construction (but has never been properly enforced); it does not require seismic retrofitting of existing public buildings, and it does not require disclosure of earthquake resistance of structures and building sites during property transactions. The code suggests an acceleration value of 0.35g for the entire Tehran region, which is slightly higher than the first values estimated at 0.31g in Berberian et al. (1985). Furthermore, corruption in the construction industry, among the city building inspectors, and in the municipalities, and the country at large, as documented by Transparency International (transparency.org, 2012) and reported by Bilham (2009), Ambraseys and Bilham (2011), and Berberian (2014), has resulted in poor construction practices throughout a country rich in petroleum, natural gas, copper, and other mineral resources. Since adoption of the revised building codes in 1988 and 1999, and despite warnings given in scientific reports since 1976, numerous buildings have been constructed on top of active surface and blind faults in major Iranian cities, including the Tehran metropolitan area. Rapid industrial growth and expansion in Iranian cities have created a very strong employment market and consequently an attraction pole for migration from rural areas and small towns to urban centers (Diba, 1980). The 1979 change of regime followed by the devastating Iran-Iraq War of 1980–1988 caused an uncontrolled urban housing boom in Tehran. Squatter housing developed on the fringes of urban centers, especially in Tehran, and the density of population in city slums increased rapidly (Diba, 1980). The growing population in earthquake-prone regions suggests that future earthquakes will become increasingly deadly, since earthquake fatalities are proportional to the exposed population and their vulnerability to earthquakes. As in other major cities of the world (Bendimerad and Comfort, 2002; Bilham, 2009; Ambraseys and Bilham, 2011; Holzer and Savage, 2013; Bilham and Gaur, 2013; Yeats, 2015), rapid urbanization in Tehran has been pushing the earthquake risk curve higher and higher, and no steps for risk mitigation have been taken by authorities in Tehran. The city of Tehran has expanded, while structures and infrastructure are constructed with little consideration to earthquake resistance; hence, 15 million people in the connected Tehran-Karaj metropolitan area live in buildings vulnerable to medium-magnitude earthquakes that endanger their lives and properties as well as the Iranian economy. Furthermore, Iran is ruled by a centralized government, with all governmental offices and emergency facilities and boards concentrated in Tehran, which would be severely affected, even in the case of a medium- to large-magnitude earthquake. Earthquake damage estimates conducted during a seismic microzoning study of Tehran (JICA, 2000) indicated that in the case of a major earthquake on the North Tehran fault system, 19.5% to 61.6% of residential buildings in Tehran would collapse (the range is based on distance from the faults), with human casualties of 126,000 (nighttime scenario) to 80,000 (day) out of 6.3 million population (JICA, 2000). The damage ratio due to activation of the distant Moshā fault was estimated as ranging

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

139

from 6.8% to 17.9% of residential buildings, with a death toll of 20,000 (night) to 12,000 (day). A floating seismic source (previously unrecognized fault) was estimated to cause an average damage ratio of 50.9% of the residential buildings, with a death toll of 301,500 (night) to 176,300 (day) out of the 6.3 million population at the time of study (JICA, 2000), when the population was almost half of the current population. Therefore, based on the current population, the death toll estimation of 126,000, would increase to ~252,000 in case of the North Tehran fault system reactivation and ~40,000 for a Moshā fault earthquake. Secondary effects such as liquefaction in southern Tehran and Ray (although with low potential due to the high clay content of sediments), and rock avalanche and landslides in northern Tehran (rated high at the Alborz Mountain front) were not considered in these estimates. Considering the 10 million population in Greater Tehran around 2000, expected casualties and injuries during an earthquake in Tehran were estimated at precise numbers of 1,446,600 dead and 4,339,900 injured (Nateghi-A., 2001). A comparison of worldwide earthquake building damage and human casualties (Coburn and Spence, 1992; JICA, 2000) shows that the losses in Tehran would be much higher than the 1985 Mw 8.3 Mexico City (Mexico), 1988 Mw 6.8 Spitāk (Armenia), 1990 Mw 7.3 Rudbār (Iran), 1995 Mw 6.8 Kobe (Japan), 1999 Mw 7.3 Chi-chi (Taiwan), 1999 Mw 7.6 Izmit (Turkey), 2003 Mw 6.5 Bam (Iran), 2008 Mw 7.9 Wenchuan (Sichuan, China), and 2010 Mw 7.0 Haiti earthquakes (though Mexico City and Kobe were much better prepared for the earthquakes there, and their losses were, therefore, low, whereas Iran, Armenia, Turkey, and China were not prepared). Equating earthquake casualties per event and earthquake magnitude (Hough and Bilham, 2006; Bilham, 2009, 2010; Berberian, 2014), the estimated figures in Tehran would be much larger than those of the 1923 Mw 7.9 Tokyo (Japan; a plate-boundary event with larger magnitude, >100,000), 1976 Mw 7.8 Tangshan (China, 240,000–500,000), 1978 Mw 7.3 Tabas-e Golshan (Iran, 20,000), 1990 Mw 7.3 Rudbār (Iran, ~40,000), 1988 Mw 6.8 Spitāk (Armenia, 25,000), 1999 Mw 7.4 Izmit (Turkey, 17,000), 2003 Mw 6.6 Bam (Iran, 32,000–43,000), 2005 Mw 7.6 Kashmir (>78,000), 2008 Mw 7.9 Wenchuan (China, ~87,000), and 2010 Mw 7.0 Haiti (~160,000) earthquakes (see figure 3 in Berberian, 2014). Despite (1) the repeated occurrence of earthquake disasters since 1900 in Iran (with destruction of the cities of Dorud, Rāvar, Bastak, Salmās, Māhān, Torud, Lār, Kākhk, Ferdows, Tabas-e Golshan, Rudbār, Manjil, Lowshān, Harzévil, and Bam, with death tolls totaling 164,000; Berberian, 2014); (2) a remarkably well-documented historical record of destructive earthquakes; (3) a remarkable indifference to the potential loss of life from earthquakes; and (4) a prolonged country-wide drought and severe drop in the groundwater tables throughout the country, in July 2012, authorities in Iran, following the order of the supreme religious leader Āyatollāh Khāmene’i, ended the controlled growth policies and scrapped the family planning and birth control programs and stated that Iran should aim for a population of 150–200 million in the near future (Iranian media news; telegraph.co.uk;

140

Berberian and Yeats

ibtimes.com, July 2012). Consequently, on 4 May 2014, Āyatollāh Seyyed Mohammad Hossein Ghazvini, director of the International Velāyāt (Guardianship) TV Network (“Islamic Governance of the Jurist”; with offices in Qom and the United States) launched “Operation 14 Children” and appealed to all Iranian Shi’as to take part in reversing Iran’s dwindling birth rate, calling for the faithful to have “14 children,” beginning “tonight,” consistent with the tradition of “14 immaculate Shi’a Saints” (Iranian news media; nooreaseman.com/forum354/thread65563. html; youtube.com/watch?v=fQcEYIBNAiY; iranwire.com/fa/ news/494/5518, accessed May 2014). All the existing data show that the earthquake fatality rate keeps pace with rising populations (Bilham, 2009; Holzer and Savage, 2013; Yeats, 2015). For the twentieth century, however, Iran, because of the large number of fatal earthquakes, the association between rising populations and increasing fatalities from earthquakes is clear (see figures 1 and 2 in Berberian, 2014). The ratio of earthquake-related fatalities in Iran to Iran’s growing twentieth-century population since 1900 indicates the two growth curves track each other. The number of deaths from earthquakes in general shows little evidence of decline (Bilham, 2009; Berberian, 2005, 2014). This clearly indicates that the approval of the codes for earthquake-resistant structures in Iran since 1969 has had little effect in reducing earthquake fatalities due to lack of building code enforcement by the authorities; the 2003 mediummagnitude Bam disaster is the latest proof. Our detailed analyses of the pre-1900 and instrumental-era earthquakes (Appendices A through C) allowed us to degrade exaggerated magnitude values of the historical earthquakes, as well as delete spurious events from the existing catalogues (Table 2). Unfortunately, the exaggerated magnitudes and dubious historical events have entered the national and global catalogues and were adopted in numerous recent publications and in probabilistic seismic hazard assessments. In most probabilistic hazard calculations for the city of Tehran (as well as the country at large), unrealistic scenarios with inputs such as: (1) very large magnitudes (7.5 < M < 8.1) at a very short epicentral distance from the city center, (2) inaccurate epicenters, and (3) vague and unrealistic boundaries of the seismic zones or seismotectonic provinces were utilized. This approach, without scrutinizing the seismic sources, has resulted in overestimation of the earthquake hazard in ground acceleration and velocity assessments for Tehran (Tavakoli, 1996; Ahmadi and Nowroozi, 1981; Bozorgnia and Mohajer-Ashjai, 1982; Berberian et al., 1985; Nowroozi and Ahmadi, 1986; Ahmadi et al., 1989; Mirzaei et al., 1997, 1998; Tavakoli and Ghafory-Ashtiany, 1999; Ghodrati Amiri et al., 2003, 2004, 2012; Ashtari Jafari, 2007; Zafarani et al., 2009, 2013; Tsang et al., 2011; Yazdani and Abdi, 2011; Ghayamghamian and Gheisari, 2013; Samaei et al., 2014). Our study, with a better characterization of seismic sources, magnitudes of events, and their locations, will enable future hazard calculations to be more realistic. In characterizing seismic sources, we selected the parameters of the 1990 Mw 7.3 Rudbār earthquake as MCE for Tehran. The

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

earthquake far exceeded previous destructive earthquakes in Iran, both in the large death toll and unprecedented destruction and damage to buildings and property along a meizoseismal area of 110 × 40 km, with a coseismic surface fault length of 80 km. Two issues should be noted: (1) The Rudbār earthquake took place on a left-lateral strike-slip fault in the Alborz; and (2) the majority of buildings destroyed during the 1990 earthquake were traditional unreinforced rural rubble or adobe masonry structures with horizontal timbers in the ceiling, and modern buildings with kiln brick with horizontal reinforced concrete ring beams in the four destroyed cities and towns. As with the Rudbār earthquake fault, the central-eastern section of the North Tehran fault system, the Niāvarān, Darakeh, Farahzād, and Kan faults (Fig. 2), have a predominantly left-lateral strike-slip mechanism, whereas the western segment of the fault system is a thrust fault (Berberian et al., 1985). Furthermore, other surface and blind reverse faults occur in inner Tehran, which, as a substitute to the North Tehran fault system, may be active (Fig. 2). It should be noted that the hazard to Port-au-Prince, Haiti, was expected to be on the Enriquillo strike-slip fault, whereas the 12 January 2010 Mw 7.0 earthquake rupture, which destroyed the city, occurred on a blind reverse fault (Hayes et al., 2010). As with other megacities in industrialized developed countries, authorities in Tehran should work to minimize the earthquake vulnerability and avoid higher death tolls and greater economic losses (Bendimerad and Comfort, 2002; Berberian, 2014); this can be achieved by strong enforcement of building codes and education of the authorities and population against earthquakes, as is apparently being done in Istanbul. The authorities in earthquake-prone countries should be legally and morally accountable for disasters. 22. CONCLUDING REMARKS This paper is an extensive review and reevaluation of substantial geological, seismological, historical, and archaeological work carried out in the Tehran region since the early 1970s and published in peer-reviewed and local literature. In our quest to reach a conclusive spatiotemporal earthquake and faulting history of the region, we have purged and modified the existing historic earthquake catalogue. We also found that the Tehran region is located in a seismic gap comparable to the pre-1990 Mw 7.3 Rudbār earthquake in the northwest Alborz Mountains. Our review shows that the seismic parameters of the Rudbār earthquake can be regarded as maximum expected earthquake (Mw ~7.3–7.4) for the Tehran region and reject recommended magnitudes of Mw 7.7–8.1 for Tehran and the Alborz mentioned in the literature. An analogue to the Los Angeles metropolitan region, southern California, metropolitan Tehran is located within a crowded active strike-slip and reverse fault system. Both areas have major reverse and strike-slip faults and have had damaging earthquakes in the past (1971 Mw 6.5 and 1992 Mw 7.3 in southern California, and 1830 Mw ~7.0 in northern Tehran). The metropolitan Tehran is located in a geologically complex area at the junction

Tehran: An earthquake time bomb between the south-central Alborz Mountains and the Central Iran seismotectonic provinces, with well documented historical earthquakes (Figs. 2, 4, and 17; Appendices A to C). Because of the close proximity and size of at least eight adjacent and inner-city active faults in the Tehran metropolitan area and their combined potential hazards to the city, these faults are of particular concern and threaten ~15 million people by large-magnitude earthquakes causing a catastrophe with great economic loss and very hard and long recovery period for the survivors and the state (Figs. 2, 4, 9, and 17). However, unlike the Los Angeles metropolitan area, with ~8.5 mm/yr slip rate, the regional GPS data indicate ~1.5 and 1.8 mm/yr left-lateral and north-south–shortening slip rates, respectively, in the metropolitan Tehran region (Mousavi et al., 2013). Our study demonstrates that the 958 Ruyān, 1665 Damāvand, and 1830 Lavāsānāt medium-to large-magnitude earthquakes along the central segment of the Moshā fault, located ~30 km northeast of Tehran, where the North Tehran fault system branched off, might have loaded the North Tehran fault system, the innercity faults, and other adjacent faults to the east and south toward the seismic gap of the cities of Tehran and Karaj (Figs. 4, 8, 9, and 17). The available macroseismic data from the two Mw ~7.0 earthquakes of 958 and 1830 along the central segment of the distant Moshā fault suggest a possible interval (recurrence period?) of 872 yr; no information is available on the seismicity of the western and eastern segments of the fault. Our interdisciplinary approach and scrutinized historical seismic data indicate that, unlike Los Angeles, no medium- to large-magnitude earthquake has occurred along the faults in the immediate vicinity or the inner-city faults in Tehran for the past 838 yr, which is consistent with the low slip rates derived from geodetic studies. As with the 2003 Mw 6.6 Bam earthquakes, with 32,000–43,000 death toll in a small desert town in southeast Iran, we may conclude that a moderate-magnitude earthquake on one of the metropolitan Tehran faults could potentially cause more damage to the metropolitan area than a large-magnitude earthquake on the distant Moshā fault (Fig. 17), unless the entire length of the Moshā fault ruptures, creating a larger earthquake, which has not been documented since at least the eighth century CE. Large-magnitude surface displacements, as well as strong ground motions on adjacent faults (the North Tehran system, Pārchin, Pishvā, and other faults with possibility of nearly simultaneous reactivation of the inner-city faults), have the potential for complete disruption of lifeline systems and search-and-rescue efforts, and the substantial destruction of buildings in the metropolitan area. Ideally, paleoseismologic data in the form of dates and magnitudes of past earthquakes should provide valuable data on the cycles of stress buildup and release, fault displacements, and slip rates, and, together with historical seismic data, increase our knowledge and ability to correctly forecast the seismic parameters and locations of future earthquakes. Unfortunately, the few paleoseismological data are too sparse and incomplete, with poorly constrained rupture dates, so we are unaware of the exact dates and seismic parameters of the most recent large-magnitude

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

141

earthquake ruptures along the eight faults in the metropolitan Tehran area. Nonetheless, based on more reliable historical seismic data, we know that these faults are capable of generating earthquakes in the range of Mw ~6.5–7.4 with large recurrence periods on individual faults, earthquake clustering, seismic gaps, along-strike migration of seismicity, and interaction along multiple parallel faults. We see some kind of alternating seismic activity, where each large-magnitude earthquake along one fault increases the probability of activity on an adjacent fault. It is tempting to devise a space-time model of activity for the Tehran region showing that we may possibly be close to the end of the current 838 yr quiescence for a medium- to large-magnitude earthquake on some of the metropolitan eight faults. However, without studying the long-term behavior of faults by properly designed additional detailed paleoseismologic trench studies, it is difficult to confidently postulate such a scenario. We have scrutinized the past regional large-magnitude earthquakes, addressed the reliability of the existing earthquake catalogues, highlighted the unconstrained nature of the current paleoseismic trench rupture dates, and mapped the faults in this region that we suggest are capable of generating damaging earthquakes affecting Tehran, some with the potential for huge losses of property and lives in the city. Since the Tehran region, with its rapidly growing urban area and numerous seismic sources, is continuously accumulating strain, we emphasize the earthquake and fault hazard and risks and call for an immediate action by the authorities. Although modern building codes exist for Tehran, the potential in this megacity for corruption in the construction industry and among building inspectors (common as well in the Middle East and other developing earthquake-prone countries) indicates the need for a major program of strengthening buildings and lifelines against the next, inevitable earthquake. The past 2000 yr of seismic history of the Ray-Tehran region, and Iran in general, reveal that the authorities have been remarkably indifferent to and ignorant of the potential loss of life and property from earthquakes. The earthquakeresistant building codes have not been enforced, unreinforced public structures and infrastructure have not been retrofitted, and recent houses in Tehran have been built upon seismically active faults. The hazardous location of Tehran, with its vulnerable structures and infrastructure, has not been dealt with to avoid strong ground shaking. In order to reduce earthquake vulnerability and improve the sustainability of Tehran, a detailed Earthquake Management Master Plan and Information System is recommended to be designed and implemented. The master plan should include disaster assessment and management technology, a disaster mitigation plan, a disaster preparedness plan, and a disaster response and recovery plan. A culture of learning from past earthquake disasters in Iran and the world at large, prevention, safer construction and urban planning, urban land use, public policy action and education, disaster accountability, and community participation should be established and become habitual in the disaster mitigation plan of Tehran and the country. The building code should be enforced, and the

142

Berberian and Yeats

public buildings and infrastructure should be retrofitted by the authorities in the two most populous provinces of Tehran and Alborz (Karaj, west of Tehran), covering an area 230 km long along the southern faulted range front of the Alborz Mountains, with a total population of 14.9 million, representing 20% of the total population of Iran in 2011. ACKNOWLEDGMENTS This paper was presented on 29 October 2013, during the 125th Anniversary Annual Meeting of the Geological Society of America (GSA) in Denver, Colorado, T188—Sessons 214 and 291, “Tethyan Evolution and Seismotectonics of Southwest Asia: In Honor of 40 Years of Manuel Berberian’s Research Contributions.” We thank GSA and the session-chair, Rasoul Sorkhābi, for organizing these special sessions, and Rasoul Sorkhābi for his warm hospitality in Denver, Colorado. Comments and corrections by Angela Landgraf, John Schroeder Jr., Eldon Gath, an anonymous reviewer, and Rasoul Sorkhābi, which contributed to improving the manuscript, are greatly appreciated. We greatly acknowledge Howard Schwartz, Håkan Wahlquist, and Anne Murray (the Sven Hedin Foundation, Stockholm; http://svenhedinfoundation.org) for the Tehran photograph taken by Sven Hedin in 1906 (Fig. 23); R.Q.M. Tābān for the 2013 photograph of north Tehran (Fig. 3); and Kāmrān Ansāri, Faramarz Golbon (Tehran), Sādeq Shajari (Paris), Karim Alizadeh (Harvard), and Prince M.H. (Mickey) Kadjar (Qājār, Dallas, Texas) for their help in tracking down some old publications. Special thanks go to the GSA publishing department for their efforts during the final editing and printing processes. This paper is dedicated to the pioneering work of professor Setrāk Ābdāliān (1894–1963; professor at the University of Yerevān, Armenia, and University of Tehran, Iran) on the earthquake hazard in Iran, especially the city of Tehran, in 1951, prior to the establishment of the Institute of Geophysics of the University of Tehran in 1957 and the Geological Survey of Iran in 1962 (Abdalian, 1951); and for his groundbreaking field investigation during the time of the petro-political and economic turmoil of 1953 (Abrahamian, 2013), when he studied the coseismic surface deformations of the 12 February 1953 Mw 6.5 Torud earthquake (Abdalian, 1953). The sustained research on this immeasurable project (which was initiated 45 yr ago in 1971 in Tehran by John Tchalenko, Nicholas Ambraseys, and Manuel Berberian; during the year that Manuel Berberian single-handedly established the first Research Department of Tectonics and Seismotectonics in the country) was not supported by any grant, organization, or individual. We hope that the authorities in Iran, and other developing countries, take the earthquake hazards seriously, to create a culture of prevention and develop solutions to disasters during the present generation, and do not hand the chronic problem over to the next generations. We expect that

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

this work will inspire renewed efforts to mitigate impacts from future earthquakes. APPENDIX A. COMMENTARY ON PREHISTORIC EARTHQUAKES OF THE GREATER TEHRAN-RAY REGION DISCUSSED IN THIS PAPER The geological and cultural evidence of earthquakes in the Greater Tehran-Ray area is mentioned here for further investigation. The prehistoric paleoseismic trench events are summarized in Tables 4–8 and 12–14 and addressed in the text. Ca. 24,000–14,000 B.P. Lāsem Lake (Damāvand) Earthquake-Induced Deformational Structures (Lāsem Lake Deposit Seismites) Synsedimentary soft-sediment deformations characterized by recumbent folds developed in flat-lying lake deposits at Lāsem (35.82°N, 52.11°E) south of the Damāvand volcano and 8 km north of the Moshā fault (Fig. 12) are possibly one of the earliest geological records of seismic activity in the Alborz Mountains, northeast of Tehran (see plate 8.1, p. 220, in Berberian et al., 1985). The soft-sediment deformational structures, located 8 km north of the Moshā fault, are overlain and underlain by horizontal layers. The formation of the structures has been attributed to a paleoearthquake; possibly, they formed by strong ground shaking and not by sedimentary processes (Berberian et al., 1985; Berberian, 1994). The deposits, in which the structures developed, are related to filling of an ancient lake dammed by lava flows of Damāvand volcano, which dammed the valley south of the present Lār-Dalichāi confluence to the southwest of Damāvand volcano (the Lār paleolake), the Lāsem Valley to the south of Damāvand volcano, and other nearby valleys (Fig. 12). A radiocarbon date of ca. 38,500 yr B.P. was reported from lake terraces formed as a consequence of damming the Lār Valley by lava flows from the southwestern flank of the Damāvand volcano (Bout and Derruau, 1961; Allenbach, 1966). The Lār Valley River was, therefore, dammed by 40–38 ka Damāvand lava flows (Allenbach, 1966; Davidson et al., 2004). The 40Ar/39Ar and apatite (U-Th)/He ages suggest periods of significant volcanic activity ca. 1.2 Ma–800 ka, ca. 280 ka–150 ka, and ca. 60 ka–7 ka (Davidson et al., 2004). The lava flows dated ca. 25 ka to the southeast of the volcano shifted the Harāz River further to the east in the Āb-e Ask area (Davidson et al., 2004). A comparison of the synsedimentary structures observed at Lāsem with the earthquake-induced sedimentary structures in the sediments of the Van Norman Dam reservoir (Sims, 1973, 1975), 12 km from the epicenter of the 9 February 1971 Mw 6.7 San Fernando Valley, California, earthquake, may indicate that a high-intensity (I > ~VIII MMI?) earthquake took place in the Damāvand region of the Alborz Mountains, 8 km north of the Moshā fault (Fig. 12). However, with a single observation, we cannot estimate the magnitude of the event, if any. Ca. 1200 BCE Ragheh (Rhagae; Modern Ray) Earthquake Myth The two cities of Rhagae (35°36′N, 51°26′E; Figs. 2 and 13), the ancient religious capital of Media (Māda/Mād), and Hecatompylus (Komesh) to the east (Fig. 15), were located on the ancient trade route of the “High Road,” connecting Mesopotamia in the southwest, through the Zāgros Mountains, and along the southern Alborz to the East. The city of Rhagae (Ragheh) in southern Tehran was known as the twelfth city in the world and the second holy city in ancient Iran created by Ahurā Mazdā, the Wise Lord, in the Zoroastrian holy

Tehran: An earthquake time bomb book of Avestā (Vendidād, 1:15; Yasnā 19:18); Doghdu, mother of Zoroaster, was born in this city around 1200 BCE (Boyce, 1989). The name of the city changed to: (1) Raga during the Achaemenids and is mentioned by Darius I in the Behistun (Bisotun [sic]) inscription (2nd column:16; 522–486 BCE; see also Nyberg, 1974); (2) Rhagae around 300 BCE (Arrian: III.9); (3) Europos during the Seleucids occupation of 312–174 BCE (Strabo: XI.9); (4) Arsakia during the Parthians (Weissbach, 1895); and (5) Ray during the Sāssānids (Kariman, 1970; Rante, 2007, 2014; Berberian, 2014), which continues to the present day. In Book 7: Marvels of Zoroastrianism, Chapter 2 (Parentage of Zartosht) of Denkard (§44, 45), the storm-god/monster Chashmag-e Div undertakes to destroy the house of Zoroaster’s father Puroshasp (Pourush-aspa/Pourushaspa in Avestan; Pourushap in Pahlavi; lit. “owner of old horse”) as well as the village, and brings typhoons (ed. West, 1897; Bahār, 1966, 1983, 1990; Berberian, 1991, 2014; avesta. org/Denkard (tr. 1876). The myth of destruction of the house of Zoroaster’s father by the storm-god/monster could have originated from an ancient large-magnitude earthquake. Airyanem-Vaedja, the motherland of Zoroaster (Zarathushtra, Zartosht), was possibly situated in the area north of Khorāsān in the Khārazm Province (south of the Lake Khārazm/Ārāl) of the Greater Khorāsān around 1200 BCE (Boyce, 1989). However, since Doghdu (Puroshasp’s wife and Zoroaster’s mother) was born and raised in Ragheh (Rhagae, Raga, modern Ray), it is practical that they married at Ragheh (Fig. 12). Therefore, the earthquake, if genuine, might have possibly happened in present-day Ray. This could be a record of a possibly major devastating earthquake in Ray, far back in the misty past, the memory of which was kept in the Zoroastrian annals and their old folklore, and legends (Berberian, 2014). It is not clear if the 0–4 ka paleo-event at the Vardāvard trench across the North Tehran thrust (Nazari, 2006; Nazari et al., 2011; Ritz et al., 2012) could be associated with the Ragheh earthquake myth, which might have taken place ca. 3.2 ka (Tables 4 and 6; for dating the time of Zoroaster, see Boyce, 1989). APPENDIX B. COMMENTARY ON HISTORICAL (PRE-1900) EARTHQUAKES OF THE GREATER TEHRAN-RAY REGION DISCUSSED IN THIS PAPER, WITH THEIR ESTIMATED MAGNITUDES AND MEIZOSEISMAL AREAS Historical seismic data are inhomogeneous and incomplete, and numerous medium-magnitude earthquakes that occurred far from the main cities have not necessarily been recorded properly. Erroneous data have entered the existing catalogues, and some seismic parameters are in error. In this commentary on the regional earthquake catalogue, we have examined the preserved contemporary and near-contemporary chronicles in Persian, Arabic, Syriac, Byzantine, Armenian, and Turkic to evaluate the historical seismic data of the study area, citing complete references to the sources of each event. The reality of dealing with historical seismic data in this part of the world is that, in most cases, the contemporary incomplete and uncritical chronicles providing collective brief yearly seismic events of the region or the country are incomplete and uncritical. In this case, there is a danger of treating the collective yearly events as a single earthquake, as has been reported by many contemporary and secondary and tertiary sources. Based on our historical knowledge of the area, the time of earthquakes and population density of cities, districts, and villages, we concluded that the casualty figures reported by the chroniclers have been unreasonably exaggerated. We also reiterate that the chroniclers very rarely identified the names of villages and small towns destroyed during earthquakes. They usually gave the names of province, districts, or the center of districts, which usually carry the same attribute, such as the city of Ray and

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

143

the district of Ray. The latter covered an area of ~70 km in the E-W direction (between the Jājrud and Karaj Rivers) and 70 km in the N-S direction (from Varāmin to Fasham), with the city of Ray located approximately at its center (Fig. 1). Another example is denoting the town and the province of Komesh, which makes it complicated to locate the damaged areas. According to many Persian and Arab writers, the city of Ray was devastated five times in 200 yr between the eighth and the tenth century CE. After analyzing the contemporary sources, we reached the conclusion that during this short time period, only three earthquakes took place on different faults. We also noted that historical large-magnitude earthquakes have been discontinuous, clustered in space and time, and interspersed with long periods of relative quiescence. In the following sections, volume and page numbers of some available chronicles are shown as II.491 (i.e., volume II, page 491). Ca. 280 (312–280) BCE Rhagae (Modern Ray) Earthquake According to near-contemporary sources, Duris of Samos (ca. 350–after 281 BCE), “Rhagae, in Media, has received its name4 because the Earth about the Caspian Gates [Caspiae Portae; Tang-e Sardarreh/Sardarreh Defile, 35°36′N, 51°26′E, Fig. 18; Hansman, 1968; Standish, 1970] had been rent by earthquakes to such an extent that many cities and villages were destroyed, and rivers underwent changes of various kind” (Strabo: I.3.19; Ambraseys, 1974). Later, Poseidonius of Apameia (second century BCE) wrote that during this earthquake, “many cities and 2,000 villages were destroyed in the district of Rhagae,” implying that the city of Rhagae itself was possibly not involved in ground deformation (Strabo: XI.9.1; Nabavi, 1972; Ambraseys, 1974; Berberian et al., 1983; Berberian, 1994, 2014). Almost two centuries after the event, Apollodorus of Artemia (mid–first century BCE) mentioned that the city of Rhagae (35°36′N, 51°26′E) was re-founded by Seleucus Nicator sometime between 312 and 280 BCE, and he named the newly built city “Europus,” but it was called “Arsacia” by Parthians (Strabo: XI.13.6; Kariman, 1970, I.76; Ambraseys, 1974; Rante, 2007, 2014). Although he did not mention the reason for its reconstruction, he located it ~500 stadia (55 miles [88.5 km]; the actual distance is ~44.7 miles [72 km]) from the Caspian Gates, i.e., Sardarreh Defile at 35°17′N, 52°08′E, +966 m (Figs. 17 and 18; Table 2; Standish, 1970; Hansman, 1968; Anderson, 1928; Ambraseys, 1974; Bosworth, 1983; Berberian, 2014). No other source records this earthquake. Ambraseys and Melville (1982), followed by Gorshkov et al. (2009) and many others, assigned a magnitude of Ms 7.6+ for this event, with its epicenter at 35.50°N 51.50°E. The basis for assigning a very large magnitude of Ms >7.6 for this event is not clear. The names of the destroyed villages and the source fault of the event are not known; hence, the assigned magnitude cannot be warranted without a proper paleoseismic trench study (Berberian, 2014). This would make it necessary to reject the reported magnitude of Ms 7.6+ (Table 2) and its source on the Garmsār fault in De Martini et al. (1998), Berberian and Yeats (1999, 2001), and Ashoori (2004). The Ms 7.6 magnitude for this event was adopted as a source input in the probabilistic seismic risk assessment of Tehran by Ghodrati Amiri et al. (2003, 2004, 2012). The number of 2000 destroyed villages reported by Greek historians in the second century BCE seems highly exaggerated and unreal for the period between 312 BCE and 280 BCE around Rhagae in Iran (even since 1900, there have never been 2000 villages in the district or county of Ray). Nonetheless, recording of this event by the Greeks, the survival of the information about it over such a long period, and

4

The claim is not correct; see Berberian (2014, p. 189–190).

144

Berberian and Yeats

widespread destruction with topographic changes may suggest that the earthquake was a large-magnitude event (Mw ~ ≥7.0) that took place during the late Achaemenid–early Hellenic occupation time with surface rupturing possibly between Rhagae and the Caspian Gates (Sardarreh Defile; Fig. 18). Sardarreh Defile is located on the hanging-wall block of the Garmsār thrust at the southeastern end of the Pārchin fault (Fig. 18), and the city of Rhagae is positioned at the northwestern end of the Pārchin fault (Berberian et al., 1985; Berberian and Yeats, 2001). It seems that reactivation of the Pārchin fault can destroy the area from the city of Rhagae (Ray) to the Caspian Gates (Sardarreh Defile), and it is probable that it ruptured during this event (Fig. 18; Berberian, 2014). Paleoseismic trench study is needed to prove this assumption. Nazari (2006), Nazari et al. (2007, 2008, 2011), and Ritz et al. (2012) assumed that their penultimate event (E2: 4000–2900 yr B.P., Mw 6.6–7.1) at the Vardāvard trench (Fig. 4) west of Tehran in the North Tehran fault zone could correspond to the 312–280 BCE earthquake at Ray. Later, Nazari et al. (2009, p. 1033) claimed that their second Tāléqān fault event (Ev.2; dated 3470–1540 yr B.P.; Fig. 4) could be associated with the 312–280 BCE earthquake, and that the 312–280 BCE earthquake was “misallocated in the historical catalogue.” These claims—that a single earthquake destroying Ray and the Sardarreh Defile (Fig. 18), occurring on two different faults far from the area mentioned by the near-contemporary sources—cannot be warranted. Tatar et al. (2015) considered that the 312–280 BCE earthquake took place along the Garmsār fault. Archaeoseismic Data Numerous archaeological mounds exist on the Tehran-Ray plain (Fig. 18), and it is vital for these sites to be reviewed and studied in detail for archaeoseismic indicators. These sites are surrounded by numerous active faults, and some of them show deep-seated, nearly vertical fractures on the trench walls. For example, figure 3 in Fazeli et al. (2004) shows the trench E4–5 wall in the background with numerous fractures cutting a 5 m section of “Transitional Chalcolithic I and II” (ca. 5300–4300 BCE) and Early Chalcolithic (ca. 4300–4000 BCE) archaeological stratigraphic layers. Archeological excavations at Cheshmeh ‘Ali mound, the ancient Sureni/Surini Spring and River (Fig. 18), named after the Parthian General Surenā (Abu Dulaf, 950; Ibn Hawqal, 978; Istakhri, 951; Yaqut, 1226; Kariman, 1970) at ancient Rhagae, indicate that Rhagae was occupied during the Late Neolithic to Late Chalcolithic (ca. 5500–3000 BCE) periods (Schmidt, 1935a, 1935b, 1936a, 1936b, 1936c; Voigt and Dyson, 1992; Matney, 1995; Mousavi, 2001; Fazeli et al., 2001, 2002, 2004; Coningham et al., 2004; Rante, 2010, 2014). After an occupation lapse of 3000 yr after the Neolithic occupation, the Sureni (Cheshmeh ‘Ali) site (Fig. 18) was occupied during the Pārthian period (250 BCE–224 CE), as shown by coins dating from 171 BCE to CE 12, as well as the Sāssānian period of 227–642 CE (Schmidt, 1935a, 1935b, 1936a, 1936b, 1936c; Rante, 2008, 2014). Later, when the name of the Sureni site was changed to Cheshmeh ‘Ali by the Arab invaders after 642 CE, it was occupied during Umayyad (660–750 CE), ‘Abbāsid (750–821), Buyid (945–1055 CE), and Saljuq (Seljuk; 1000–1218 CE) periods (Schmidt, 1935a, 1935b, 1936a, 1936b, 1936c; Miles, 1938; Keall, 1979; Treptow, 2007). Apparently, after a gap at Sureni/Cheshmeh ‘Ali, the Ray Citadel mound in the ancient fortified city of Ray (composed of the Citadel in the north and the large Shahrestān or inner city to its south; Figs. 13 and 18), located ~300 m southeast of Cheshmeh ‘Ali (Sureni) on the nose of the Sorsoreh Mountain, was occupied in the third millennium BCE, and the Early Iron Age (ca. 1000 BCE), the early Pārthian (250 BCE–CE 224), Sāssānian (224–642 CE), and the eighth to the twelfth centuries (Mousavi, 2001; Rante, 2008, 2010, 2014). The Citadel rampart was destroyed by the order of Noāaym b. Moqarrin’s ca. 643 during capture of the city by the

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

Arab invaders (Rante, 2008, 2010, 2014). Southwest of the Tappeh Mil fire temple, southeast of Ray (Fig. 18), lay other Sāssānian settlements of Chāl Tarkhān, Eshqābād (sic, Ashkābād), Nezāmābād and Qal’eh Gabr (Thompson, 1976; Schmidt, 1935a, 1935b, 1936a, 1936b, 1936c). A burned layer 2 cm thick covered the Saljuq phase of the city of Ray (1000–1218). The very same burned layer was also found on the Citadel and can be associated with the destruction of the entire settlement by the Mongols in 1220 and not by any earthquake. Despite archaeological history of seven millennia, no archaeoseismic indicators have yet been cited in the excavation reports, though they were not particularly looking for such indicators. The Sureni/ Cheshmeh ‘Ali site was suddenly abandoned in the third millennium BCE for several centuries (Schmidt, 1940), was again inhabited at the end of the second millennium BCE, and later was used by the Parthians (312 BCE–CE 224) and Sāssaniāns (224–642 CE). The cause of abandonments (droughts, wars, earthquakes) is not known. Communications with Renata Hold and Fredrick Hiebert of the University of Pennsylvania, Philadelphia (February 2000), indicate that no specific reference was made to earthquakes in Eric Schmidt’s unpublished personal field notes (Berberian and Yeats, 2001; Schmidt’s 1934 Ray excavation log books are available at www3.uakron.edu/cheshmehali/ Schmidt%20Log%201934-1936.pdf). The Sāssānian period (224– 642 CE) monumental Mil palace and fire temple, located south of Cheshmeh ‘Ali (Sureni in Fig. 18; Schmidt, 1936a, 1936b, 1936c, 1936d, 1937, 1940; Naumann, 1964; Choksy, 2007), do not show evidence of large-magnitude earthquakes due to reconstruction carried out during the twentieth century. However, the pre-repair photograph of the Mil fire temple shows the collapse of the dome and one of the four vaults (see figure on page 55 in Kariman, 1976). The later period archaeological sites at Tehran (Qaytarieh, Darrus, Saltanatābād, Bustān-5, ‘Abbāsābād, and Galanduak), Ray (Kahrizak, Pishvā, and Ayvān Kay sites; Fig. 18), and Karaj–south Tāléqān (Chandār and Khorvin sites; Fig. 16) were actively occupied during the period of 1200–1000 BCE. The whole region (at least 50 km in the E-W direction and 40 km in N-S direction) was abandoned around 1000–900 BCE, and the sites were never occupied again (KambakhshFard, 1969, 1991; Curtis, 1989). The people at the Sartakht mound (Figs. 13 and 18) in south Tehran (discovered in 1983 by the first author) left behind more than 1000 active pottery kilns on the Kahrizak escarpment (Kambakhsh-Fard, 1991). Kambakhsh-Fard (1991) stated that the Qaytarieh Spring near the Qaytarieh mound in central north Tehran (QA in Fig. 18) has very little water at the present time. The reason for this abandonment of the site and change in the flow of the Qaytarieh Spring is not clear (Berberian and Yeats, 2001). 743 Ms ~7.2 (?) Caspiae Portae (Caspian Gates) Earthquake Theophanes (eighth century) mentioned an earthquake during the reign of Byzantine Emperor Constantine V Copronymus (741–774 CE) and wrote that: “In the year 6235 [of Alexander era, 25 March 743–24 March 744 CE] there was an earthquake in the region of Caspian Gates” (Theophanes, eighth century, 418.14–15). This report was repeated later by Cedrenus (twelfth century, 805–806). No other source records this event (Ambraseys, 1974; Berberian et al., 1985; Guidoboni et al., 1994; Guidoboni and Traina, 1995; Berberian, 2014). Based on this very brief phrase mentioned in the Byzantine sources, Ambraseys and Melville (1982) assigned a magnitude of Ms 7.2 and a macroseismic epicenter at 35.3°N, 52.2°E at the Sardarreh Defile (Caspian Gates) southeast of Ray for this event (Fig. 18). The same parameters were presented by Gorshkov et al. (2009) and NOAA (2014). However, Kondorskaya and Shebalin (1977, 1982), and later Guidoboni et al. (1994) and Guidoboni and Traina (1995) declared that the earthquake took place at the “Caspian Gate” in the northern Caucasus at the western Caspian Sea coast of Derbent (“Darband”; lit.

Tehran: An earthquake time bomb “the Gate” in Persian; located in modern Dagestan in the northern Caucasus), where the coast with fortress wall sank into the Caspian Sea. Kondorskaya and Shebalin (1977, 1982) assigned an ~Ms 5.5 ± 1.0 and MSK (Medvedev-Sponheuer-Karnik scale) intensity ~VII for this event, with its epicenter at 42.1°N, 48.2°E which is ~860 km northwest of the location given by Ambraseys and Melville (1982) in the area southeast of Ray (Berberian et al., 1985; Berberian, 2014). It is impossible to assign a magnitude to an earthquake in southeast Ray for which the exact location has not yet been established. At least five historic “Caspian Gates” were named by the Greek and Latin authors at: the pass of Derbent/Darband (Dagestan), the pass of Dariel (Georgia), in Armenia, in Tālesh, and between Media and Parthia in Iran (Jackson, 1911; Anderson, 1928; Hansman, 1968; Standish, 1970; Bosworth, 1983; Simonian, 1989; Guidoboni et al., 1994; Guidoboni and Traina, 1995). Despite comments and arguments expressed by Guidoboni (1994) and Guidoboni and Traina (1995) about the location of the 743 event at Dagestan, northern Caucasus, Ritz et al. (2003, 2012), Nazari (2006), Nazari et al. (2005, 2009), and Asadi and Zare (2014) still located the 743 earthquake southeast of Ray. Nazari et al. (2005) drew an ellipse representing the meizoseismal area of the 743 earthquake along the North Ray fault. Nazari et al. (2005, 2011) introduced another event in 793, which is a belated duplicate of the 743 earthquake. Nateghi-A. (2001) assigned this event along the Garmsār fault, whereas Ashoori (2004) stated that renewed activity of the Garmsār fault and salt domes was responsible for the 743 earthquake at the “Caspian Gates” of Iran. Later, Djamour et al. (2012) drew an ellipse representing the meizoseismal area of this misallocated event along the Pārchin fault (Berberian, 2014). Finally, Nazari (2015) stated that the 743 earthquake took place along the Pārchin fault. The Ms 7.2 magnitude for this misallocated event was adopted as a source input in the probabilistic seismic risk assessment of Tehran by Ghodrati Amiri et al. (2003, 2004, 2012). It is necessary to reject the event in Ambraseys and Melville (1982), Berberian et al. (1985), and Berberian and Yeats (1999, 2001) at Sardarreh Defile, southeast Ray, as well as the aforementioned cited references, and its usage in the national hazard map (Fig. 18; Berberian, 2104). We, therefore, delete this event from our catalogue of earthquakes in the Ray-Tehran region.

145

rary and near-contemporary sources, we delete it from our catalogue for the Tehran-Ray region. 855 May 22–856 May 11 MS ~7.1 Ray Earthquake

A dubious earthquake on 15 or 17 July 850 (236 H [Hijra]) is reported in Ray with severe damage and 45,000 dead by Hasani Sayed Mortezā ibn Qāsem Rāzi (Hasani Rāzi, ca. 1204–1225, p. 196), followed by Kariman (1970, II.243–248), Nabavi (1972, 1978), Moinfar (1978), Ravandi (1985, V.416), CHN (2005), Nazari (2006), NOAA (2014), and Utsu (2002, 2014). The death toll of 45,000 is also given for the 856 Komesh (Dāmghān) earthquake (discussed later). Because there is no record of the earthquake in contemporary and near-contemporary sources, and it may be a duplicate of the 855–856 Ray earthquake (discussed below; Berberian et al., 1985), we delete it from our catalogue.

The native poet of the city of Ray, Badr al-Din Qavāmi Rāzi (born in the second half of the eleventh century; died after 1165; Qavām Rāzi, second half of the eleventh century, p. 1), in a “qasideh” poem, lamented 350,000 lives lost during an earthquake that ruined the city of Ray during the time of Yahyā ibn Mo’āz, who died in 258 H/CE 871–872 (Kariman, 1970, I.106, II.244; Nabavi, 1972, 1978; Ambraseys, 1974; Berberian et al., 1985; Ravandi, 1985, V.416; CHN, 2005). As mentioned already, Hasani Sayed Mortezā ibn Qāsem Rāzi (Hasani Rāzi, ca. 1204–1225, p. 196), wrote that in Shahr-e Ray, 45,000 people perished during the 850 earthquake. The number is also mentioned for the 856 Komesh (modern Dāmghān) earthquake; however, he categorically stated the damage in the city of Ray. Although the casualty figure is highly exaggerated and the data have been amalgamated with those of the 22 December 856 Komesh earthquake (discussed below), it could, nonetheless, be indicative of the magnitude of the calamity. Its timing, being very close to the 22 December 856 Komesh earthquake, ~250 km ENE of Ray (Fig. 15), may persuade us to delete the 855–856 Ray earthquake from the catalogue. However, the narrative of Qavāmi Rāzi (eleventh century CE, p. 1), the native of the city of Ray, written ~300 yr after the event, may prevent this action, because he might have had access to reliable sources. This can only be resolved with detailed paleoseismologic trench studies along the faults in the Ray-Tehran region (Fig. 4). Apparently, Zamakhshari Khārazmi (1075–1144; see Yaqmāee, 1947, p. 10) wrote that in year 241 H (22 May 855–11 May 856 CE), Ray, Jorjān (Arabicized Gorgān, modern Gonbad Kāvus), Tabarestān (modern Māzandarān province), Neyshābur/Neishābur, Qom, Kāshān, and Dāmghān were destroyed by an earthquake, and 25,000 people were killed (see inset bottom right in Fig. 15 for the locations). Although he apparently amalgamated three earthquakes of Ray (855– 856), Neyshābur (856–857), and Komesh (856), his citing of Qom (85 km SSW of Ray) and Kāshān (177 km south of Ray) may indicate that the 22 May 855–11 May 856 Ray, or the 856 Komesh earthquake was strongly felt at these two cities (Fig. 15, inset lower right). We have not been able to trace Zamakhshari Khārazmi’s book named al-Modhesh; however, there is another book with the same title by ibn al-Jauzi (1114–1201). Zamakhshari Khārazmi’s statement (if any) was later repeated by ibn al-Jauzi (1181, abridged version of Mukhtasar: fol. 85v-86r; who added the city of Esfahān to the existing list; Fig. 15), Mirkhānd (ca. 1498), al-Suyuti (pre-1499), Khāndmir (1521), and many more. Two centuries after Zamakhshari Khārazmi, Ibn al-Athir (1231, VIII.53) placed the earthquake in the year 241 H (22 May 855–11 May 856) and wrote that an unknown number of people were killed; aftershocks continued for 40 d and added more destruction. Utsu (2002, 2014) reported this earthquake on 22 May 855. Ambraseys and Melville (1982) estimated a magnitude of Ms ~7.1 with a macroseismic epicenter at 35.6°N, 51.5°E. The extent of the damaged zone of this event is not clear, and its source fault is unknown. Nonetheless, Djamour et al. (2012) mentioned that the 855 earthquake might have taken place along the North Tehran fault. In contrast, Ritz et al. (2012, p. 13) suggested that “the Pishvā fault might be the source of the 855 A.D. event.” Finally, Nazari (2015, p. 267) stated that “the 855 Ray earthquake is the same earthquake of 856 Komes with [the] source fault being on the Astaneh or Astaneh and Firuzkuh faults.”

853 Spurious Ray Earthquake

856 December 22 Komesh (Modern Dāmghān) Earthquake

Nazari et al. (2005) reported an event in 853 at Ray with no reference cited. Because there is no record of this earthquake in contempo-

The Komesh (modern Dāmghān) earthquake of 22 December 856, with Ms estimated as 7.9 (Ambraseys and Melville, 1982; Fig. 15),

793 Spurious Ray Earthquake In addition to the 743 earthquake, Nazari et al. (2005) proposed a damaging earthquake in 793 at Ray. Later, Nazari et al. (2011) only referred to the 793 earthquake at Ray. The source for the 793 earthquake at Ray is not known, and no reference of it has been found in the historical chronicles. We, therefore, delete this spurious entry from the data that will not be cited as an earthquake in future publications. 850 Spurious Ray Earthquake

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

146

Berberian and Yeats

has been considered in numerous papers since 1982 to be the largest earthquake ever recorded in Iran. East of Tehran, the eastern Alborz Mountains was the source of the earthquake (Fig. 9), reportedly killing 40,000–200,000 people (Ambraseys and Melville, 1982; Berberian, 1994; Berberian et al., 1996). Gorshkov et al. (2009) assigned a magnitude of Ms 8.1 for this event. If these estimates are correct, the Komesh earthquake would be the worst-case earthquake (maximum credible earthquake, or MCE) to be expected for the Alborz, including the Tehran region. We reviewed the contemporary and near-contemporary historical evidence for this earthquake and conclude that the meizoseismal area and the magnitude of this earthquake were exaggerated, and its magnitude was probably much smaller. The closest surface rupture to the destroyed city of Dāmghān (originally “Deh-Moghān”) is the Dāmghān fault (Fig. 15; Berberian, 1976c); recent faulting along the North Dāmghān and the Āstāneh faults is visible to the north and the west (see figure 10, p. 197 in Berberian, 1976c; plate 11.1 in Berberian 2014). Like Ray/Tehran, the medieval city of Komesh (“Kumes”; Arabicized “Qumis”; modern Dāmghān: 36°10′N, 54°20′E; see Fig. 15) was situated at the southeastern foot of the Alborz Mountains (Fig. 9) along the ancient High Road/Silk Road between Mesopotamia and China. However, as we report here, several separate earthquakes occurring and felt in different parts of the region in the same year, including Neyshābur/ Neishābur (400 km to the east of Dāmghān), Jorjān (modern Gonbad Kāvus), Tabarestān (modern Māzandarān), Ray, Qom, Kāshān (380 km to the southwest), and Esfahān (480 km to the southwest; see Fig. 15 inset lower right for the locations), led to an overestimation of the meizoseismal area of the 856 Komesh earthquake by Ambraseys and Melville (1982) and others after them. The 856 earthquake might have been felt in all of these areas; however, it is unlikely that these areas were all damaged and/or destroyed by a single earthquake on 22 December 856. Although the 22 May 855–11 May 856 Ray earthquake, which was possibly felt in Qom, Kāshān, Esfahān, and Tabarestān (Fig. 15), took place prior to the Komesh earthquake, this earthquake was apparently combined in later chronicles with the 22 December 856 Komesh event. As for the earthquake at the city of Neyshābur (Fig. 15, inset lower right), that event (856–7) was first mentioned 18 yr later by Ya’ghubi (892, II.600) with an exaggerated death toll of 200,000 people, which, due to its distance, should be considered as a separate event (if any). It is interesting that the contemporary writer categorically mentioned only: “occurrence of a strong earthquake with long duration during the night in the land of Komesh [present Dāmghān], where all the houses were destroyed killing many people” (quoting the contemporary poet Dāvud ibn Tahmān al-Baihaqi, in Ibn Funduq Baihaqi, 1169; see endnote to page 138 by Ahmad Bahmanyār, the editor of Ibn Funduq Baihaqi, 1169, p. 362). Eighteen years after the earthquake, Yaghubi (892, II.600) stated that 200,000 were killed by the earthquake. About 47 yr after the 856 Komesh earthquake, ibn Rosta (Rosteh; sic, Rasteh) Esfahāni (903, I.199) visited the area and wrote that: “from Komesh [present Dāmghān] to Haddādeh village [near the modern Dehmollā on the Dāmghān fault; Fig. 15] the houses and caravanserais destroyed by the earthquake were still visible.” Later, Tabari (915, III.3, XII.1433–1434) wrote that “a fearful earthquake took place in Komesh (Dāmghān) and its districts this month; houses were destroyed and about 45,096 people were killed, mostly in Dāmghān.” (The given precise number of 45,096 seems to be the death toll in the Dāmghān district and not the town; a common practice in Iran.) The event was then reported by Ibn al-Jauzi (1181, fol. 85v-86r), Ibn alAthir (1231, VIII.53), Bar Hebraeus (1286, p. 143), and many more. Although we consider that the death toll is still exaggerated, this report is indicative of the extent of damage. The population of Dāmghān during its zenith in the first half of the eleventh century CE was ~25,000. During the Timurid (1370–1502) and Safavid (1491–

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

1722) eras, the population dropped to ~2000–3000. It was only in 1991 when the population of Dāmghān reached 40,954 (Adle, 1971, 1993; SCI, 2013). We include these population estimates as evidence that the casualty number of 40,000 for the arid Dāmghān in 856 CE has been highly exaggerated, and the alleged death toll of 200,000 for the meizoseismal area is simply impossible because the population of the Dāmghān would have been far less than 25,000 during the time of the 856 earthquake (though there is no record of population during the earthquake). Unfortunately, the unproven numerical account of 200,000 fatalities of this earthquake entered into the world catalogues (Utsu, 2014; USGS, 2014; NOAA, 2014; Holzer and Savage, 2013, among others). Utsu (2014) reported two entries on 3 December 856 (with severe damage at “Qum, Semnan, Damgha, Khorasan,” with 48,690 death toll) and 22 December 856 (with extreme damage and 45,096 death toll at “Kumes, Damghan, Ray”), which is incorrect. Review of the historical data in the first 59 yr after the 856 earthquake only confirms that: (1) the town of Dāmghān was destroyed; (2) the villages from that town to Mehmāndust and Haddādeh (~40 km ENE of the town; Fig. 15) were destroyed; and (3) the death toll was reported to be 45,096, possibly in the whole district including the town. Comparing the 856 earthquake death toll of 45,096 with the 1990 Mw 7.3 Rudbār earthquake in the western Alborz (with population of the country at 59.9 million), where four cities were damaged or destroyed (Figs. 6 and 9), the number is still a very high figure for Iran in 856. We, therefore, do not regard this earthquake covering an area of 181 × 45 km as the most destructive earthquake in Iran (Fig. 15). No direct reference has been made to any destruction or damage to the west and southwest of Dāmghān. It was only ~642 yr after the earthquake that Mirkhānd (ca. 1498, III.478), followed by Khāndmir (1521), wrote that: “half of Dāmghān and 1/3 of town of Bastām (mentioned for the first time in the chronicles) were destroyed”; they also added Ray, Jorjān (modern Gonbad Kāvus), Neyshābur, and Esfahān as being hit by the same earthquake, combining the damage from at least three known earthquakes at Ray (855–856), Dāmghān (22 December 856), and Neyshābur (856–857; Fig. 15, inset bottom right). Therefore, the size of the damaged area discussed in previous literature does not support a region of maximum destruction covering an area of 181 × 45 km, cutting the structural trend of the area, and encircling several active faults, and leading to a mega-earthquake with a magnitude of Ms 7.9 with 200,000 death toll, as estimated by Ambraseys and Melville (1982), nor does it support a maximum area of destruction 150 km long, Ms 7.9, and 200,000 death toll as presented by Hollingsworth et al. (2010), followed by Nemati et al. (2011), Rizza et al. (2011), Nazari (2006), Nazari et al. (2014), and Nazari (2015) (Fig. 15). Figure 3.1 in Ambraseys and Melville (1982, p. 38) does not show any destroyed or heavily damaged sites within their proposed meizoseismal area of the 856 earthquake. No evidence exists that the towns of Shāhrud (established in the early nineteenth century as a town during the reign of Fath ’Ali Shāh Qājār: 1792–1834), Bastām (originally, Bestām), Āstāneh, the ancient city of Komesh (“Hecatompylos”; lit. “the City of 100 Gates”), and villages such as Āhuvān and Tāsh (shown in figure 3.1 in Ambraseys and Melville, 1982; figure 1 in Hollingsworth et al., 2010) were destroyed by the 856 earthquake (Fig. 15). Their destruction is not confirmed by any contemporary or near-contemporary sources. Neither the historical documents nor archaeological investigations at the ancient city of Komesh (“Hecatompylos”; see following) support its destruction during the 856 earthquake (Fig. 15). Therefore, the statements by Ambraseys and Melville (1982), Hollingsworth et al. (2010), and Nazari (2015) regarding destruction of Hecatompylos by the 856 Ms 7.9 earthquake are not confirmed. The declaration by Hollingsworth et al. (2010) that (1) “The 856 A.D. Qumis earthquake (M 7.9) is the most destructive earthquake to have occurred in Iran, killing more than 200,000 people and destroying the cities of Damghan and

Tehran: An earthquake time bomb the old Partian capital of Shahr-I Qumis (Hecatompylos) (p. 1), and (2) “The 856 A.D. Qumis earthquake is one of the five most destructive earthquakes to have occurred in recorded history” (p. 18) cannot be confirmed by the annals. Paradoxically, Nazari et al. (2014, p. 134) and Nazari (2015, p. 264) claimed that their indirectly dated 763–819 event on the Firuzkuh fault trench was caused by the 856 Komesh earthquake. They stated that the Firuzkuh fault either ruptured during the 856 earthquake or was triggered by the 856 earthquake on the Āstāneh fault. They concluded that the 856 Komesh earthquake rupture continued as far west as the Firuzkuh fault to the west, a distance of 148 km to the southwest of Dāmghān (Figs. 9, 14, and 15). Nazari (2015) added that the old city of Ray was located in the realm of the 856 large earthquake, which is not true. The distance between the confirmed destroyed site of Haddādeh (modern Dehmollā; 40 km east northeast of Dāmghān; Fig. 15) and Firuzkuh is 190 km (Fig. 14), and the distance between Haddādeh and Ray is 230 km; no earthquake on the Iranian Plateau has ever shown coseismic rupturing of almost 200 km in length. The length of the E-W–trending Dāmghān fault, from its bend near the Āstāneh valley fault in the west to the eastern end of the fault is ~65 km. The fault bends to the southwest and enters the Āstāneh Valley. To the east, it dies out ~52 km east-northeast of Dāmghān (Fig. 15). The Ms 7.9 magnitude was reported later by Berberian et al. (1996), Hollingsworth et al. (2010), and Nemati et al. (2011). Hollingsworth et al. (2010) claimed that the region of maximum destruction for the 856 earthquake extended for a length of more than 150 km and coincided exactly with the Āstāneh fault (Fig. 15). This is not confirmed because the destroyed village of Haddādeh (mentioned above) is located on the Dāmghān fault, and its distance across strike to the Āstāneh fault, where the trench was cut (36.25°N, 54.00°E), is ~70 km (Fig. 15). The destroyed city of Dāmghān is located ~20 km south of the Dāmghān fault and 30 km east of the Āstāneh fault (Fig. 15). Based on the empirical relations between surface fault rupture and earthquake moment magnitude (Wells and Coppersmith, 1994), the 65-km-long east-west strike-slip Dāmghān fault (from the fault bend in the west to its eastern tip) would have been capable of generating an ~Mw 7.2 earthquake with ~2.4 m left-lateral displacement. We delete the spurious duplicate earthquake of 26 April 662 titled “Qumis, Semnan and Damghan” cited by Nabavi (1978), and shown with damage and 40,000 dead (Utsu, 2002, 2014), from the catalogue, since there seems to be no record of this event in contemporary and near-contemporary sources. The statement by Nazari et al. (2014) and Nazari (2015), that during the 856 earthquake, both the Āstāneh and Firuzkuh faults were activated, and the statement by Hollingsworth et al. (2010), that multiple faults (Āstāneh, Dāmghān, and North Dāmghān) were activated, cannot be confirmed (Fig. 15). Archaeoseismic Evidence No clear archaeological evidence for the 856 Komesh earthquakes has yet been reported. The Tārikhāneh (“Khodāi-Khāneh,” lit. “House of God”; also known as “Chehl Sotun”; lit. “Forty Columns”) congregational mosque at Dāmghān, built in 760 CE with Sassanid [224–642 CE] architecture, form, materials and technique characters (Godard, 1934, 1936; Pope, 1965; Rafi’, 1983; Bloom, 1989; Blair, 1992; Adle, 1993, 2011), was damaged during the 856 earthquake, and its square minaret collapsed during the event. Only ~3 m of the basal original attached minaret base survived (Godard, 1934), and it was later replaced by a detached minaret built in 1026–1029 (Godard, 1934, 1936). During the 2008 excavations in the northeastern corner of the Tārikhāneh wall, remnants of the old minaret, two square-based adobe columns with a plaster floor around them, and a brick floor (in the NW) were discovered (CHN, 2008), which could possibly be the remnants of the pre-856 earthquake structures (communications with Zarrintāj Shaibāni, the excavation project archaeologist, on 4 March 2014 were

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

147

not returned). The present vaults at Tārikhāneh (exact date unknown) are constructed upon square wooden boards placed on the top of the huge columns (see figures 71–73, p. 79 in Pope, 1965). This innovative step, together with the use of timber in the walls, possibly was taken during reconstruction after the 856 earthquake in order to dissipate the strong ground motion effects on the structure. However, we lack detailed archaeological data from this site. The Komesh archaeological mound, in the ancient Parthian capital city of “Shahr-e Komesh” (“Hecatompylos”) with Pārthian (250 BCE– CE 224) and Sāssānian (224–642 CE) remains (Hansman, 1968; Hansman and Stronach, 1970, 1974; Stronach, 1977, 1979), as well as the “Hessār” mound (fifth millennium BCE to ca. 1700 BCE; Dyson and Howard, 1989; Schmidt, 1937; Thornton and Rehren, 2009; Thornton et al., 2013) have not been carefully investigated for archaeoseismic indicators during the excavation (filled triangles in Fig. 15). David Stronach (16 October, 1998, personal commun.) indicated that both he and John Hansman, while investigating the limited excavated areas within the bounds of a very large settlement of the “Komesh” site (33 km SW of Dāmghān; Fig. 15), did not observe any major cracks or structural displacements in the Komesh mound. Apparently, around 600 CE (long before the 856 earthquake), the seat of government was transferred from “Shahr-e Komesh” (“Hecatompylos”) to Dāmghān (Hansman, 1968; Adle, 1993, 2011). Hence, the destruction and abandonment of “Hecatompylos” by the 856 earthquake and transfer of the seat to Dāmghān stated by Ambraseys and Melville (1982) and Hollingsworth et al. (2010) cannot be attributed to the size and magnitude of the 856 earthquake. Robert H. Dyson Jr. (1 December 1998, personal commun.), discussing the Hessār archaeological mound (ca. 4590–1705 BCE), located ~4 km southeast of Dāmghān (Fig. 15), indicated that “in the ‘Burned Building’ (ca. 3670–2955 BCE), a piece of fallen wall found in the northern courtyard and a block of brickwork fallen from the stairway could have been due to an earthquake. However, it is also possible that they were pulled down with the collapse of the burning roof.” Development of a vertical fracture and displacement of a large basin-shaped furnace of tempered clay observed in Phase 7 in the “Dumping Yard” in DF89/3/, below the western wall of Room 9 (period II: ca. 3670–3500 BCE) at the South Hill of Hessār mound (see figure 7, p. 41 in Tosi and Bulgarelli, 1989), may have been caused by the 856 earthquake or an earlier unknown event. The external buttressed wall innovation (repeating buttresses as a bonded part of the wall, as opposed to a single buttress added at odd points to an existing wall for stabilization) with a sophisticated system of brick constructed around 3380–2880 BCE (Dyson and Remsen, 1989; Tosi and Bulgarelli, 1989) may be an earthquake-resistant design at the Hessār mound following an ancient earthquake in order to support the walls from strong ground motion. However, this cannot be demonstrated with the limited available evidence and lack of paleoseismic trench study. Extant historical monuments in Dāmghān (Meshkati, 1970; Rafi’, 1983) indicate that no large-magnitude earthquake has taken place in the region since the 856 earthquake. Earthquakes were reported in the area in 1102 (reported in 1880) and 1852 (Sani’ al-Dauleh, 1880–1882; Ambraseys and Melville, 1982; Berberian, 1994), but their source parameters and effects on the structures in the region are unknown. In conclusion, the 856 earthquake (Fig. 15) seems to have been much smaller than proposed by Ambraseys and Melville (1982, and successive references), and the previously proposed magnitudes for it are not consistent with other earthquakes in the Alborz for which magnitudes are well known (Fig. 9), including the 20 June 1990 Mw 7.4 Rudbār earthquake. 864 January 15–February 12 Ray Earthquake According to Tabari (915, III.3, 1515; ed. Pāyandeh, V.1975), a violent earthquake in Dhu’l Hijja 249 (Arabic lunar month) destroyed

148

Berberian and Yeats

many houses and killed a multitude of people at Ray, where the survivors fled into the countryside. This statement was later repeated by ibn al-Athir (1231, VII.82; ed. Azhir, IX.4141), al-Suyuti (pre-1499, p. 28), and Sani’ al-Dauleh (1880–2, I.97). The latter added the cities of Neyshābur (670 km to the east of Ray; see Fig. 15, inset lower right), Qazvin (Fig. 9; 148 km to the NW), and Tabriz (520 km to the NW) as affected areas (Kariman, 1970, II.243–248; Nabavi, 1972; Ambraseys, 1974). It is probable that the 864 Ray earthquake was felt in Qazvin. The reference to Tabriz could be the 13 February 863 Dvin, Armenia, earthquake (Guidoboni, 1994, 1997; Guidoboni and Traina, 1995), which was definitely felt in Tabriz. A damaging earthquake was reported in the Khorāsān Province of NE Iran (Neyshābur city was not mentioned), but it took place in the year 245 H/8 (April 859– 27 March 860; al-Suyuti, pre-1499, p. 27). Ambraseys and Melville (1982) assigned a magnitude of Ms 5.3 for the 864 Ray earthquake, with its macroseismic epicenter at 35.7°N 51.0°E, which cannot be warranted due to lack of enough macroseismic information. Hence, the source of the event is not known. Gorshkov et al. (2009) increased the magnitude to Ms 6.5. We tentatively assign an intensity of >VII+ (MMI) and ~Ms >6.0 based on destruction of the houses and casualties reported in the city. Nabavi (1978), followed by NOAA (2014), reported two separate earthquakes at Ray on 24 February 863 and 15 January 864, which probably refer to a single event. Utsu (2014) showed two earthquakes in (1) 863, with severe damage at “Ray/Armenia: Yerevan,” and (2) on 24 February 863, with severe damage at Ray. Asadi and Zare (2014) located this earthquake in the area west of Tehran. This clearly indicates that the later authors used different sources without analyzing the contemporary or near-contemporary sources and built corrupt entries with no solid foundation. 898 January 28–899 January 16 Dubious Ghār (Ray)– Tabarestān Earthquake Hājji Khalifa (1648, p. 56) reported: “An earthquake of extraordinary violence in 285 H (898–899 CE) ruined the district of Ghār and Tabarestān.” This statement was later followed by Wilson (1930, p. 113, on the authority of Mallet, 1850, and of Hājji Khalifa, 1648), Nabavi (1972, 1978), Ambraseys (1968, p. 488, 1974, p. 71), NOAA (2014), Berberian et al. (1985, p. 226), and Nazari (2006). The district of Ghār included Ray (with 40 villages such as Tehran, Firuz Bahrām, Daulatābād, Shāh ‘Abdol’azim, etc., in southern Tehran). The Tabarestān (modern Māzandarān) Province has been located to the north of Ghār in the Alborz Mountains south of the Caspian Sea. The source for Hājji Khalifa (1648), who wrote 749 yr after the alleged event, is not known, and since this event is not confirmed by any contemporary or near-contemporary Persian sources, we delete it from our catalogue. 958 February 23 Ruyān Earthquake The only available reliable contemporary chronicle for the 23 February 958 Ruyān earthquake is by ibn al-‘Amid (ca. 960), the prime minister of the Buyid ruler Rokn al-Dauleh (Rukn al-Dawla, r. 935–976), who resided at the city of Ray. Biruni (1025, 48.1–5) recorded that Ibn al-‘Amid related in his book, Fi Binā’ al-Mudun [On the Construction of Cities, 970], that ‘an earthquake took place in ‘Ruyān,’ not long ago, which made two mountains collide and tumble down, and that the debris of the collision blocked the course of the rivers which ran between them, and that the waters of the rivers receded and formed a ‘lake.’ This is what usually happens when the water has no outlet; like the Dead Sea which is formed by the water of the Jordan River. (Biruni, 1025, 48.1–5, ed. ‘Ali, 1967; ed. Ahmad Ārām, 1973, quoting contemporary writer ibn al-‘Amid)

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

Although the lake related to the 958 Ruyān earthquake was formed by a rock avalanche blocking a river and the Dead Sea is a rift formed by the Dead Sea transform fault, we preserve the contemporary author’s observation as written. Almost two to three decades after the earthquake, both ibn Hawqal (978) and al-Muqaddasi (985) spoke of Ray as already much gone to ruin (repeated by Le Strange, 1905); however, the cause was not addressed. Nineteen years after the earthquake, ibn Hawqal (978, p. 176), who traveled from 943 to 969 CE, wrote that: “The citadel [of Ray] is in good repair; and there is a wall round the suburbs, which is, however, falling to decay and almost desolate” (ibn Hawqal, 978, p. 176; tr. and ed. Ouseley, 1800). It is interesting to note that archaeological excavations at a trench in the ancient city of Ray at 35°36′14.57″N, 51°27′00.95″E revealed traces of many destroyed mud brick adobe structures of the tenth century CE (Rante, 2014; Rocco Rante, Musée du Louvre, Paris, 4 February 2014, personal commun.). This indicates that the 958 earthquake in the northern Ruyān Mountains of Ray caused destruction or damage at the city of Ray in the south (Figs. 19 and 20). Rante (2010, 2014) reported a four-century lacuna in the ancient Ray Citadel Hill prior to occupation by the Timurids (1370–1500), where a 1432–1433 minted coin of Shāhrokh (r. 1405–1447) was discovered. The cause of this lacuna at the ancient Ray Citadel Hill is not clear. It may have followed the 958 Ruyān earthquake; however, we need additional archaeological evidence to support this assumption. About 72 yr after the earthquake, ibn Miskawayh (sic, Ibn Maskuyeh Rāzi; ca. 1030, II.167), a historian from the affected city of Ray, wrote that: “The earthquake affected Ray and its districts causing a very large number of casualties.” Neither of these two reliable contemporary (ibn al-‘Amid, ca. 960) and near-contemporary (ibn Miskawayh, 1030) local sources residing at Ray mentioned earthquake effects in Tāléqān (either Tāléqān of Iran located more than 100 km NW of Ray [Fig. 4], or Tāléqān of the ancient greater Khorāsān of Iran in modern Afghanistan, 1630 km to the east). Furthermore, as mentioned, archaeological excavation confirmed destruction or damage of adobe structures at Ray during the 958 earthquake. We, therefore, rely on the local contemporary and near-contemporary sources backed by the limited archaeological investigation that the source was in the Ruyān Mountains north of Ray, and the city of Ray was destroyed or damaged by the 958 earthquake (Figs. 19 and 20). About two centuries after the earthquake, Tusi (1167, p. 82) wrote that: “I have heard from people of Guilān [province, SW of the Caspian Sea] that the Kabudān Sea [lit. “the Blue Lake”] was agitated by high tides, the city of Ardébil was shaken; and the distance of Ardébil to the Kabudān Sea is 12 farsangs [~72 km]” (Tusi, 1167, p. 82, ed. Sotudeh). “Kabudān Sea” is the old name of Lake Urumiyeh in northwest Iran, but it is located ~240 km to the west of the city of Ardébil, whereas the Caspian Sea is located 55 km to the east of Ardébil. Tusi (1160) seems to have confused the names of the two bodies of water reported to him. It is probable that the reference was to the 958 Ruyān earthquake effect at the Caspian Sea, which might have caused some changes in the Caspian Sea level, if both took place in the same time (Berberian, 2014). We know that the 1962 Mw 7.0 Bu’in earthquake (located ~146 km southwest of the Caspian Sea, much more that 108 km distance between Ray and the Caspian Sea; cf. Fig. 9) showed abnormal large sea waves (seiches) along the southern Caspian Sea shoreline (Ambraseys, 1962, 1963; Berberian et al., 1983; Naderi Beni et al., 2013; Berberian, 2014). Two-hundred and twenty-three years after the earthquake, ibn alJauzi (1181, VI.384, residing in Baghdād), in describing the 958 earthquake at Holvān (at the modern Iran-Iraq border) and Baghdād, wrote about earthquakes at Ray (23 February 958) and Holvān (April 958), and added “Tālèqān” for the first time in the chronicles. It is not clear if ibn al-Jauzi (1181) was describing two different events at Ruyān/ Ray and at Tāléqān. He definitely combined two different earthquakes

Tehran: An earthquake time bomb at Ruyān (23 February 958 in the southern foot of the Alborz) and Holvān (April 958) at the modern Sar-e Pol-e Zahāb in the western Zāgros near the Iraqi border, ~520 km southwest of Ray. Furthermore, it is not clear that ibn al-Jauzi (1181) in Baghdad received the earthquake report from Tāléqān of the western Alborz (in Iran) or Tāléqān in the greater Khorāsan Province (modern capital of Takhār Province in northeast Afghanistan, 36°43′N, 69°31′E). Almost 363 yr after the earthquake, ibn al-Athir (1231, VIII.390), living in Mosul and Baghdād, wrote that: “Much of the city of Ray was destroyed and many perished; Tāléqān with its districts was also affected and large numbers of people were killed.” Ibn al-Athir’s (1231) reference to Tālèqān could possibly be related to ibn al-Jauzi’s text of 1181. Later, Gregory Bar Hebraeus (1286, ch. 183, p. 165) in Shām (Syria) stated that: In the year three hundred and forty-six of the Arabs [957 CE]… the Great Sea [Caspian Sea] diminished, and it shrunk into itself for a distance of about three hundred cubits; and many rocks and islands were laid bare which had not been known before. And one year later there was a terrible earthquake, and it destroyed many districts, and many were suffocated beneath the overthrow in the mountains of the Desian/ Dailomaye and Kashaa/Kashan. (?; probably Dizeh Qasrān in Ruyān Mountains; DQ in Fig. 19 along the northern Jājrud River, between the North Tehran and Moshā faults) The reported corrupt name of the Dizeh Qasrān (Arabicized “Kuhsārān”) seems to be important, since the village was located in the Ruyān Mountains on the Jājrud River (at Galukān [GAL] in Fig. 19, between Lashgarak and Ushān), in the area near the North Tehran and Moshā faults. At this location, a large landslide of 2 × 2 km at 35°51′40.27″N, 51°32′24.11″E pushed the Jājrud riverbed to the east, where it meanders through the base of the landslide (Dizeh, GAL, and DQ in Fig. 19). This landslide blocked the Jājrud River, since the river now flows through the base of the landslide mass. There are numerous rock mass movements possibly triggered by earthquakes along the Moshā fault (such as the rock avalanche blocking a linear drainage along the Moshā fault that formed the two Tār Lakes [Fig. 12], Lavāsān [Fig. 19], Āb-e ‘Ali [7.5 km west of Moshā village; AB in Fig. 12], and many more), but this one is the largest, and once it had blocked the Jājrud riverbed (Fig. 19). Following ibn al-Athir (1231, VIII.390) and ibn al-Jauzi (1181, VI.384), al-Dhahabi (1315, I.168) described a strong earthquake at Ray and Tāléqān, where Tāléqān “was engulfed” and only 30 people survived, and 150 villages were destroyed. About five centuries after the earthquake, al-Suyuti (pre-1499, p. 30), influenced by ibn alAthir (1231, VIII.390), wrote that the earthquake at Ray lasted 40 d; Tāléqān was engulfed; 150 settlements sunk into the ground at Ray, and a mountain sank, and an enormous chasm opened from which fetid water and smoke gushed out. In his History of the Caliphs, al-Suyuti (pre-1400, p. 399–400) added that the earthquake extended to Holvān (in the western Zāgros), most of which was swallowed up (see also Ambraseys, 1961, 1968, 1974). Later, ibn al-‘Imād (1678, II.371), following al-Suyuti (pre-1499), wrote that the 150 villages belonged to the Ray district and not Tāléqān. It is therefore vital to find out the exact location of the ancient Ruyān (Rudān) Mountains that were mentioned in the contemporary and nearcontemporary annals written at the affected city of Ray. In Persian, “Ruyān” (or “Rudān”) means a “mountainous area with numerous rivers” (Rud = River). We learn from ibn al-Faqih Hamédāni (903, p. 304), Istakhri (951, p. 122), ibn Hawqal (978, p. 320), Bakrān (1208), Yāqut (1226a, II.873), ‘Abd al-Mo’men (1300 in Kariman, 1977), Oliyā’-Allāh Āmoli (1402; who lived in Ruyān), and Mar’ashi (1489, p. 112) that: (1) The Ruyān and the Ray Mountains are connected to each other and the way to Ruyān is from Ray.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

149

(2) The Ruyān Mountains are located between Ray and Tabarestān [modern Māzandarān province, south of the Caspian Sea], and cover the watersheds of the Jājrud, Lār, Lurā, and Shahrestānak Rivers. (3) The area is also known as the Qāren and Rubanj Mountains. The Ray Mountains correspond with the crest of the present Touchāl Mountains of the southern Alborz and the southern Qasrān (Arabicized form of “Kuhsārān”; i.e., “Mountains”) north of Ray-Tehrān (in the south; Fig. 19). The Ruyān Mountains north of Ray cover the Inner Qasrān/Kuhsārān area, with the watersheds of major rivers such as the Jājrud (in the center), Lār (to the east), and a portion of the Karaj River (in the west), and with the mountainous districts of Greater and Lesser Lavāsānāt, Rudbār-e Qasrān, Shahrestānak, and Siyāhrud (Fig. 19). The “Outer Qasrān” (Kuhsārān) covered the Touchāl Mountains (the ancient Ray Mountains) of northern Tehran (at Shemirān), and the present isolated Ray Mountain (Bibi Shahrbānu), with the city of Ray and village of Tehran (Rabino, 1928, 1957; Sotudeh, 1969; Kariman, 1977). Therefore, the location of the ancient Ruyān given by Le Strange (1905) in the area southwest of Kalār (Kalārdasht; SW Chālus: 36°30′N, 51°09′E, and the Valasht Lake at 36°32′N, 51°17′E), located 102 km to the NNW of Ray (VA in Fig. 9), followed by Ehteshami Moinabadi (2007), cannot be warranted. The old city of Ruyān (later called Kojur: 36°23′N, 51°43′E, +1495 m) was located further north (Rabino, 1928, 1957; Stark, 1934). The ancient Ruyān (Rudān) Mountains encompassed 70 km of the Moshā fault (Shahrestānak to Moshā) and 38 km of the eastern segment of the North Tehran fault system (Figs. 19 and 20), and most probably reactivation of one of these two active faults was responsible for the 958 Ruyān earthquake destroying the Ray district. The city of Ray is located 36 km south of the Moshā fault, and 20 km south of the North Tehran fault. Average cross-fault attenuation characteristics from five Mw 7.0–7.3 earthquakes in Iran (1968, 1978, 1981, 1990, and 1997; despite slight variations of their source parameters) show that the average distances of isoseismal lines from the coseismic surface ruptures were 5 ± 1 km (for IX MMI), 15 ± 2 km (VIII), 30 ± 5 km (VII+), and 45 ± 10 km (VI+). Hence, the 958 Ruyān earthquake might have caused intensity of VII– at Ray, located 26 km south of the Moshā fault (Fig. 20). If the 958 Ruyān earthquake took place along the Moshā fault, the same segment of the fault reactivated 872 yr later during the 1830 Mw 7.0–7.4 Lavāsānāt earthquake (discussed in the following). If the account of ibn al-Jauzi (1181, VI.384), written 223 yr after the 958 earthquake while residing at Baghdad (695 km SW of Ray), is correct about the destruction of the Tāléqān district northwest of Ray, then we may consider a second and separate probable event in the same year affecting the Tāléqān district (if the event took place in the Tāléqān district of modern Iran and not in Afghanistan, for which there is no data). In this case, the Tāléqān fault might be considered as the source for the second event (Berberian et al., 1983; Berberian and Yeats, 2001, 2014) rather than the Ruyān earthquake (Figs. 16 and 17). Based on historical data recorded in the chronicles from 960 to 1678, Ambraseys and Melville (1982) assumed a NW-SE–trending meizoseismal area of 150 × 60 km, covering Ray, Tehran, Karaj, Fasham, Tāléqān, and Alamut (36°26′40.72″N, 50°35′10.30″E), with intensities >VIII MMI (see figure 3.3, p. 40 in Ambraseys and Melville, 1982), and assigned a magnitude of Ms 7.7 for the 958 earthquake named by them the “Ray-Taleqan earthquake.” However, figure 3.3 in Ambraseys and Melville (1982) does not show any destroyed or heavily damaged sites within their proposed meizoseismal area. Furthermore, the causative fault of the assumed Ms 7.7 earthquake was not designated among numerous faults covered by their meizoseismal area (figure 8.5, p. 227 in Berberian et al., 1985). As with the assumed magnitude of the 856 Komesh earthquake, if the estimated Ms 7.7 magnitude for the 958 Ruyān earthquake is correct, then this event could be considered as a maximum credible earthquake for the Alborz covering the Ray-Tehran region. The magnitude

150

Berberian and Yeats

of the 958 earthquake was increased to ML 8.0 by Ben-Menahem (1991, mentioning that it was felt in Syria, Turkey, and Israel) and Ms 8.0 by Gorshkov et al. (2009); Utsu (2002, 2014) reported major damage at “Ray, Taliqan” with magnitude 7.7 and intensity X (MMI). Following Ambraseys and Melville (1982), the estimated Ms 7.7 for this event was also reported in Berberian et al. (1985), Berberian and Yeats (1999, 2001), Nateghi-A. (2001), Nazari (2006, 2015), Nazari et al., (2009, 2010), Ehteshami Moinabadi (2007), Solaymani Azad et al. (2011), Ritz et al. (2012), Ghassemi et al. (2014), and Asadi and Zare (2014). The last author located this event not in the Ray region but at Tāléqān (Fig. 16). Ehteshami Moinabadi (2007) stated that the 958 Ms 7.7 Ruyān earthquake triggered a landslide at Valasht (102 km NNW of Ray) and created the present “Valasht Lake,” 15 km southwest of Chālus (36°32′19.93″N, 51°17′21.94″E; VA in Fig. 9). We reviewed the historical evidence for this earthquake and conclude that the previously estimated earthquake magnitude and meizoseismal area were too large. As with the 856 earthquake in Komesh (Dāmghān) discussed already, no evidence exists that the city of Ray, together with villages of Tehran, Karaj, and Zavārak, as well as the famous historical castle/fort of Alamut (lit. “Eagle’s Nest”; later to become the Assassin strongholds) and all the villages in the Tāléqān district to the northwest, were destroyed during a single earthquake of Ms 7.7 in 958 with a vast meizoseismal area of 150 km × 60 km (see figure 3.3, p. 40 in Ambraseys and Melville, 1982). Acceptance of simultaneous rupturing of several major active faults in the Alborz during a single mega-earthquake was followed by De Martini et al. (1998), Nazari et al. (2009), Djamour et al. (2012), Ghassemi et al. (2014), Nazari et al. (2014), and Nazari (2015). For both the 856 (Komesh/Dāmghān) and 958 (Ruyān) earthquakes, the reliable contemporary and near-contemporary local sources give no technical justification to support the conclusion that both events were mega-earthquakes with widespread and vast destruction of a 180-km-long meizoseismal area (intensity >VIII MMI) and simultaneous rupturing of more than one active fault at each event. Sedimentary debris flow units 27, dated 1105 ± 30 yr B.P. (bone sample in debris flow), and 29 in the Irā road-cut trench across the Moshā fault near its junction with the North Tehran fault system (Fig. 4) coincide within uncertainty with the 958 event (Ghassemi et al., 2014), although no evidence of surface rupture associated with the 958 earthquake was documented in the Irā trench (Angela Landgraf, 25 October 2015, personal commun.). Therefore, (1) data from two reliable contemporary local sources; (2) the exact location of the ancient Ruyān Mountains; (3) archaeological data from the city of Ray; (4) data from the Irā trench (with no evidence of surface rupture); and (5) geomorphological data (including landslides blocking rivers and forming a lake) support the reactivation of the central Moshā fault during the 958 Ruyān earthquake (Figs. 19 and 20). Berberian and Yeats (1999, 2001) assumed the western Moshā and/or the Tāléqān fault was responsible for this event, which is corrected here. However, Nazari et al. (2009) claimed that the source of the 958 earthquake might be the Tāléqān fault. They wrote that the 80-km-long Tāléqān fault (Fig. 16) is capable of an Ms 7.7 earthquake. The authors introduced a 120 × 36 km E-W–trending meizoseismal area covering both the Tāléqān as well as the western segment of the Moshā fault. Furthermore, Ghassemi et al. (2014) stated that considering the vast macroseismic area of the 958 earthquake (figure 1 in their report), it probably involved a major fault system that included the Tāléqān, North Tehran, and even Moshā faults. Based on the analyses addressed herein, these interpretations, including activity of more than one fault and the Ms 7.7 magnitude, as suggested by Nazari et al. (2009), Nazari (2015), and Ghassemi et al. (2014), and used as a source input in the probabilistic seismic risk assessment by Ghodrati Amiri et al. (2003, 2004, 2012), are not accepted here. All the evidence shows that the event took place near

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

the Ray region with much lower magnitude (Table 2). Amighpey et al. (2016, p. 169, 170) misquoted Berberian et al. (1985) and Berberian and Yeats (1999) and erroneously assigned the 958 earthquakes to the activity of the North Tehran fault, whereas, this earthquake took place on the Moshā fault. 1119 December 10 Qazvin Earthquake The city of Qazvin, located ~140 km northwest of Tehran, has a similar seismotectonic situation as the city of Tehran; both cities are located at the faulted southwestern range front of the Alborz Mountains (Fig. 9). We therefore briefly mention the only reported earthquake in the city of Qazvin. Rāfe’i Qazvini (1160–1226), a native scholar of Qazvin, wrote that a very violent earthquake occurred in Qazvin with considerable damage; aftershocks continued for a year. By the order of Minister Sadr al-Din Mohammad ibn ‘Abdolāh Marāghi, and the help of Rāfe’i Qazvini’s father, the city was rebuilt (Golriz, 1958, 872–873). The Ashāb Abu Hanifeh (later Haydariyeh) congregational mosque at Qazvin was destroyed. The Abu Hanifeh followers of Qazvin asked Amir Zāhed Khomārtāsh ben ‘Abdollāh ‘Emādi (the governor of Qazvin during the reign of Malekshāh Saljuq) to rebuild the mosque, which was intact until the Safavid period, when it was abandoned. Al-Suyuti (pre-1499, p. 36–37) quoting Rāfe’i Qazvini wrote that Qazvin was destroyed and that the earthquake returned the following year precisely at the same time (i.e., on 28 November 1120 CE, converting from the lunar year). This was cited by Chardin (1811, II.398), Golriz (1958, 872–873), Nabavi (1972), and Ambraseys (1961, 1974). Chardin (1811, II.398), who was in Qazvin in 1673, gave an incorrect date of 1067 and mentioned that the city wall and one-third of the buildings at Qazvin were destroyed. He added that 3 yr after the earthquake, the city was rebuilt by the order of Prince Kahnun Saljuq. Golriz (1958) added that one third of the city and the city wall collapsed in the earthquake. Except the city of Qazvin, no other town or village name is mentioned in the annals (Berberian et al., 1983). Ben-Menahem (1991) stated that the December 10 1119 Qazvin area earthquake that destroyed Qazvin was felt in Israel, which cannot be warranted. Ambraseys and Melville (1982) assigned a magnitude of Ms 6.5 for this event. The closest active fault to the city, the North Qazvin fault (Berberian et al., 1983), located ~11 km north of the center of the city (Fig. 9), might have been the source of this event (Berberian et al., 1983; Berberian and Yeats, 1999, 2001). The North Qazvin fault is the northwestern continuation of the North Tehran fault system. Activity of the eastern section of the Tāléqān and Moshā faults (Figs. 16 and 17) is ruled out because of their distance to the city (50 and 70 km, respectively). No paleoseismic trench study has been carried out across the North Qazvin thrust. The population and the casualty numbers of the city of Qazvin during the 1119 earthquake are not known. The city was the capital of Iran from 1546 to 1597. Its population has increased from 66,420 in 1956 to 381,598 in 2011 (SCI, 2012). Morier (1812), who visited Qazvin in 1809, stated that the city was largely in ruins as a result of a fairly recent earthquake. This statement has not been confirmed. Bosworth (2008) stated that earthquakes were recorded in Qazvin in 863–864, 970–971, 1119–1120, 1120, and 1169. The 864 earthquake took place at Ray, and the 970–971, 1120, and 1169 alleged events also did not occur in Qazvin. Archaeoseismic Data As mentioned already, the Ashāb Abu Hanifeh (later Haydariyeh) congregational mosque at Qazvin was destroyed during the 1119 CE earthquake. According to the northern inscription of the Shabestān (sanctuary) of the present great congregational mosque of Qazvin, the

Tehran: An earthquake time bomb construction took 7 yr (1113–1120). In another inscription, year 1115 was recorded as the end of the construction (Meshkati, 1970). This may indicate that after finishing the construction in 1115, the 1119 earthquake might have caused some damage to the newly built mosque. 1177 May 1–30 Ray-Qazvin Earthquake The first account of the 1177 CE earthquake was given by ibn alJauzi (1181, X.266), who received news that many towns were shaken by an earthquake in Dhu’l-Qa’da 572 (May 1177), and some were swallowed up in the ground, including Ray and Qazvin. About five decades later, ibn al-Athir (1231, XI.287) wrote that Persian towns from the direction of Iraq to beyond Ray were shaken by an earthquake in 571 H (1175–1176 CE), and many houses were destroyed, killing many people, with the worst damage in Ray and Qazvin cities. These two authors were the only two contemporary and near-contemporary accounts for this event, and no other local name is given. Furthermore, it is not clear if these two cities, 152 km apart (Fig. 9), were damaged by a single earthquake. Only Mostaufi Qazvini (1340, p. 63), without mentioning the earthquake, wrote that in 572 H (10 July 1176–29 June 1177), the walls of Qazvin were restored and faced with burnt bricks by the order of the minister of Ālp Arsalān Saljuq. This could have been repairs carried out after the 1177 earthquake. Ambraseys and Melville (1982) suggested that the 1177 Ms 7.2 earthquake source was the same as the 1962 earthquake southwest of Tehran on the Ipak fault located ~150 km west of Ray/Tehran and 65 km south of the city of Qazvin (Fig. 9; Berberian et al., 1983; Berberian and Yeats, 2001; Berberian, 2014). The 1 September 1962 Mw 7.0 Bu’in earthquake (Fig. 9; Ambraseys, 1963; Berberian and Yeats, 2001; Berberian, 2014) did not cause serious damage to either the cities of Ray/ Tehran or Qazvin; so the comparison is not appropriate. The source of the 1177 event could possibly have been the western segment of the North Tehran thrust, the movement on which could have had similar effects in Qazvin (to the NW) and Ray (to the SE) (Fig. 9), if both cities were destroyed/damaged during a single event. We, therefore, may draw an unconstrained schematic meizoseismal area between the two cities until paleoseismic trench studies are done. Ritz et al. (2012), reporting on the trench across a fault south of the North Tehran fault line (Fig. 4), suggested that the most recent earthquake in their trench (E1: 0–4000 yr B.P., Mw 6.2–6.8) could be the 1177 earthquake. An earthquake source in Tehran cannot destroy the city of Qazvin located 140 km NW of Terhan. We delete the 22 July 1175 Ray-Qazvin entry by Nabavi (1978) and NOAA (2014) from our catalogue. Utsu (2014) shows three entries on 22 July 1175 (in “Ray, Qazvin,” with intensity IX, many killed), 1176 (in “Ray, Qazvin,” with M 7.2, intensity IX–X, many killed), and May 1177 (in “Ray, Qazvin,” with M 7.2, intensity IX–X, many killed). Nazari et al. (2005) located the July 1177 earthquake at two different locations at Karaj in the northwest and Ray in the southeast. Following Ambraseys and Melville (1982), Nazari (2006) listed this event as “Boin-Zahra,” the source of the 1962 earthquake, which is not correct. Ca. 1384 Alleged Earthquake at Ray In a Persian autobiography of Timur (the Lame; Tamerlane; the Turko-Mongol warlord) of dubious authenticity edited by Mansuri (1983, p. 95–96, Tehran), reference is made to a great earthquake that leveled the city of Ray with the ground and where bodies were buried under debris 2 yr before the arrival of Timur at Ray (Mansuri, 1983, p. 95–96; Kariman, 1970, II.247; Ambraseys, 1974; Ambraseys and Melville, 1982; Berberian et al., 1985). Timur, the Turco-Mongolian warlord and his hordes captured Ray in November 1384 and spent the winter in the area (Yazdi, 1419–1425).

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

151

This entry is not recorded in the contemporary and near-contemporary Persian sources and books written about Timur. Ruy Gonzalez de Clavijo, the Spanish ambassador (Henry III of Castile) to the court of Timur, on his way to Samarkand, arrived at Tehran on 8 July 1404, ~20 yr after the supposed earthquake. He found both Tehran and Ray with intact buildings and mosques and did not describe any ruined village or report any news about the alleged earthquake. Clavijo (1403–1406) wrote that: This city of Tehran was very large, but it had no walls, and it was a very delightful place, well supplied with everything; but it was an unhealthy place, according to the natives, and fevers were very prevalent… The territory in which it stands is called Rei [Ray], which is a great and extensive lordship, possessed by the son-in-law of the lord. On Tuesday, in the afternoon, they departed, and, at a distance of two leagues, they came in sight of a great city, all in ruins, on the right hand side of the road [referring to the ruins of the ancient city of Rhagae]; but there appeared towers and mosques, and the name of the place was Xahariprey [the ancient city of Ray]. This was once the largest city in all that land, though it is now uninhabited. The ruins of the ancient city of Rei (the Rhages of the Apocrypha), are a few miles south of the city of Tehran. They cover a vast extent of ground, and have supplied materials for the modern capital of Persia. (Clavijo, 1403–1406, p. 98–99) The history of this period is well documented by contemporary and near-contemporary Iranian scholars. Zafarnāmeh-ye Taimuri (The Victory Book of Timur), the Persian contemporary autobiography of Timur, is one of the most important literary works of the Timurid era, written by Sharaf al-Din ‘Ali Yazdi during the period of 1419– 1425, ~35–41 yr after the alleged earthquake destroying Ray (Yazdi, 1419–1425). The book is based on an older Zafarnāmeh written by Nezām al-Din Shāmi, the official biographer of Timur during his lifetime (Timur died on 18 February 1405). No reference to the alleged 1384 earthquake is made in this contemporary book, or in Hāfez Abru (1414), Roemer (1986), and Manz (1999). It is impossible for an earthquake to ruin the city of Ray without any damage to Tehran and villages in between. We, therefore, delete this entry from our catalogue, as well as the 17 March 1382 Ray event reported by Nabavi (1978) and NOAA (2014). Nazari (2006) wrote that the 1384 earthquake was caused by the reactivation of the North Tehran, North Ray, South Ray, Kahrizak, or Pishvā faults. Nazari et al. (2005) located this event at Ray. Later, Majidi Niri et al. (2010) and Nazari (2015, p. 256, 266) in a paleoseismologic trench study assigned the spurious 1384 “event” to the activity of the Pishvā fault in southeast of Ray (discussed earlier in the text; Fig. 4). Finally, Asadi and Zare (2014), without citing a reference or presenting evidence, located this alleged earthquake not at Ray but at the Caspian Gates (Sardarreh Defile; Fig. 18), 90 km to the southeast of Ray. 1428 Tāléqān (Modern Alborz/Iran, or NE Afghanistan) Earthquake: Dubious Locations About 383 yr after this alleged earthquake, al-‘Umari (1793), without citing any reference or mentioning the names of the destroyed villages or towns, briefly wrote that, “In the year 831 H [1428 CE], there was an earthquake in Tāléqān, with shocks lasting for 10 days; many people were killed” (al-‘Umari, 1793, fol.159v). It is impossible to determine from this short phrase whether Tāléqān referred to the Alborz or to Tāléqān of Takhār in the Greater Khorāsān Province of ancient northeast Iran (now the capital city of Takhār Province in northeastern Afghanistan). The distance between these two Tāléqāns is 1680 km. Ambraseys and Melville (1982) and Ambraseys and Bilham (2003) preferred the location of Tāléqān in modern Afghanistan, since al-‘Umari (1793) also reported another earthquake in 1410 at Balkh/Bactra (in modern northern Afghanistan, 234 km west of Tāléqān of Takhār: 36°45′N, 66°53′E)

152

Berberian and Yeats

and Bokhārā (in modern Uzbekistan: 39°46′N, 64°25′E) prior to this event. There is no reference to this earthquake in the Tāléqān of Alborz (Figs. 4 and 16) in the Persian sources. Without any macroseismic data, Nazari et al. (2009) drew an E-W meizoseismal area ~25 km long in the Tāléqān valley of the Alborz Mountains of modern Iran, north of the Tāléqān fault (figure 1 in Nazari et al., 2009), which cannot be confirmed. Similarly, Djamour et al. (2012) assigned the 1428 earthquake to the Tāléqān fault without presenting any supporting data. As mentioned by Ambraseys and Bilham (2003), a tentative location of the event in Tāléqān of northern modern Afghanistan may be preferred until a solid data are presented. 1485 August 15 Upper Polrud Earthquake Although the 1485 Polrud earthquake is farther away from Tehran (120 km NW of Tehran; Fig. 9), evaluating its effects and magnitude is important to understanding the seismicity of the Alborz and the Tehran region. The 1485 large-magnitude earthquake was experienced by Zahir al-Din Mar’ashi (1489, 453–454) and the ruler of Daylamān (in the western part of the meizoseismal area), Soltān Mirzā ‘Ali, who was at prayers when the buildings at Daylamān collapsed, but he was able to escape the disaster (Mar’ashi, 1489, 453–454; Rabino, 1928, 1957; Sotudeh, 1969; Nabavi, 1972; Ambraseys, 1974; Ambraseys and Melville, 1982; Berberian et al., 1983, 1992; Berberian, 1994, 2014; Berberian and Walker, 2010). Based on detailed description of the damage, Ambraseys and Melville (1982) estimated an equivalent surface-wave magnitude of 7.2 for this event, which seems to be a reasonable estimate for this earthquake in Alborz. The earthquake caused widespread destruction and damage in the eastern Guilān and western Māzandarān Provinces of northern Iran, in the High-Alborz Mountains south of the Caspian Sea (Fig. 9). Considering contemporary macroseismic data analysis (Berberian and Walker, 2010; Berberian, 2014), the Kelishom left-lateral strike-slip fault is the most obvious fault along the long axis of the 1485 meizoseismal area (Fig. 9). This fault shows apparent cumulative stream course displacements of >1 km (Berberian and Walker, 2010). We postulate that the Kelishom fault was ruptured during the 15 August 1485 Upper Polrud earthquake. The destructive effect of the earthquake along the Kelishom fault bears resemblance to that of the Rudbār earthquake of 20 June 1990 of Mw 7.3 along the Rudbār fault to the west of the 1485 epicentral area (Fig. 9), showing westward migration of seismicity to the Rudbār gap. The meizoseismal area that we propose for the 1485 earthquake, therefore, assumes that the source faults had similar mechanism, length, trend, and source dimensions, as well as being located in the same structural zone of the High-Alborz (Berberian and Walker, 2010; Berberian, 2014). We delete the 27 August 1484 (“Gilan, Dilaman, Karajian, Shakur, Gulijan”) and 14 August 1485 (“Alamut-Jenat Rudbar-Dailaman”) entries reported by Nabavi (1978), and followed by NOAA (2014), from our catalogue. Archaeoseismic Data Numerous first millennium BCE archaeological sites occur along the Kelishom fault, and some of them, such as Khoshkushān, Kelishom, Pishkāljān, Shāhijān, and Darreh (Kāmbakhsh-Fard, 1991), are close to the fault line (Berberian and Walker, 2010). It would be interesting to study these sites to look for archaeoseismic indicators of the ancient earthquakes in addition to paleoseismic trench study across the Kelishom fault. 1608 April 20 Alamutrud Earthquake About 123 yr after the 1485 Upper Polrud earthquake in the northwestern High-Alborz, another large-magnitude earthquake

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

shook the mountain belt in the area east of Qazvin and northwest of Tehran (Fig. 9). The 1608 earthquake was reported to be severe in five villages (out of many) of northeastern Qazvin, where more than 3000 people were killed (Monajem Yazdi, 1611, fol. 57r–v; Sani’ al-Dauleh, 1880–1882, II.190; Golriz, 1958; Meshkāti, 1970; ‘Asgari, 1971; Ambraseys and Melville 1982; Berberian et al., 1983; Berberian, 1994, 2014; Berberian and Walker, 2010). Ambraseys and Melville (1982), followed by NOAA (2014), estimated a magnitude of Ms 7.6 for the 1608 event, which seems slightly large; we lower the magnitude to Ms ~7.4. Sani’ al-Dauleh (1880–1882, II.190), followed by Golriz (1958), Ambraseys (1974), Berberian (1977), Nabavi (1978), and NOAA (2014), reported a destructive earthquake in 1639 in Qazvin with death toll of 12,000. Utsu (2002, 2014) also reported an earthquake on 4 May 1639 at Qazvin with M 6.1 and a 12,000 death toll. No record of this entry appears in contemporaneous sources, and it seems that it is an erroneous date for the 1608 Alamutrud earthquake that was felt in Qazvin (Ambraseys and Melville, 1982). Sani’ al-Dauleh is not a reliable source, and we delete the 1639 entry from our catalogue. The unconstrained meizoseismal area of the 1608 earthquake may suggest that the Alamutrud reverse fault could be a likely source of the event (Fig. 9; AL in Fig. 6). Djamour et al. (2012) assigned the 1608 earthquake to the activity of the Tāléqān fault. The Ms 7.6 magnitude for this event was adopted as a source input in the probabilistic seismic risk assessment by Ghodrati Amiri et al. (2003, 2004, 2012). Archaeoseismic Data Ambraseys and Melville (1982) stated that the restoration and repair work carried out on the Damāvand congregational mosque as recorded by an inscription dated 1024 H/1615 CE (Meshkāti, 1970) may have been associated with the 1608 Alamutrud earthquake. Damāvand is located ~180 km southeast of the meizoseismal area of the 1608 earthquake, 3.5 km south of the Moshā fault (Fig. 9; Berberian and Walker, 2010; Berberian, 2014). The Damāvand congregational mosque, built in 1322, has been repaired several times (Meshkati, 1970; Shaibani, 1986, 1988, 1990) in 1416, 1521, 1615, 1621, 1670 (possibly after the 1665 earthquake), and 1831 (after the 1830 earthquake). The old mosque was demolished in 1958 to be replaced by a more up-to-date structure! At present, only the lower part of the two massive columns and the Saljuq minaret (ca. 1000–1218) remain (Matheson, 1972; Shaibāni, 1986). 1665 June 15–July 13 Damāvand Earthquake Although this is an estimated medium-magnitude (Ms ~6.5) earthquake, it is close enough to Tehran (Figs. 4, 9, and 12) that we review it here because it was part of a westward migration of seismicity along the Moshā fault culminating in an earthquake in 1830 in northeast Tehran. The town of Damāvand (35.71°N, 52.06°E) lies 60 km east of Tehran in a pleasant and fertile valley with mineral springs (Fig. 12). The 1665 earthquake was severe in Damāvand and nearby regions, where it ruined many settlements and killed a large number of people. The damages at Damāvand were repaired by the order of Shāh ‘Abbās II Safavid (r. 1642–1666) and recorded on an inscription in Kufic Arabic on a beam on the wall to the left of the mehrāb (niche in the southern wall of a mosque; mihrab [sic]) of the Damāvand congregational mosque (Sani’ al-Dauleh, 1880–1882, II.204; Nabavi, 1972; Ambraseys, 1974; Ambraseys and Melville, 1982; Berberian et al., 1985; Berberian, 1994, 2014). The effect of this earthquake in Ray and Tehran is not known. Amighpey et al. (2016, p. 169, 170) misquoted Berberian et al. (1985) and Berberian and Yeats (1999) and erroneously assigned the 1665 earthquake to the activity of the North Tehran fault, whereas,

Tehran: An earthquake time bomb this earthquakes took place on the Moshā fault (see Berberian et al., 1985, p. 230, figure 8.7, p. 231; Berberian and Yeats, figure 3, p. 125). The town of Damāvand is located 3.5 km south of the Moshā fault, and we relate this event to reactivation of that fault (Fig. 12; Berberian et al., 1985; Berberian and Yeats, 1999, 2001; Berberian, 2014). The 1665 population of Damāvand is not known; its population was 37,315 in 2011 (SCI, 2012). 1786 April 15 Spurious Ray Earthquake (?) Mansuri (1957, p. 518), allegedly translating from a dubious French edition (we have not been able to trace the alleged book by Jean Goré), wrote that: During the night of 15 Jumada II of the same year (1200 H) when Āghā Mohammad Khān [Qājār, re. 1794–1797] was asleep in one of the Karim-Khāni buildings in Tehran, an earthquake occurred and lasted a few seconds. The buildings in Tehran shook. Despite Āghā Mohammad Khān was a brave person, he was scared since he had never experienced an earthquake. The earthquake caused no damage in Tehran, and nobody killed or injured; however, few houses collapsed and few people were killed in the area where in the old city of Ray was located. (Mansuri, 1957, p. 518; translated from the Persian text) No other source records this spurious event created by Mansuri himself (Berberian et al., 1985). Zabihallāh Mansuri was not a reliable author, translator, or source. He did not have a high school diploma, and for marketing purposes of his books, he created nonexistent French writers. He also wrote about another dubious earthquake prior to devastation by Timur ca. 1384 at Ray (see above). In both cases, Mansuri tried to create a human and kind picture of brutal mass murderer dictators such as Timur (in the case of the 1384 event) and Āghā Mohammad Khān Qājār (in this case). We, therefore, delete this entry from our catalogue, preventing its entrance into the national and international catalogues. 1802 Damāvand (?) Earthquake Morier (1818, p. 355) reported a damaging earthquake in Damāvand that ruined a number of houses in Māzandarān Province, followed by numerous aftershocks. Morier’s reference to destruction of a number of villages in Māzandarān Province could possibly be a reference to Rabino (1928, p. 54, 1957), who wrote about the 1809 Ms ~6.5 earthquake at Āmol (Fig. 9). Wilson (1930, p. 117), on the authority of Morier (1818, p. 355), maintained that during the 1802 earthquake, 70 villages and towns were destroyed, and Damāvand and Semnān were heavily damaged. This is inaccurate, for it was not Morier but Watson (1866, p. 257) who mentioned the widespread destruction in connection with the 1830 earthquake (see following) at Damāvand and Semnān (Ambraseys, 1974; Nabavi, 1978; Berberian et al., 1985). 1808 December 16 Earthquake An earthquake of estimated Ms ~5.9 in northern Qazvin (Ambraseys and Melville 1982) was strongly felt at Tehran where the people rushed outdoors (Dupré, 1819, II.187; Morier, 1812, p. 254; Hedāyat, 1856, IX.289; Rabino, 1928, p. 69, 1957). Adrien Dupré (1819) wrote that: “On 16 December violent shocks were felt, each lasting 30 seconds; they devastated Kazwin [Qazvin], and other places in Mazanderan [Māzandarān province to the north], but caused little damage in Tehran” (Dupré, 1819, II.187). In the Tajrish quarter of northern Tehran, the mausoleum of Qāsem (35°48′N, 51°25′E) was damaged

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

153

and was repaired in 1809 (Meshkati, 1970). Aftershocks caused panic at Tehran. The meizoseismal area of the earthquake is not known yet; it could have happened along the Alamutrud or the Tāléqān faults (Berberian et al., 1983). No reference to this event is found in the Iranian literary sources. Djamour et al. (2012) without citing their source(s) associated this event with activation of the Tāléqān fault. Archaeoseismic Data The last line of an inscribed poem on a blue clay tile in the mausoleum of Ja’far at Pishvā (built in 1549) in southeast of Tehran (35°18′N, 51°43′E; Fig. 4), refers to a repair date with “Abjad Numerals” in 1227 H/1812 CE during the reign of Fath ‘Ali Shāh Qājār (Meshkāti, 1970). Ambraseys and Melville (1982) mentioned that the repairs at the mausoleum may have been associated with the 1808 earthquake in Tāléqān (Fig. 9). The distance between Varāmin and Tāléqān is ~130 km (Berberian, 2014). 1811 June 20 Earthquake While in Damāvand, Morier (1818, p. 355) felt an earthquake (Wilson, 1930; Nabavi, 1978; Ambraseys, 1974; Berberian et al., 1985). The earthquake could possibly be one of the aftershocks of the 1809 Āmol earthquake, south of the Caspian Sea (Fig. 9). 1815 June Earthquake Both Morier (1818, p. 355) and Stahl (1911, 4.v.6) reported an earthquake strongly felt at Damāvand, where the hot spring dried up (Ambraseys, 1974; Nabavi, 1978; Ambraseys and Melville, 1982; Berberian et al., 1985). 1830 March 27 Mw ~7.0–7.4 Lavāsānāt Earthquake Enough macroseismic data are available for this event to constrain the meizoseismal area of the most recent earthquake to heavily damage Tehran (Figs. 9 and 20; Ambraseys, 1974; Ambraseys and Melville, 1982; Berberian 1994) and locate the source fault along the central segment of the Moshā fault northeast of Tehran (Berberian et al., 1985; Berberian and Yeats, 1999, 2001; Berberian, 2005, 2014). The 1830 earthquake was centered in the Lavāsānāt district ~38 km northeast of Tehran, with damage to the east and west (Fig. 20). With its central town of Lavāsān (35°49′N, 51°46′E), the district had a population of 22,289 in the 2006 census (SCI, 2007); however, there are no data on the population of the district at the time of the 1830 earthquake. Contemporary sources state that areas east of the Jājrud River (Fig. 20) along the road to the east (35°45′N, 51°44′E, 7 km south of the Moshā fault), including the town of Damāvand (35°43′N, 52°04′E, 4 km south of the Moshā fault), and a total of 70 villages to its east and west, were devastated or damaged, some with casualties. The town of Damāvand was heavily damaged, and more than 500 people were killed there (Bell, 1840, p. 579; Watson, 1866, p. 257). In the spring of 1837, Bell (1840, p. 579) traveled from Tehran to the Caspian shore and described the ruins of the Jājrud caravanserai and Damāvand mosque as well as slight damage at Sāri (Fig. 14). The caravanserai at Jājrud was also badly damaged and partly destroyed (Conolly, 1838, I.15). Damage extended to the capital city of Tehran (Watson, 1866) (Fig. 20). Watson wrote that: The year 1830 was marked in Persia by occurrence of a series of shocks of earthquakes. In the month of April [Old Style] the town of Demevend [Damāvand] suffered severely; not less than five hundred persons are said to have [been] buried under their ruins of the houses

154

Berberian and Yeats

which were overthrown. The towns of Semnan and Damghan, and the villages in their neighbourhood, likewise sustained great injury; and in all seventy towns and villages are said to have been partially destroyed. (Watson, 1866, p. 257)

Semnān (35°24′N, 53°23′E), 163 km ESE of Lavāsān, ~180 km ESE of Tehran (MMI VII+), and 120 km ESE of the town of Damāvand (VIII+ in MMI), apparently sustained intensity around VII– (Fig. 20). Adobe buildings at Āmol (36°28′N, 52°21′E; 95 km NE of Lavāsān), Sāri (36°27′N, 52°51′E; 150 km NE), and Dāmghān (36°10′N, 54°20′E; 247 km ENE) were slightly damaged (Fig. 20). The earthquake was followed by strong aftershocks, and the one at midnight local time on 6 April (~Ms ≥ 6.0) destroyed the old caravanserai (inn) at Jājrud (35°44′N, 51°41′E; Fig. 20) and caused additional damage in the epicentral area (Conolly, 1838 [was in Tehran during the earthquake]; Mallet, 1853; Watson, 1866, p. 257; Milne, 1911; Wilson, 1930, p. 104, 119; Nabavi, 1972; Ambraseys, 1974; Ambraseys and Melville, 1982; Berberian et al., 1985; Berberian, 1994). Bell (1840), who toured the area from Tehran via Damāvand, and Talārrud Valley to Sāri, returning to Tehran through the Harāz Valley in 1839, wrote that: On this line of road [Harāz Valley road between Karoo (Kuhrud/ Kohrud) and Bulkulum (near Bāijān; 27.5 km north of the Moshā fault, see Fig. 20 in the epicentral region of the 1983 earthquake)] are many ruined bridges, and portions of broken masonry; also remains of roads, now impassable; but, in this narrow fissure, the ruins of a stone bridge were too remarkable not to attract attention. The piers, built on the solid rock of the opposite sides of the [Harāz] River, seemed as if they could never have been intended to support the same arch, so different was their parallel. Upon inquiry, I found that this, like almost all the bridges and galleries on the road, had been destroyed (and the opposite sides of the ravine had no doubt suffered displacement) by a tremendous earthquake, which occurred about eleven or twelve years before [1839], and destroyed an immense number of villages, and rendered the road totally impassable for nearly two years. On the more eastern line of route, we saw the destructive effects of the same earthquake, in the almost total ruin of the caravanserai of Jajerood [Jājrud], the ruin of a handsome mosque at the village of Demavend [Damāvand], and the destruction of other buildings, as far as Sarree [Sāri]. (Bell, 1840, p. 579)

Bell (1840) confirmed the observation of Conolly (1838, I.15) at the Jājrud caravanserai, and that of Watson (1866, p. 257) at Damāvand, and added the destruction of the mosque at Damāvand by the 1830 earthquake (Fig. 20). The 1830 Lavāsānāt earthquake destroyed or damaged almost all the adobe buildings, the bazaar, and the British Embassy compound in Tehran, and at least 30 people lost their lives in the city (Fig. 20). The chief officers of the royal court were encamped in the open courts of the damaged citadel, transacting business. The survivors at Tehran

Figure 21. “Shemirān Gate” illustration by Eugéne Flandin (Flandin and Coaste, 1851) of the Shemirān rampart and gate (Darvāzeh Shemirān; see Fig. 1 for location) northeast of ca. 1840 Tehran (courtesy of en.wikipedia.org). Note the well-developed through-going shear fractures in the structures and rampart (enhanced by arrows) portrayed 9 yr after the 1830 earthquake.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

Tehran: An earthquake time bomb

155

Figure 22. “Streets of Tehran” illustration by Eugéne Flandin (Flandin and Coste, 1851) of ca. 1840 downtown Tehran; see Figure 1 for location (courtesy of en.wikipedia.org). The picture shows the Soltāni (Shāh) mosque constructed ca. 1810–1825 in the covered bazaar of Tehran. Though in the background, the mosque and it entrance portal and dome look intact, whereas a well-developed through-going shear fracture can be seen in the corner bastion to the right, and a collapsed section of the rampart is seen to its immediate right (enhanced by arrows).

lived outdoors, and business was consequently at a standstill (Conolly, 1838, I.15; Wright, 1977, p. 26). During the time of the 1830 earthquake, Tehran was a small town with a population estimated around 65,000, covering an area ~4 km2. The population estimates of Tehran are 50,000 in the early 1800s; 85,000 in 1867; and 90,000 in late 1800s, with the total population of the country estimated at ~4 million, which grew to 9.86 million in 1900 (Porter, 1821; Minorsky, 1934; Issawi, 1971; Ladier, 1993; Hourcade, 1994; Planhol, 2004). Except for the King’s palace and some mosques like Soltāni (Shāh) in the bazaar, most of the buildings of the city were traditional adobe masonry structures of unreinforced rural rubble or sun-dried clay brick with poor workmanship. Augène Flandin and Pascal Coste (Flandin and Coste, 1851) arrived in Tehran 9 yr after the 1830 earthquake and stayed in Iran until 1841. While residing 23 d in Tehran, they wrote that some sun-dried mud brick houses collapsed during rains. They did not write about the 1830 earthquake and its effect at Tehran; however, some of their illustrations give some information about the building quality, style, and fracturing of some structures, presumably from the 1830 earthquake, which provides an estimate of the 1830 earthquake intensity in Tehran. On the “Shemirān Gate” illustration (Flandin and Coste, 1851), located to the northeastern section of the city rampart (see the 1876

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

outbound of the city in Fig. 2), well-developed through-going shear fractures are clearly visible cutting: (1) piers of a four-arch pavilion (“Chahār-Tāq”; an open-sided pavilion with four pillars supporting a domed roof, traditionally holding the Zoroastrian sacred fire); (2) piers of an adjacent vaulted structure to the left of the former; (3) two perpendicular walls with collapsed tops, and the remains of the lower part of another wall near the tower; (4) a windcatcher tower (bādgir) at the corner of the two mentioned walls; and (5) the wall to the right of the Shemirān Gate (Fig. 21). Another illustration titled “Streets of Tehran” shows the intact Soltāni (Shāh) mosque of the bazaar of Tehran (built in 1810–1825 during the time of Fath ‘Ali Shāh Qājār; r. 1797–1834) standing intact in the background. However, to the right side of the domed roofs of the bazaar, a corner bastion with a well-developed through-going shear fracture and a ruined wall connected to its right are shown (Fig. 22). Other illustrations of Tehran do not show any structural damage from the 1830 earthquake. A perspective distant view of the four-story “Qājār Palace” (Qasr-e Qājār), as well as the interior views; the “brick kiosk/gazebo” at the palace garden; the interior and exterior views of the Golestān Palace with the marble throne room; the Bārut-Khāneh (ammunition storage building near Tehran); the two-story DivānKhāneh building (Court House); and buildings in the Shāh Square

156

Berberian and Yeats

Figure 23. Historic photograph of the city of Tehran taken in 1906 by Sven Hedin (Hedin, 1910), giving a glimpse of the quality of the buildings in Tehran 76 yr after the 1830 earthquake (see Fig. 1 for location). The snow-covered Alborz Mountains are seen in the background (courtesy of the Sven Hedin Foundation, Stockholm).

(Maidān), with a two-story building, all seem intact with no major structural damage or failure. At Ray (south of Tehran), the Cheshmeh ‘Ali vault and Shāh ‘Abdol‘azim shrine, as well as the Solaymānieh mausoleum and the adjacent bridge at Karaj look intact without any major damage (see illustrations in Flandin and Coste, 1851). The photograph taken in 1906 by Sven Hedin also gives a glimpse at the quality of the buildings in Tehran 76 yr after the 1830 earthquake (Hedin, 1910; Fig. 23). These structures could have been easily damaged by an earthquake of intensity VI+–VII– (MMI). Throughout 1830–1831, Iran was suffering from diseases such as cholera, typhoid fever, smallpox, and malaria, followed by a quickly spreading plague. It was the most devastating plague epidemic in the country in 1830–1831, with the exception of the Kermān and Khorāssān Provinces in the east, killing many people. The northern provinces of Guilān and Māzandarān (north of the 1830 epicentral area, south of the Caspian Sea) suffered most, and apparently ~20,000 people died in the capital city of Tehran (Fraser, 1838; Tholozan, 1882; Lorini, 1900; Seyf, 1989; Kohn, 2007; Shahraki et al., 2016). As a result, the population of the country was reduced significantly, with prolonged destructive economic consequences. Although lack of economic growth, health facilities, a sanitary sewage system, polluted drinking water, absence of any public health administration and of quarantine procedures, and the religious practice of washing dead bodies prior to burial by running water were the main causes of the rapid spread of epidemics, the 1830 Mw ~7.0–7.4 Lavāsānāt earthquake contributed to the epidemics. Nonetheless, the exact time, location, and source of the 1830–1831 plague are not clear. Based on the available data, Berberian et al. (1985), and Berberian and Yeats (1999, 2001) identified the central segment of the Moshā fault north and northeast of Tehran as the source of the 1830

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

earthquake (Figs. 9 and 20). The city of Tehran is located 28 km south of the Moshā fault. As mentioned earlier, average cross-fault attenuation characteristics from five Mw 7.0–7.3 earthquakes in Iran (1968, 1978, 1981, 1990, and 1997; despite some differences in their source parameters) show that the average distance of isoseismal lines from coseismic surface ruptures was 5 ± 1 km (for IX MMI), 15 ± 2 km (VIII), 30 ± 5 km (VII+), and 45 ± 10 km (VI+). Hence, the 1830 earthquake along the Moshā fault might have caused intensity of ~VII+ at Tehran (Fig. 20). This estimation is compatible with damages reported in the city of Tehran (Conolly, 1838, I.15; Watson, 1866, p. 257) and with historic illustrations (Flandin and Coste, 1851). This observation, together with destruction of Damāvand and Jājrud (>VIII), and the estimated magnitude of the earthquake indicate that the 1830 destruction should have possibly extended to the NW of Tehran along the Moshā fault to its northwestern tip. Therefore, the destroyed villages from east of Damāvand to the north and probably northwest of Tehran cover a distance exceeding 80 km along the Moshā fault (Figs. 9 and 20), which seems compatible with a ~70–80 km coseismic surface rupture with Ms ~7.1 along the Moshā fault. About 7.5 km northwest of Damāvand at the Āb-e ‘Ali ski resort (Fig. 12), the Moshā fault cuts young colluvial sediments dated 1.21 ± 0.15 ka (Solaymani Azad et al., 2011). Those authors considered the paleoseismic rupture could possibly correspond to the 1665 or 1830 earthquake, and they preferred the association with the latter event. They stated that the 1830 earthquake source was closer to Tehran than proposed earlier. This cannot be warranted since, as mentioned above, (1) the main damage was to the east of Jājrud (Fig. 20), (2) more than 500 people were killed at Damavand (Watson, 1866, p. 257), and (3) historical illustrations do not show destruction of buildings (Flandin and Coste, 1851).

Tehran: An earthquake time bomb Fracturing (“faulting 24–28” in Ghassemi et al., 2014) associated with landsliding in the Irā trench across the Moshā fault near its junction with the North Tehran fault system (Fig. 4) cut sediment units 23, 26, 29, and 30 as fractures developed in the breakaway zone of the Irā landslide. The “F24 fault” in the trench (Fig. 4), with 2.3 m slip, is probably associated with historic earthquakes; however, no independent dating was conducted for the 1830 or 1930 events in the trench (Ghassemi et al., 2014; Angela Landgraf, 22 April 2014, October 2015, personal commun.). Data from two reliable contemporary sources, and landslides in the area (though not dated) may indicate reactivation of the central Moshā fault during the 1830 earthquake (Figs. 4 and 20). Although the assigned equivalent of Ms ~7.1 magnitude (Ambraseys and Melville, 1982) is a reasonable assessment, it might possibly be slightly underestimated. We believe that the 1830 earthquake, with ~80 km surface rupture and destruction, was possibly comparable to the 20 June 1990 Mw 7.3 Rudbār earthquake (both with left-lateral strike-slip mechanisms; Fig. 9). Aside from the older destructive earthquakes in the Ray-Tehran region with unknown sources (Table 2), the 958 Ruyān and 1830 Lavāsānāt earthquakes, along the central segment of the Moshā fault (Figs. 19 and 20), were the closest large-magnitude earthquakes to Tehran, with 872 yr return period, and had the greatest bearing on the earthquake hazard to the city. These earthquakes on the Moshā fault might have loaded the North Tehran fault system (Fig. 20). It should be noted that the Latyān dam supplying water to Tehran is located ~7 km south of the Moshā fault and 4 km south of the North Tehran fault system (Fig. 4). Utsu (2002, 2014) introduced two entries on 27 March 1830 (in “Damavand-Shamiranat, Tehran,” M 7.1, intensity IX–X, and 530 killed), and on 9 May 1830 (in “Damavand-Shamiranat, Tehran, Damghan,” with 500 dead); the date of the latter cannot be warranted, and we delete the event from our catalogue. Despite the availability of macroseismic data for the 1830 earthquake, Asadi and Zare (2014) reported two 1830 earthquakes: one with an epicenter 45 km southeast of Damāvand, and the second epicenter ~60 km northeast of Damāvand (see figure 4 in Asadi and Zare, 2014). Amighpey et al. (2016, p. 169, 170) misquoted Berberian et al. (1985) and Berberian and Yeats (1999) and erroneously assigned the 1830 earthquake to the activity of the North Tehran fault, whereas, this earthquakes took place on the Moshā fault (see Berberian et al., 1985, p. 233, figure 8.8, p. 234; Berberian and Yeats, 1999, figure 3, p. 125). Archaeoseismic Data The dome of the Zaid mausoleum in the bazaar of Tehran was reconstructed in the late 1830s during the reign of Faith-‘Ali Shāh Qājār. The tile inscription above the entrance recording the renovation is dated 1245 H/1830 CE (Makinezhad, 2006). This may indicate that the dome structure of this old mausoleum was damaged by the 1830 earthquake. The effects of the 1830 earthquake on other monuments in Tehran, Shemirān, Ray, Lavāsān, Sālehābād, Rudak-e Rudbārān-e Qasrān, Beryānak, and other locations in the greater Tehran-Ray and Ruyān region have not yet been studied (Berberian et al., 1985). An inscription in the Qom congregational mosque (Fig. 20), built in 1135, refers to repairs carried out in 1246 H/1830–1831 CE and 1248 H/1832–1833 CE (Meshkati, 1970), which might have been related to the 1830 earthquake. Ambraseys and Melville (1982) mentioned only the 1832 repairs. The city of Qom (34°38′N, 50°52′E; Fig. 20) is located ~128 km to the SSW of the city of Tehrān and ~153 km to the SSW of the Moshā fault (Berberian and Yeats, 1999, 2001; Berberian, 2014). 1830 April 6 Ms ~>6.0 Strong Lavāsānāt Aftershock Lieutenant Arthur Conolly (1838, I.15) wrote that:

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

157

While we were endeavouring to find some merchant who would arrange our affair, the city was visited by a severe earthquake, and business was consequently at a stand… On the 6th of April [1830] we took our departure from the capital, and rode out fifteen miles easterly to Jujjer-road [Jājrud], a rapid stream in a narrow barren glen, where we forded with some difficulty, by reason of the loose stones in its bed. On the other bank was a brick caravanserai, solitary instance of the reigning monarch’s extravagance. As part of this building appeared to have lately fallen, we did not venture to lodge in any of its cells, but spread our beds in the centre of the square; well was it that we did so, for about midnight we were awakened by a heavy rumbling sound that seemed to pass under the ground we lay upon, and, starting up in alarm, we saw the bricks of the different apartments falling all round us. (Conolly, 1838, I.15) The ruined caravanserai of Jājrud (Fig. 20) was also visited in 1839 by Bell (1840, p. 579). This strong aftershock, which destroyed the caravanserai at Jājrud (Fig. 20), caused panic at Tehran and added damage to the already damaged poor-quality buildings (Ambraseys and Melville, 1982; Berberian et al., 1985). 1847 September 7 Earthquake An earthquake at 01:25 a.m. local time was felt in Tehran (Perrey, 1848, p. 451; Ambraseys, 1974; Nabavi, 1978; Berberian et al., 1985). The source and magnitude of this event is not known. 1850 Earthquake In the diary kept during her sojourn in Iran between 1849 and 1852, Lady Mary Leonora Woulfe Sheil (1803–1871), knight general and diplomat and wife of the British ambassador in Tehran, wrote that: A year afterwards [1850], in Tehran, while lying ill in an upper room, I felt a curious sensation, like the shaking of steamboat. I rushed out of the room, down the stairs; for I suspected what it was, and feared a repetition of it. There was, however, only one shock; and I never felt any other during my stay in Persia. (Sheil, 1856, p. 90–91) This is the only surviving record of this earthquake being felt in Tehran and was not included in the existing earthquake catalogues (Ambraseys and Melville, 1982; Berberian et al., 1985; Berberian, 2014). Based on this single report, Vasheghani Farahani et al. (2014) in their figure 10 drew an isoseismal of VI covering the southern Tehran and Ray area, which cannot be warranted. 1853 September 5 A rather strong earthquake was felt in Tehran 10.5 h after the night (possibly after sunset) with no damage in the night of 1 Dhu’l Hijja 1269 (Vaqāye’ Etefāiqiyeh Newspaper, Thursday, 4 Dhu’l Hijja 1269 [8 September 1853], no. 136, p. 854). 1890 July 11 Ms 7.2 Tāsh Earthquake Although the epicentral region of this destructive earthquake was located ~300 km northeast of Tehran in the eastern Alborz Mountains, southeast of the Caspian Sea (Fig. 9), causing sea waves (seiche) along the southern shore of the Caspian Sea (Mushketov, 1891, p. 5, 6, 46, 58; Mushketov and Orlov, 1893; Ambraseys, 1974; Ambraseys and Melville, 1982; Berberian, 1994), we include it in our study as another example of the occurrence of large-magnitude earthquakes in the Alborz Mountains.

158

Berberian and Yeats

1895 December 24 Earthquake Without quoting any authority, Lysakowski (1906, p. 48; 1910, p. 47) mentioned a strong earthquake in Tehran killing many horses and other animals in the city (Ambraseys, 1974; Berberian et al., 1985). No other sources for this event have yet been found. Nazari et al. (2005) drew an ellipse representing the meizoseismal area of the 1895 earthquake along the Ray fault on their map, for which there are no supporting macroseismic data. APPENDIX C. COMMENTARY ON THE INSTRUMENTAL PERIOD (1900–2015) EARTHQUAKES OF THE GREATER TEHRAN-RAY REGION DISCUSSED IN THIS PAPER, WITH LOCATIONS OF MEIZOSEISMAL AREAS AND SEISMIC SOURCES The mean epicentral location errors of the Iranian large-magnitude earthquakes range from 300 km (ca. 1918) to 30 km (ca. 1963). The error decreases progressively between 1918 and 1963. Commencement of operation of WWSSN (World-Wide Standard Seismographic Network, established in 1963) produced improvement, and the error gradually decreased from 30 km (ca. 1963) to 15 and 10 km (ca. 1977) for large-magnitude earthquakes, presumably as the number of operational stations increased. Nonetheless, for the medium- to small-magnitude earthquakes, the amount of location error is much more. Errors in the focal depth estimates have been more than those of the epicenters. It seems that factors such as seismological station distribution, earth models used in location programs, timing error of the instruments, and finally reading errors of the seismograms were presumably the cause of the error in instrumental epicenter location of the Iranian earthquakes. Using or interpreting the Iranian early teleseismic data by preparing epicenter and seismotectonic maps has already shown that there is a limit to the use of the instrumentally located and relocated epicenters of the Iranian earthquakes with the existing data. Therefore, at this stage the macroseismic information is invaluable for the seismotectonic studies in Iran, where international data are not always sufficient in quantity or accuracy to locate active faults (Berberian, 1979a).

2 people were killed. Eight houses were destroyed and the rest damaged at Mobārakābād. In Ahmadābād, where the Moshā fault is covered by an earthquake-triggered landslide and rock avalanche, a few people lost their lives, the qanāt water flow was blocked, and a rock avalanche blocked streams and roads. In the nearby spa of Cheshmeh ‘Alā (between Moshā and Damāvand villages, see figure 8.10, p. 240 in Berberian et al., 1985), the bottling factory was damaged beyond repair. The shock did not cause any change in the water flow of the hot springs. In the Irā-rud valley, the shock ruined a few houses in Ārdineh and caused the collapse of walls at Sang Darvāzeh. Some houses were slightly damaged at Irā, Moshā, Dasht-e Mazār, and Damāvand (Fig. 12). The main shock and its strong aftershock of 7 October (M 5.0) caused additional damage in the Irā-rud valley and triggered landslides from the north banks of the stream between Javārd and Irā, blocking streams and roads. The earthquake was felt in Tehran, where a few qanāt lines collapsed. The shock was also felt at Firuzkuh, Garmsār, Karaj, Kāshān, and Natanz (Ettela’āt, 1930; Kaihān, October 1930; Kushesh, 1930; Tchalenko, 1974; Berberian et al., 1985). Despite its small magnitude, the well-constrained meizoseismal area of the 1930 Āh earthquake aligns with the Moshā fault (Berberian et al., 1985), which did not show any evidence of coseismic surface rupture (Fig. 12). This was the first earthquake of the instrumental period occurring along the central segment of the Moshā fault. Earlier, the 985 Mw ≥~7.0 Ruyān, 1665 Mw ~6.5 Damāvand, and 1830 Mw ~7.0–7.3 Lavāsānāt earthquakes took place along the same fault (Figs. 9, 19, and 20). The instrumental relocation of the event (Nowroozi, 1976) shows a focal depth of 70 km, which is inconsistent with the observed macroseismic effects of the earthquake and seismicity of the Alborz (Berberian, 1979b). 1930 October 7 Āh Aftershocks Two aftershocks at 00:23 h (20:53 UTC, M 5.0) and 01:30 h local time were strongly felt in Tehran, where people rushed outdoors. The first shock was felt in Shemirān (north of Tehran at the mountain foot), Qom, Damāvand, and Firuzkuh (Figs. 12 and 20). In Damāvand, a few damaged houses collapsed. These two aftershocks were followed by several other earthquakes till morning (Ettela’āt, 9 October 1930; Berberian et al., 1985).

1937 April 7 Spurious Earthquake 1947 September 5 Lavāsān Vasheghani Farahani et al. (2014) introduced an earthquake of M 5.25 and intensity of VIII (EMS98) on 7 April 1937 in the area south of Tehran at the edge of the Daryāyeh Namak (salt playa) damaging the qanāts between the villages of Qale’h Nau and Hessār Qoli. The authors made a mistake and described the effects of the 22 July 1927 Ms ~6.3 Central Kavir (desert) earthquake (Ambraseys and Melville, 1982; Berberian et al., 1985). The locations of the meizoseismal area and macroseismic epicenter of the 1927 earthquake are not known, since it took place in the uninhabited desert salt playa with very little sedentary population. Furthermore, it is not clear how an event of M 5.25 in Vasheghani Farahani et al. (2014) can create an intensity of VIII EMS98 (European Macroseismic Scale)? 1930 October 2 Āh Earthquake A century after the 1830 Ms ~7.1 Lavāsānāt earthquake along the central segment of the Moshā fault (Figs. 9 and 12), an earthquake of Ms 5.2 ruined the small district of Āh and damaged a number of neighboring villages in the area northwest of Damāvand village (SW of the Damāvand volcano) along the same fault segment (Fig. 12). At Āh Bālā Kushki village, 28 houses collapsed, one person was killed, and five were injured. Five houses were destroyed at Qābus (Kāvus) Mahalleh. Almost all the houses at Sarpurak were ruined, and

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

An earthquake at 09:45 h local time caused damage to poorly built houses in Lavāsān, northeast of Tehran (Fig. 20). The shock lasted for 5 min (Bozorgnia, 1962). The earthquake has no instrumental data. 1947 October 12 Damāvand Region An earthquake at 22:30 h (local time October 11) was strongly felt in Damāvand (Fig. 12; Iran Newspaper 1326.07.20/1947.12.13; Bozorgnia, 1962). 1955 November 24 Moshā Earthquake A small-magnitude earthquake with no reported teleseismic data ruined a house and damaged the others in Moshā (Bozorgnia, 1962; Nabavi, 1972; Tchalenko, 1974; Berberian et al., 1985). The village is located on the central segment of the Moshā fault (Fig. 12). 1957 July 2 Mw 7.1 Band-e Pay Earthquake In our search for the MCE for the Tehran region, we report the 1957 earthquake here as the second largest magnitude (Mw 7.1) event

Tehran: An earthquake time bomb (after the 1990 Mw 7.3 Rudbār earthquake) during the instrumental period occurred in the Alborz, 100 km northeast of Tehran (Fig. 9). The 1957 earthquake totally destroyed more than 120 villages in the eastern Alborz Mountains. It was followed by typhoid fever and influenza epidemics in the epicentral area, and damages estimated at about US$25 million original value. Terrified people in Tehran rushed outdoors, and a few houses were damaged in southern Tehran (VI, MMI; Ettela’āt Newspaper, 4–18 July 1957; New York Times, 3–17 July 1957; Āsiā-ye Javān Magazine, 1957, v. 7, no. 361, p. 3; Khāndanihā Magazine, 1957, p. 87–88; Savage, 1957; Hagiwara and Naito, 1959; Bozorgnia, 1962; Rothé, 1969; Nabavi, 1972; Tchalenko, 1974; Ambraseys and Melville, 1982). 1970 October 3 North Tehran Earthquake An earthquake of mb 4 was strongly felt in Tehran; terrified people rushed outdoors. It caused slight damage at Qasrān-e Rudbār, north of Tehran (Inner Qasrān in Fig. 19), where the shock also caused panic (MMI V). It was less intense at Rud-e Hen and Tehran (IV), where the electricity was cut off, and the traffic became gridlocked. The shock was intense at the northern quarters of Tehran in Shemirān and Nārmak, as well as Mehrābād, Āriyāshahr, and Ray (IV) to the south. A few old houses cracked at Nārmak (northeastern quarter of Tehran) and Damāvand village. It was also felt at Āb-e ‘Ali, Polur, Firuzkuh, and slightly at Karaj (III; Ettela’āt Newspaper, 3–5 October 1970, Kayhān Newspaper, 4–10 October 1970; Nabavi, 1972; MoazamiGoudarzi, 1972; Berberian, 1977; Berberian et al., 1985). The meizoseismal area of this earthquake covers the area west of the Moshā–North Tehran fault junction (Fig. 20; Berberian et al., 1985). The ISC (with 13 observations) and the U.S. Coast and Geodetic Survey (with seven observations) epicenters are located in the vicinity and to the north of the Moshā fault, but the location errors are not known (Fig. 8). 1999 June 20 Rudbār Earthquake The M w 7.3 Rudb ā r earthquake took place 220 km northwest of Tehran (Fig. 9) and was the largest documented earthquake in the Alborz Mountains during the instrumental period (Berberian et al., 1992; Berberian and Walker, 2010; Berberian, 2014). We, therefore, consider using the 1990 Rudb ā r earthquake magnitude as the maximum credible earthquake (MCE) hazard to the city of Tehran. The 1990 Mw 7.3 Rudbār earthquake took place in one of the most urban and agriculturally developed rural regions of Iran in an area southwest of the Caspian Sea (Fig. 9). The event took place near midnight at 00:30 h local time (21:00 GMT), contributing significantly to the large loss of life and a day delay in nonorganized and chaotic rescue operations. The national Islamic Republic News Agency (IRNA) reported that rescue teams had found 28,950 dead bodies by 23 June 1990. On 28 July 1990, the official estimation of losses was 40,000 dead, 60,000 injured, and 500,000 homeless (UNDRO, 1990a, 1990b). The reported official casualty figures ranged from 40,000 dead and 105,000 injured in 1990 (UNDRO, 1990a; Berberian et al., 1992; Berberian and Walker, 2010) to 13,000 dead and 60,000 injured reported a few years later (Ghafory-Āshtiāny and Eslāmi, 1997), a death toll that we believe is too low considering that the first author studied the area immediately after the earthquake and visited the meizoseismal area (Berberian et al., 1992; Berberian and Walker, 2010; Berberian, 2014). The earthquake demolished the towns of Rudbār, Manjil, Harzéhvil, and Lowshān. It also damaged the provincial capital city of Rasht, located 50 km NNE of the coseismic surface fault; it destroyed

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084737/spe525-04.pdf by University of Wisconsin Madison user

159

700 villages in total, demolishing more than 100,000 houses, as well as the underlying economic infrastructure of two provinces of Guilān and Zanjān in the stricken area of at least 5500 km2 covering MMI intensity zones of IX+ to VIII. About 300 additional villages and many towns sustained some damage. Almost 1300 schools with 7000 classrooms in two provinces were destroyed or damaged beyond repair. The Rudbār two-story hospital, built three years before the earthquake in 1987, and the Rostamābād hospital, together with 85 additional health-care centers and hospitals, were demolished in two provinces. Direct economic loss was estimated to be US$7.2 billion, ~2.5% of gross domestic product (GPD; UNDRO, 1990a, 1990b, 1990c; Astāneh and Ghafory-Ashtiany, 1990; Underwood, 1991; Berberian et al., 1992; Berz, 1992). If such an earthquake occurs in Tehran, the destruction and damage will be unfathomable due to its demography. The Rudbār earthquake source fault was not identified prior to the 1990 earthquake, and additional, unmapped, active faults may be present within the Alborz Mountains and in Tehran. Three discontinuous fault segments were mapped along a previously unrecognized fault, 80 km long, north of Rudbār city. The subvertical fault has a strike of N95°120°E with principal left-lateral strike-slip motion (Fig. 9). Maximum observed surface displacements were 100 cm left-lateral and 120 cm vertical (Berberian et al., 1992; Berberian and Walker, 2010; Berberian, 2014). Assuming a fault length of 80 km, a depth of 15 km, and rigidity of 3 × 1010 Nm–2, the calculated seismic moment accounts for an average seismologically determined slip of 240 cm, which is much more than the observed displacements in the field (Fig. 9). This discrepancy could be due to the following: (1) The fault and its displacements were not thoroughly mapped immediately after the earthquake because of the high mountainous area of the High Alborz; (2) the rupture might have propagated deeper than 15 km, although microseismicity data of Tatar and Hatzfeld (2008) suggest that most seismic deformation is limited to 1300 yr history of the Borujerd old congregational mosque by means of scrutinizing ancient historical chronicles, archaeology, architecture, historical and modern seismicity, geology, and regional active faults. Our study resulted in recognition of at least four major phases of destruction, rebuilding, and renovations of the megastructure located 4 km to the northeast of the Zāgros Main Recent fault in western Iran. The grand structure shows significant paleo-architectural and archaeological evidence of destruction and damage. Although some damage events, recorded in seven phases since the seventh century C.E., could have been due to the poor construction of early periods and their decay, there is strong evidence of at least one extensive, simultaneous, and abrupt destruction and damage pattern of mosque III (ca. post–1090/pre–1139 C.E.) in the early fourteenth century. We suggest that the poorly known 1316 C.E. strong earthquake (which destroyed more than 20 villages in the general area, with erroneous epicentral location in the historical seismic catalogues) was possibly responsible for the simultaneous sudden collapse of the Borujerd congregational mosque lofty dome chamber and its tall free-standing minaret; we infer that this earthquake occurred with intensity >VIII+ (modified Mercalli intensity scale) conceivably along a seismic gap zone of the Zāgros Main Recent fault. No pre–1316 C.E. monument exists in the epicentral region, and no strong earthquake has occurred along that segment of the Zāgros Main

*E-mails: [email protected] and [email protected]; [email protected]. Berberian, M., Moqaddas, M., and Kabiri, A., 2016, Archaeological and architectural evidence of historical seismic activity along the Zāgros Main Recent fault at Borujerd (western Iranian Plateau); the 1316 C.E. earthquake, in Sorkhabi, R., ed., Tectonic Evolution, Collision, and Seismicity of Southwest Asia: In Honor of Manuel Berberian’s Forty-Five Years of Research Contributions: Geological Society of America Special Paper 525, p. 171–212, doi:10.1130/2016.2525(05). © 2016 The Geological Society of America. All rights reserved. For permission to copy, contact [email protected].

171

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084761/spe525-05.pdf by University of Wisconsin Madison user

172

Berberian et al. Recent fault for the last seven centuries. Retrospectively, apparent indigenous paleoarchitectural renovations were utilized during construction of the new congregational mosque (mosque IV: ca. post–1405/pre–1447 C.E.) to enhance the coherency and elasticity of the rigid brick structure to withstand future earthquake shear stress. The hazard-reducing efforts included: (1) retrofitting the surviving load-bearing structural elements; (2) avoiding grandeur and majesty and implementing simplicity by reducing the size, height, and shape of the dome chamber; (3) avoiding free-standing minarets; (4) minimizing the size and reducing the light/ventilation openings; and (5) utilizing several levels of timber bracings to neutralize earthquake strong ground motion. Our research reveals that the return period of large-magnitude earthquakes along the two major segments of the fault is in the range of 1000 and 2000 yr, thus making historical earthquakes unrecognizable through routine historical research. It also shows how the use of archaeoseismology and paleo-architectural investigations on deformed monuments may improve our knowledge of long-term seismicity and seismic hazards of a region. This kind of study permits us to hypothesize the occurrence of strong earthquakes in an area for which historical seismicity does not show significant earthquakes. Finally, based on the described historic seismic damage and destruction, the regional national monuments should be properly retrofitted to withstand future earthquake hazards.

INTRODUCTION1 1

Understanding the impact of large-magnitude earthquakes and the intervening destructive episodes of historical and ancient cultural heritage sites located in earthquake-prone areas near active faults is of scientific, architectural, engineering, cultural, and social importance. The obtained knowledge from these preserved ancient “archaeo-seismograms” or “archaeoaccelerometers” (Sintubin, 2011; Sintubin et al., 2010) can: (1) strengthen and extend the prehistoric and historical seismic record; (2) show the extent of meizoseismal areas of ancient and historical earthquakes; (3) reveal the long-term pattern of seismicity and recurrence periods of major earthquakes; (4) unveil the maximum credible earthquake; (5) provide crucial data on earthquake-fault hazard assessments, risk analyses, and management; (6) deliver information on vulnerability or strength of historical monuments; and (7) help to develop retrofitting techniques with accurate stress distribution in the monuments and other megastructures in order to withstand future earthquakes. Archaeoseismological studies are especially important in areas such as the Iranian Plateau, where ~45,000 archaeological sites and monuments exist (CHTHO, 2011); written old documents are scarce due to numerous devastating invasions; meizoseismal areas of some historical earthquakes are either unknown, misallocated, or unconstrained; the historical seismic record is incomplete and heterogeneous (Ambraseys and Melville, 1982; Berberian, 1994, 2014); and archaeoseismological and paleoseismological trench studies are in their infancy in the country (Berberian et al., 2012, 2014; Berberian, 2014). Limited archaeological excavations below the present floor of the old Borujerd congregational mosque (Fig. 1) as well as

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084761/spe525-05.pdf by University of Wisconsin Madison user

All dates have been converted to the New Style/Gregorian calendar dating system used in the West (C.E./A.D.). In order to assist location of accounts in ancient Iranian and Arabic sources, Arabic Hijra lunar (H.; pre–1900 C.E.) and Iranian solar (post–1900 C.E.) calendar years are added throughout the text, tables, and the reference list. Where the conversion is given, the Arabic lunar and/or Iranian solar years are separated by a forward slash from the Christian year [1292/1875], where obviously the latter has a larger number. We tried to substantiate some of the earthquakes with historic chronicles; hence, tables are designed to facilitate understanding the lengthy chronology of reported events in the region as well as constrain the location of the earthquakes in space and time. Volume and page numbers of some available chronicles are shown as II:491 (i.e., volume II, page 491). Note that in this report, Persian geographical names and other Persian words are written as they are correctly pronounced and written originally, with direct and simplified transliteration from Persian into English. Diacritical marks and special characters are used to differentiate vowels “A” (short; e.g., ant) from “Ā” (long; e.g., Ārmenia or Irān), and Arabic “ain” (used also in Persian) as “’A” (e.g., ‘Abbās). Well-known names such as Iran (Irān) and Tehran (Tehrān) are used without diacritical marks. The recognition of the Persian possessive (afzudeh; or ezāfé in Arabic), which inaccurately appears variously in English and French literature, especially as “-i” (thus: Shahr-i Kurd [sic]) is correctly shown as “-e” (cf. French “é”; thus: Shahr-e Kord), as conforms to the correct current usage in the Persian language (Fārsi). We used the correct spelling of the Iranian terms of “ayvān” (porch), “mehrāb” (niche), and “menbar” (pulpit) instead of their misspelled and incorrect usages of “ivan/iwan,” “mihrab,” and “minbar” in the literature. The name of the city of “Borujerd” has been misspelled as “Burujird” in the European-language literature. Elevations are given in meters above mean sea level; the archaeological stratigraphic horizons are given in centimeters below the present dome chamber floor of the Borujerd congregational mosque. Coordinates of the sites discussed are given in the text and/or tables. Some of the photographs used in this study were taken 30–20 yr ago and are of low quality, and efforts to trace the invaluable prints or negatives failed at the Iranian Cultural Heritage Organizations in Tehrān and in the Borujerd office. Because of their important archaeoseismic value, which were destroyed by reconstruction activities, we decided to use a few irreplaceable unique photographs, though with lower resolution, in this report. Whenever possible, the Persian translation editions of the documents utilized in this work are included in the reference list for the Iranian scholars and students who have not had easy access to the foreign publications during the past three decades.

Evidence of historical seismic activity along the Zāgros Main Recent fault at Borujerd on the megastructure itself during the 1980s revealed several major phases of destruction, reconstruction, essential repairs, and restorations (Mehryār, 1985; Moqaddas, 1997). Both the project archaeologist (Moqqadas) and architect (Mehryār) independently concluded that the high dome, upper portions of the thick

173

load-bearing walls, and the minaret of the mighty congregational mosque most probably collapsed in an early fourteenth-century strong earthquake (Mehryār, 1985; Moqaddas, 1997). However, they could not assign the destruction to any dated earthquake because no earthquake was listed in the Iranian catalogues occur-

Figure 1. Major active faults of Iran showing the Borujerd study area in western Iran (filled circle) along the Zāgros Main Recent right-lateral strike-slip fault (Zāgros Recent on the figure adjacent to the Zāgros Main Recent fault). Fault names are in italics. Reverse faults are shown with teeth on hanging wall. Strike-slip faults are shown with arrows. Faults without teeth or arrows—sense of slip unknown. ± indicates relative vertical motion. Inset top right: Map of Iran showing boundary with Arabian plate (line with teeth). AZ—Āzarbāijān; KB—Karehbas fault in the Zāgros, SW Iran; KP—Kopeh Dāgh; M—Makrān; S—Sistān suture zone; S—Sarvestān fault in SW Iran; SP—Sabz Pushān right-lateral shear zone in the SW; TA—Tabriz; TP—Turān plate. Rigid blocks are cross-hatched. Figure is modified from Berberian (1976, 1977, 1981, 1995, 2005, 2014) and Berberian and Yeats (1999, 2001).

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084761/spe525-05.pdf by University of Wisconsin Madison user

174

Berberian et al.

ring in the city of Borujerd (Ambraseys and Melville, 1982; Berberian, 1994, 1995, 2014). The excavations carried out at the site were haphazard and limited in scope, and the observations presented here are based on archaeological deposits exposed in small soundings below the crawl space with no radiometric dating, and architectural indicators of different phases detected from the structure. The 1980s investigations were not systematic or conclusive, and archaeoseismological investigation was not included as a project task. Since then, no more permission has been granted for further investigation. The monument has undergone numerous substantial rebuilding phases since the ninth century C.E., in the course of which typical elements of construction, architecture, and decorations, which would have been valuable in dating, have been mixed together or vanished. Therefore, some issues described in the two original archaeological reports (Mehryār, 1985; Moqaddas, 1997) are still clouded, and certain dates are

not very precise or constrained. Hence, further research is needed to improve the quality of the data and constrain the dates and the events described here. Nevertheless, despite the obvious limitations of the available archaeological record, we try to shed some light on major events that might have affected the monuments in this region. The intention of the present work is, therefore, to utilize the surviving limited archaeological, architectural, and historical accounts; search for architectural evidence with respect to possible earthquake damage and destruction; and contribute to the existing data sets on earthquakes, active faulting, and their hazards around the city of Borujerd located along the Zāgros Main Recent active fault in the Lorestān province of the western Iranian Plateau (Fig. 2). In the following sections, we try to organize and present the reported archaeological and architectural observations, and, with the help of historical, seismicity, and active fault data, search for probable earthquakes and

Figure 2. Location of Borujerd (large filled circle) along the Zāgros Main Recent fault in the framework of the medieval period provinces of western Iran (discussed in the text) and locations addressed in the text. The underlined cities are reported as being the epicentral area of the earthquake in historical documents (see the text). Dashed line—medieval province boundary. AZ—Āzarbāijān; KD—Kopeh Dāgh; M—Makrān; S—Sistān suture zone.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084761/spe525-05.pdf by University of Wisconsin Madison user

Evidence of historical seismic activity along the Zāgros Main Recent fault at Borujerd possibly locate the causative seismic source for the 1316 C.E. strong earthquake. Archaeoseismological indicators, preserved in archaeological records (settlement layers of archaeological mounds or on the structural elements of structures), are usually difficult to determine explicitly. Clear archaeological and architectural evidence of earthquake strong ground motion is sufficient to conclude that damage, destruction, and deformation were indeed earthquakeinduced and not a result of: (1) anthropogenic action (damage and destruction [this usually does not apply to the congregational mosques], restorations, and modifications of structures); (2) meteorological processes, such as rainstorms, erosion, leaching, or weathering of archaeological deposits and monuments; or (3) uneven settlement of the structure due to foundation subsystem, site-specific geology, and geotechnical characteristics. We tried to avoid: (1) indiscriminate use of archaeological and architectural data in amalgamating a false earthquake or duplicating the past seismic events of the area (Ambraseys, 2005, 2009; Rucker and Niemi, 2010); (2) circumstantial evidence and speculations; and (3) circular reasoning (Rucker and Niemi, 2010; Sintubin et al., 2010). We should also emphasize that unlike in the developed industrialized countries, archaeoseismological study in Iran has not yet been developed. Therefore, there are not enough data for a quantitative paleoseismological approach and modeling of the damage pattern (Hinzen, 2005, 2009; Hinzen et al., 2011, 2013; Jusseret et al., 2013), or seismic response analysis and analyses of stability of structural elements (Argyriou et al., 2007; Psycharis et al., 2000, 2013; Hinzen, 2011; Ambraseys and Psycharis, 2011), especially for sites like Borujerd with poor data. Therefore, as a starting point, we took a qualitative approach toward damage patterns so far documented in only two introductory archaeological and architectural investigations mostly carried out during the 8 year Iran-Iraq war. We tried to substantiate the archaeological and architectural findings and major natural and anthropogenic events with the historic chronicles to avoid any error with respect to earthquake history of the region. Tables 1 through 3 are, therefore, designed to facilitate understanding, validating, and correlating the complicated chronology of the recorded events in the region in space and time. Appropriate original references to each entry are clearly cited in the tables. SEISMOTECTONIC SETTING Active tectonic processes have played an important role in the geomorphology of the Iranian Plateau since the Neogene collision of the Eurasian and Arabian plates and in continued convergence since then (Berberian, 1981, 1983, 2014; Berberian and King, 1981; Berberian et al., 1982; Jackson and McKenzie, 1988; Allen et al., 2004). The study area is located along the continental collision zone between Arabia (the Zāgros fold-and-thrust belt) in the SW and Eurasia (Central Iran) in the NE (Fig. 1), with 25 mm/yr total NNE-SSW shortening between Arabia and Eur-

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084761/spe525-05.pdf by University of Wisconsin Madison user

175

asia at longitude ~53°E (Sella et al., 2002; Nilforoushan et al., 2003; Vernant et al., 2004). This ongoing compression has given rise to the present physiography and active morphotectonics of the plateau, which is composed of high mountain belts and foothill fertile lands with reliable water supplies along mountain-bordering active reverse faults, or strike-slip faults, which were important for ancient agricultural colony settlements. This compression also created pathways through the mountain belts for migration and settlement of people through millennia. The physiography is likewise responsible for development of different Iranian social and cultural assortments. All these life-supporting outcomes of the active morphotectonics in the semiarid to arid Iranian Plateau have also created serious life-threatening seismic hazards to the inhabitants and numerous settlements because of high population concentrations along the active fault features since the Neolithic agricultural revolution (Berberian et al., 2012, 2014; Berberian, 2014). The seismic activity is distributed along numerous active fault lines throughout the plateau (Fig. 1), with individual faults rupturing every few centuries to thousands of years during strong earthquakes (Berberian, 1976, 1977, 1981, 2005, 2014; Berberian and Yeats, 1999, 2001, 2016). Such long recurrence intervals have resulted in an incomplete and inhomogeneous historic seismic record of the region (Berberian, 1994, 2014). Local global positioning system (GPS) observations (Tatar et al., 2002; Nilforoushan et al., 2003; Vernant et al., 2004; Walpersdorf et al., 2006) show that the northwestern Zāgros (including the study area; Fig. 1) accommodates the oblique convergence of Arabia and Eurasia by slip partitioning between (1) 3–6 mm/yr of orthogonal shortening (NE-SW) within the Zāgros fold-and-thrust belt, and (2) 4–6 mm/yr of orogenparallel (NW-SE) right-lateral strike-slip motion along regional faults, with 2.5 mm/yr along the active Zāgros Main Recent fault in the study area. The study area (Fig. 2) is a highly seismic-prone region with little recorded historic seismicity (Table 1). It is clear that the instrumental seismic data record is very short, especially for the Iranian Plateau, where pre–C.E. 1963 seismic data are either not recoded or are largely mislocated with large magnitude of location errors (Ambraseys, 1978; Berberian, 1979). Furthermore, the Iranian historical (pre–C.E. 1900) seismic records are inhomogeneous and incomplete, and very small proportions of medium- to large-magnitude earthquakes located in ancient urban areas have been recorded (Ambraseys and Melville, 1982; Berberian, 1994, 1997, 2005, 2014). This historic seismic data deficiency has resulted in unreliable seismic hazards assessments (for the case of the Bam earthquake and the citadel, see Berberian, 1976, 2005). However, archaeoseismological, paleo-architectural, and paleoseismological investigations can contribute to expand, and fortify, the historic seismic record, give a better seismotectonic picture, and create a complementary seismic data set for a more reliable seismic hazards and risk assessment. So far, no paleoseismic trench study has been carried out in the study area.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084761/spe525-05.pdf by University of Wisconsin Madison user

– 34.08, 48.40 – –



– –



07:24 09:31 16:17 19:36 01:17 01:31 11:54 08:31 12:08

– – –

– – –

2005.05.03 2005.06.18 2006.03.30 2006.03.30 2006.03.31 2006.03.31 2006.03.31 2006.04.01 2006.04.01

34.61, 47.50 34.60, 47.50

Night –

1909.01.23 1957.12.13 1958.08.14 1958.08.14 1958.08.16 1958.09.21 1961.10.14 1961.10.28 1963.03.24 1970.10.25 1971.01.20 1972.07.11 1975.03.03 1975.09.01 1987.05.29 1998.08.21 2002.04.24 2005.05.03





33.78, 48.88† 32.97, 48.43† 33.58, 48.73† 33.55, 48.73† 33.58, 48.79† 33.75, 48.65† 33.63, 48.67† 33.88, 48.75† 33.83, 48.81†

33.38, 49.28 34.34, 47.62* 34.27, 48.10 34.33, 48.02 34.24, 47.84* 34.63, 47.46 34.00, 48.51 33.53, 48.49* 34.43, 47.89* 36.75, 45.18* 35.02, 46.86* 36.10, 45.64* 35.70, 45.70 33.34, 49.12* 34.05, 48.28* 34.12, 48.15† 34.64, 47.40† 33.48, 48.52†

32.57, 50.20 33.00, 49.60

34.61, 47.50



33.30, 49.26

Coordinates (°N, °E)





Origin time (h:min; UTC) – –

– 03:00 L.T. 02:48 01:45 11:27 15:26 19:13 16:18 07:00 10:46 12:44 11:22 21:32 22:49 16:05 23:15 06:27 05:13 19:48 07:21

1853.06.11 1876.09.28

912.08.18– 913.08.07 956.04.15– 957.04.05 1008.04.27 1107.08.22– 1108.08.11 1135.07.25 1135.08.13 Late 1191– Early 1192 1310.05.31– 1311.05.19 1316.01.05 1429.09.30– 1430.09.15 1494.10.02– 1495.09.21

Prehistoric 1650–1600 B.C.E. 276–578 C.E.

Date (year.month. day)

5.1

5.2 6.1

4.6 4.5 5.2 4.2

5.3 4.9 5.4 4.9 3.8 4.0 4.6 5.9 5.2 4.4 3.9

5.8 5.2 5.0

7.4 6.7 5.7 5.5 6.6 5.2

5.5 5.8

>6.0

>7.0 >6.0

>5.5?

>6.5 >6.5 ?

7.0 6.5

>6.0

Ms

5.8 5.5

6.5

7.4 6.8

Mw

VII

5.3 4.8

4.1 3.8

5.7 4.7

4.1 4.8 4.8 VIII–

VIII+ VI+ VII VII VII+

Chālānchulān Chālānchulān Chālānchulān Chālānchulān

Armanijān Choghālvandi, S. Borujerd

Taqiābād Kāhriz

VI+ VII–

5.0 5.5 5.3 5.0 4.3 4.2 4.9 5.0

6.2

IX+ VIII

VII

Jebāl/Media Province, Hamédān SE Faridan (Periā) NW Faridan (Periā) Silākhor (Dorud) Fārsinaj Givaki Kalādeh Firuzābād Kargsār Mogh Heydarābād Kārkhāneh Pāveh Dānān Bāneh

>VII+

Shahr-e Zur

W. Kordestān W. Kordestān Hamédān

Dinévar Kargsār

Hamédān

Dinévar

Kangāvar

Lake Gahar Gowdin–Giyān

Earthquake name

Silākhor (Borujerd) Hamédān

>VII?

>VIII >VIII F

IX VIII

>VII+

I

>VIII >VII+

7.2 6.5

Mb

319/67/-168

321/70/-167 318/63/174

218/80/2 025/39/-84 036/74/16 220/71/-17

314/52/-164 319/50/-154

325/70/10

006/52/-125

Nodal plane 1 st./dip/ra.

224/79/-24

226/77/-20

128/88/170 197/51/-095 302/75/163 316/74/-161

214/77/-39 212/70/-43

058/81/20

137/50/54

Nodal plane 2 st./dip/ra.

RLSS (b)

RLSS (b) RLSS (c)

RLSS (b) N (b) RLSS (b) RLSS (b)

RLSS RLSS

RLSS (a)

RLSS RE (a)

FT

308 287

304 302

328

Slip vector

06

H

H P+

H H H H

N&B J&M

S

B M

T

B B&Y, Table 1 B&Y, Table 1

Ref.

(Continued)

CDw (km)

TABLE 1. RECORDED SEISMICITY OF THE ZĀGROS MAIN RECENT FAULT (ZMRF) WITH AVAILABLE EARTHQUAKE SOURCE PARAMETERS (FOR REFERENCES, SEE THE TEXT; BERBERIAN, 2014; FIGS. 6 AND 22)

176 Berberian et al.

TABLE 1. RECORDED SEISMICITY OF THE ZĀGROS MAIN RECENT FAULT (ZMRF) WITH AVAILABLE EARTHQUAKE SOURCE PARAMETERS (FOR REFERENCES, SEE THE TEXT; BERBERIAN, 2014; FIGS. 6 AND 22) (Continued) Ms Mb I Earthquake name Nodal plane 1 Nodal plane 2 FT Slip CDw Date Origin Coordinates Mw Ref. st./dip/ra. st./dip/ra. vector (year.month. time (°N, °E) (km) day) (h:min; UTC) 2006.06.06 17:03 35.62, 46.02 4.9 4.1 313/41/-168 214/82/-49 RLSS (b) H 4.9 4.6 4.5 239/78/11 147/79/168 RLSS (b) H 2006.09.26 08:14 31.91, 50.67† 2007.03.06 23:32 33.40, 48.86† 4.1 Dorud 2010.11.06 03:52 33.37, 48.91† 5.0 4.9 Dorud 4.9§ Darband (W. Aznā) 2012.11.27 06:22 33.44, 49.30§ Note: UTC—Coordinated Universal Time; Mw—moment magnitude; Ms—surface-wave magnitude; Mb—body-wave magnitude; st./dip/ra.—strike/dip/rake deduced from (a) first motion fault plane solution; (b) Harvard centroid moment tensor (CMT) solution; and (c) constrained focal mechanism by P and SH body wave modeling. FT—fault type (mechanism): (a) first motion fault plane solution; (b) Harvard CMT solutions; and (c) constrained focal mechanism by P and SH body wave modeling; RLSS—rightlateral strike-slip; RE—reverse; N—normal; CDW—centroid depth deduced from P and SH body wave modeling; I (MMI)—epicentral intensity (modified Mercalli intensity scale); F—earthquake strongly felt; L.T.—local time. References: B—Berberian (1995, 2014); B&Y—Berberian and Yeats (2001); H—Harvard (2016); J&M—Jackson and McKenzie (1984); M— McKenzie (1972); N&B—Ni and Barazangi (1986); P+—Peyret et al. (2008); S—Shirokova (1967); T—this study. *Engdahl et al. (2006). † ISC (2016). § The Iran Telemetered Seismograph Network (ITNS) operated by the Iranian Seismological Research Center (IRSC; http://irsc.ut.ac.ir) at the Institute of Geophysics of Tehran University (IGTU); with local magnitude.

Evidence of historical seismic activity along the Zāgros Main Recent fault at Borujerd

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084761/spe525-05.pdf by University of Wisconsin Madison user

177

ARCHAEOLOGICAL, ARCHITECTURAL, AND ARCHAEOSEISMOLOGICAL CONTEXT The enchanting Iranian Plateau, the home of one of the ancient civilizations of the world, with architectural history of more than 6000 yr, has an extraordinary wealth of old monuments, culture, literature, and art. There are ~45,000 archaeological sites on the Iranian Plateau (CHTHO, 2011). Some of these sites, especially those located near active faults, show numerous evidence of seismic damage and destruction, but so far the earthquake indicators have been overlooked (Berberian, 1994, 2014). Furthermore, little comprehensive analysis exploring the effects of archaeoseismological events on the archaeological sites, monuments, and ancient societies located along the seismic faults has been published (Berberian and Yeats, 2001; Berberian et al., 2012, 2014). In this section, we review possible archaeoseismological indicators of off-fault earthquake strong ground motion effects on structural elements and archaeological stratigraphy of the Borujerd congregational mosque in the vicinity of the Zāgros Main Recent active fault in the Silākhor Plain (Fig. 2). Little is known about the ancient town of Borujerd (33°53′50″N, 48°54′6″E, +1567 m), located in the Lorestān province of western Iran (Figs. 1 and 2). There are references to its Sāssānid (224–642 C.E.) Zoroastrian fire temple in the chronicles (Estakhri, 951, p. 165; Hazin Borujerdi, 1972, p. 82; Maulānā Borujerdi, 1974, p. 9); however, there is no extant pre–fourteenth-century monument left at the present city or the nearby towns and villages. Archaeological investigation revealed some typical Sāssānid-era bricks reused in the construction of the Borujerd congregational mosque after destruction of the fire temple during invasion of the Moslem Arabs in 643–652 C.E. (Table 2). Throughout the eighteenth century, Borujerd was turned into a large garrison during the Qājār Dynasty (1779–1925) in their drive to appease the nomadic tribes of the Lorestān province (Ja’fari Moqaddas, 1975; Rezvāni, 1997). The population of the town at the end of the nineteenth century is reported to have been ~22,000 inhabitants (Ehlers, 1989). The population of Borujerd in 2011 was 240,654 (SCI, 2012). The present Borujerd congregational mosque (masjed jāmé Borujerd; National Heritage ID 228 dated 7 December 1935), with a dome chamber from Shāhrokh Timurid period (ca. 1405– 1447), Safavid Ayvān2 (1614), two minarets (1794), and outer dome shell (1970s), is one of the ancient monuments of the city (Fig. 3; Table 2). The nearby mausoleum of Ja’far (National Heritage ID 1855) with the “sugar-loaf” dome and construction date of 1125 C.E. has been restored several times since its construction date and thus has lost its original architecture and historic value. The third monument in the city is the Qājār period (1794) Soltāni (Shāh) mosque (National Heritage ID 393) has remained

2 Ayvān (ivan/iwan [sic]): porch; a barrel-vaulted hall in which the fourth side is open to a courtyard. Minaret: from the ancient Persian word “menār,” meaning a “watch- or signaling-tower” lit with fire.

178

Berberian et al. TABLE 2. CONSTRUCTION PHASES OF THE BORUJERD CONGREGATIONAL MOSQUES SINCE THE EIGHTH CENTURY C.E. (FIGS. 3, 4, 5, 7, AND 13)

Borujerd congregational mosques Mosque IV

Date/ period C.E. [H] 2006

Mosque IV

Qajar: 1794 [1209]

Mosque IV

Safavid: 1681 [1092]

Mosque IV

Mosque IV

Structure Repairs after the earthquake

Repair of Borujerd congregational mosque & building, the two minarets (inscription) & ayvān Repairs of congregational mosque

Safavid: 1658– 1659 [1069]

Repairs of Borujerd congregational mosque; installing the wooden ninestep menbar on the –20 cm floor below the present floor (inscription) Timurid: ca. post– The present smaller & lower 1405/pre–1447 dome chamber built [800–850]

Mosque III [Rāzi style]

Late Buyid–Early W/single isolated minaret Saljuq: ca. post– constructed on the remains of 1090/pre–1139 the earlier mosque [483–533]

Mosque II [Rāzi style]

Buyids: ca. 945– 971 [333–360]

Mosque I [Khorāssāni style]

‘Abbāsid: 9th century [3rd century]

Traces & debris w/potsherds at –180 to –240 cm below the present floor

3 Mehrāb (mihrab [sic]): niche; an ornamental doorway-like indentation in the southern wall of a mosque. Originally adapted from the Mithra (Mehr) cult; Mehrābé.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084761/spe525-05.pdf by University of Wisconsin Madison user

Source

Mosque and minarets IIEES (2006); this study damaged by the 31 March 2006 Mw 6.1 Chālānchulān earthquake Hazin Borujerdi (1972); Mehryār (1985); Moqaddas ? (1997) ?

Hazin Borujerdi (1972); Mehryār (1985); Moqaddas (1997) Mehryār (1985); Moqaddas (1997); this study

?

Pre-1658 [Pre-1069] Collapse of the mudbrick ribbed vaults & arches in the dome chamber basement of Borujerd congregational mosque IV Destruction of the pre14th-century Borujerd congregational mosque III dome chamber & its minaret by the 1316 earthquake Tilted brick pillars to the SE & E (unknown time & cause) ? (Destroyed before 971)

Early Mosque [Khorāssāni style] UmayyadAt depth of –240 to –420 cm ‘Abbāsid: below the present floor of the Ca. 8th century congregational mosque, above [ca. 2nd century] the native soil Zoroastrian fire temple Sassanid: Chahār Tāq 224–642 Note: H—Arabic lunar (Hijra) calendar used in pre–1900 C.E. chronicles.

standing since its construction and was repaired in 1937 and after the 2006 earthquake (Meshkāti, 1970; Mehryār, 1985; Moqaddas, 1997). Due to: (1) numerous renovations, alterations, and modifications to the architectural elements of the Borujerd historic congregational mosque; (2) several phases of damage and destruction; (3) gradual addition of various wings; and (4) numerous phases of reuse of older bricks with different dimensions in the newer constructions, the structure has become an amalgamated hodgepodge complex. At each stage, the floors of the complex as well as the elevations of the mehrābs3 and the corridors have been raised. The size, shape, and decorations of the mehrābs have also been modified at least seven times. Furthermore, the shape

Possible earthquake effect



Mehryār (1985); Moqaddas (1997); this study

Meshkāti (1970); Mehryār (1985); Moqaddas (1997); this study

Estakhri (951); Hazin Borujerdi (1972); Mehryār (1985); Moqaddas (1997); this study Estakhri (951); Ibn Hawqal, (978); Yāqut (1225); Hazin Borujerdi (1972); Mehryār (1985); Moqaddas (1997) Hazin Borujerdi (1972); Mehryār (1985); Moqaddas (1997)

– (Destroyed by the Moslem Arabs)

of the structure and the height of the walls and dome chamber have been altered numerous times. Moqaddas and Mehryār, who were involved in the excavations below the present floor of the Borujerd congregational mosque as well as restoration of the dome chamber and the walls, categorically stressed “ca. pre–early eighth-century H [ca. pre– early fourteenth-century C.E.] severe destruction” of the older structure, which was followed by a “major reconstruction of the dome chamber” later during the prosperous reign of Shāhrokh Timurid (r. 1405–1447 C.E.). All the datings are based on artifacts, numismatic material, and style of the structures, and no radiometric ages are so far available to constrain the discussed phases. The approximate given timing of one of the major destruction phases corresponds with the 1316 C.E. earthquake recorded in a contemporary historic text. The archaeologists working on the excavations did not have any idea about the 1316 C.E. earthquake and did not mention it in their reports; however,

Evidence of historical seismic activity along the Zāgros Main Recent fault at Borujerd

179

Figure 3. A south (left)–north (right) profile of the Borujerd congregational mosque (revised after Moqaddas, 1997). Dates on the structures as well as the construction periods of mosques II through IV and recent (left margin) are added from this study. See Figure 4 for the plan of the structure.

they considered destruction of the structure by “a major earthquake” (we return to this issue later in the text). The structure of the Borujerd congregational mosque contains abundant disrupted architectural relicts and collapses, at least since the tenth century C.E. (Figs. 3 and 4; Table 3). In order to configure the timing of the complex collapse/ reconstruction phases with probable natural and anthropogenic events, we focused on the major discovered phases, and analyzed the set of architectural relict disturbances and deformations for their characterization and classification as being true coseismic features or not. To avoid any misperception, we attribute each major structural phase with Roman numerals. At each phase, the floor of the congregational mosque, its corridors, and mehrābs were elevated by ~30–40 cm. At least four floor levels were discovered in a sounding between the tenth-century C.E. Buyid (elevation –180 cm below the present floor) and the seventeenth-century Safavid-era (–20 cm) reconstruction phases (Figs. 3, 4, and 5).

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084761/spe525-05.pdf by University of Wisconsin Madison user

Early Borujerd Mosque (Ca. Eighth Century C.E.) Apparently, the original early mosque (ca. second century H/eighth century C.E.) was built somewhere at depths of –240 cm to –420 cm (virgin soil; Fig. 5) below the present floor of the congregation mosque, where debris with potsherds was detected in a small sounding (Mehryār, 1985; Moqaddas, 1997). The 180 cm interval in the sounding contains a black debris layer with potsherds at the top (–190 to –230 below the present floor) underlain by sand and earth mixture with a lime mortar layer at –270 to –305 of unknown age. A sandy layer with potsherds (date unknown) was detected resting on top of the native soil at depth of –405 to –420 cm below the present floor (Fig. 5). Unfortunately, the archaeological layers were never dated or linked properly to the major destruction and rebuilding phases during the excavation period or discussed in the published reports. The “early mosque” of Borujerd was referred to as “the old mosque” sometime prior to 841 C.E. (Estakhri, 951, p. 165; Ibn

180

Berberian et al.

Figure 4. Plan of the Borujerd congregational mosque (revised after Mehryār, 1985). Dates of different structures added from this study.

Hawqal, 978, p. 258, 262; Yāqut, 1225, I:596, II:737; Hazin Borujerdi, 1972, p. 82; Table 2). The “early mosque” should have been a very simple single quadrangular poorly constructed structure using local native building material. Because of Islamic hostility to statues and ornaments, the simple structure was void of ornaments, and vanity. Furthermore, during those early days, the country was still in anarchy and turmoil resisting the occupation, and the Moslem Arab invaders did not have the opportunity and architectural knowledge to construct sophisticated structures, which were later developed by the Iranian master architects. Hazin Borujerdi (1972, p. 82), without presenting any historical evidence or citing a reference, stated that this “early mosque” was destroyed by the followers of the Iranian revolutionary freedom fighter and nationalist, Bābak Khorramdin (ca. 797–838 C.E.), sometime in the early ninth century C.E. However, his

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084761/spe525-05.pdf by University of Wisconsin Madison user

statement cannot be affirmed without further systematic investigation and excavation. Borujerd Congregational Mosque I (Ca. Ninth Century C.E.) The Borujerd congregational mosque I was built on the site or adjacent to the Sāssānid (ca. 224–642 C.E.) Zoroastrian fire temple and the early mosque (eighth century C.E.), possibly sometime in the early ninth century C.E. (Table 2). Remnants of this phase were apparently found buried below the –180 cm floor (Fig. 5; Maulānā Borujerdi, 1974; Mehryār, 1985; Moqaddas, 1997; Rezvāni, 1997). Construction of the congregational mosque I is attributed to Hamuleh ben ‘Ali al-Borujerdi (d. 856 C.E.), the ruler of Borujerd, who was appointed by the governor of the Jebāl (ancient Media/

Evidence of historical seismic activity along the Zāgros Main Recent fault at Borujerd

181

TABLE 3. PERIODS, IMPORTANT EVENTS, AND MAJOR RECORDED HISTORIC EARTHQUAKES RELATED TO THE CONFIGURATION OF THE MAJOR HISTORIC MONUMENTS IN THE CITIES OF BORUJERD (33°53′N, 48°54′E, +1569 m), HAMÉDĀN (34°47′N, 48°30′E, +1819 m), GOLPĀYÉGĀN (33°27′N, 50°17′E, +1817 m), AND AZNĀ (33°27′N, 49°27′E, +1878 m) ORGANIZED IN CHRONOLOGICAL ORDER (SEE FIGS. 2 AND 16 FOR LOCATIONS) Date C.E. [Hijra/Persian]

Historic period

Archaeological milestones of structures (and architectural style)

2010.11.06 [1389.08.15] 2006.03.31 [1385.01.11]

Islamic regime

2006.03.31 [1385.01.11]

Islamic regime

2006.03.30 [1385.01.10]

Islamic regime

2005.05.03 [1384.02.13]

Islamic regime

1987.01.12 [1365.10.22]

Islamic regime Borujerd congregational mosque damaged by Iraqi bombardment Pahlavi a

1961.10.28 [1340.08.06]

Earthquake Dorud: Mb 4.9 Chālānchulān Aftershock: Mw 5.1; I VII Chālānchulān: Mw 6.1; I VIII–

Islamic regime

Heydarābād: Ms 5.0 5.0,, I VII

Pahlavi a

1958.08.14

Pahlavi

Firuzābād: Ms 5.7 5.7,, I VII+

1958.08.14

Pahlavi

Firuzābād: Ms 5.5 5.5,, I VII

1937 [1316] 1909.01.23 [1287.11.03]

Pahlavi

1876.09.28 [1293.09.09]

‘Aliābād Mogh: Ms ?, I VII Firuzābād: Mw 6.5, I VIII+

Renovation of Borujerd congregational mosque

Qājār

1859 Qājār [1275] 1830–1831 [1246] Qājār 1853.07.11 [1269.10.04] 1798 [1213]

Qājār

1794 [1209]

Qājār

Qājār

– Silākhor: Mw 7.4; I IX+

Qājār

NW Faridan (Periā§): Ms ~5.8 ~5.8,, I ~VII+ Borujerd houses were in ruins – in 1859 Addition of the Howzkhāneh – to the Borujerd congregational mosque SE Faridan (Periā§): Ms ~>5.5 ~>5.5,, I ~>VII+ Northern Shabestān added – to Borujerd congregational mosque Repair of Borujerd congregational mosque & building the two minarets (inscription) & ayvān

F at Bor ujerd (V) Felt Borujerd +

V at Borujerd

Source This study This study

VII– at Borujerd. Two minarets IIEES (2006); this of Borujerd congregational study mosque severely fractured, one top minaret collapsed & ayvān damaged;; ayvān arch of Soltāni mosque damaged; partial collapse of dome of Emāmzādeh J’afar of Borujerd This study. V+ at Borujerd

Chālānchulān Foreshock: Mw 5.1; I VII South Borujerd: Borujerd r strongly shaken (IV); irna.ir (2005); IIEES Mw 4.9; I VI+ 10 suburb villages damaged, 1 (2005); this study person killed, 26 injured & 200 villagers demanded tents – – Moqaddas (1997)

1961.10.14 [1340.07.22] 1958.08.16

Pahlavi

Earthquake effect

?

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084761/spe525-05.pdf by University of Wisconsin Madison user

V+ at Borujerd: one turnet Berberian (1995) (goldasteh) of congregational mosque collapsed, dome cracked V+ at Borujerd; some houses Berberian (1995) cracked VI at Borujerd Ambraseys & Moinfar (1974); Tchalenko & Braud (1974); Berberian (1995) V+ at Borujerd Ambraseys & Moinfar (1974); Tchalenko & Braud (1974); Berberian (1995) V+ at Borujerd Ambraseys & Moinfar (1974); Tchalenko & Braud (1974); Berberian (1995) – Mehryār (1985); Moqaddas (1997) VII at Borujerd; congregational Ambraseys & Moinfar mosque fractured & repaired (1973); Ambraseys later & Melville (1982); Berberian (1995, 2014); this study 120 km SE of Borujerd Ambraseys (1979) –

Sani’ al-Dauleh (1881) Moqaddas (1997)

– 200 km SE of Borujerd –

?

Ambraseys (1979) Hazin Borujerdi (1972); Mehryār (1985); Moqaddas (1997) Hazin Borujerdi (1972); Mehryār (1985); Moqaddas (1997) (Continued)

182

Berberian et al.

TABLE 3. PERIODS, IMPORTANT EVENTS, AND MAJOR RECORDED HISTORIC EARTHQUAKES RELATED TO THE CONFIGURATION OF THE MAJOR HISTORIC MONUMENTS IN THE CITIES OF BORUJERD (33°53′N, 48°54′E, +1569 m), HAMÉDĀN (34°47′N, 48°30′E, +1819 m), GOLPĀYÉGĀN (33°27′N, 50°17′E, +1817 m), AND AZNĀ (33°27′N, 49°27′E, +1878 m) ORGANIZED IN CHRONOLOGICAL ORDER (SEE FIGS. 2 AND 16 FOR LOCATIONS) (Continued) Date C.E. Historic Archaeological milestones of Earthquake Earthquake effect Source [Hijra/Persian] period structures (and architectural style) 1794 Qājār Borujerd Soltāni (Shāh) – – (Cause of destruction of the Hazin Borujerdi [1209] mosque built atop an older older mosque is unknown) (1972); Meshkāti Sunni mosque, w/numerous (1970); Mehryār repairs & renovations: 1831, (1985); Moqaddas 1874, 1876, 1879, 1937 (1997) 1681 Safavid Repairs of congregational ? ? Hazin Borujerdi [1092] mosque (1972); Mehryār (1985); Moqaddas (1997) 1665.07.14– Ms ~>6.5, I ~>VIII NW Ardal (250 km SE of Safavid a Al-‘Umari (1793, fol. 1666.07.03 Borujerd); strongly felt in 216v); Ambraseys [1076] Hamédān (330 k km to the NW (1979); Ambraseys of Ardal), Shirāz & Esfahān & Melville (1982); Exact seismic parameters Berberian (1994, unknown 1995) 1658–1659 Safavid Repairs of Borujerd – – Hazin Borujerdi [1069] congregational mosque; (1972) installing the wooden 9-step menbar on the –20 cm floor below the present floor (inscription). Pre–1658 Collapse of the mudbrick ? (Exact date of collapse has not Mehryār (1985); [Pre–1069] ribbed vaults & arches in the yet been documented) Moqaddas (1997) dome chamber basement of Borujerd congregational mosque IV (1405–1447). 1619 Safavid Repairs of Borujerd ? ? Mehryār (1985); [1028] congregational mosque Moqaddas (1997) 1613 [1022] 1495 [900]

?

?

Timurid

Ms ~>6.0 ~>6.0,, I ~>VII+ (?)

Strongly felt in Hamédān (where caused landslides), Esfahān & district of Ray Exact epicentral location & seismic parameters unknown

1430 [833]

Timurid

Ms ~>6.0, I ~>VII+ (?)

Destruction in Hamédān Exact epicentral location & seismic parameters unknown

ca. post–1405/ pre–1447 [800–850]

Timurid

1370–1375 1386–1388 1399–1405 1340 [740]

Safavid

Hazin Borujerdi (1972) Al-‘Umari (1793: fol. 160r); Sani’ al-Dauleh (1880–1882, II:57); Ambraseys & Melville (1982); Berberian (1994) Al-‘Umari (1793: fol. 160r); Sani’ al-Dauleh (1880–1882, II:57); Ambraseys & Melville (1982); Berberian (1994) Mehryār (1985); Moqaddas (1997)

BORUJERD – – CONGREGATIONAL MOSQUE IV: The present smaller & lower dome chamber built Invasions of Timur Turco-Mongol Hordes. Borujerd was sacked and people were massacred in Rabi’-II 788/April 1386 C.E. (Hazin Borujerdi, 1972). lkhānids

1337 [737] (wood lkhānid or door) or Timurid 1405–1446 [807– 850] (2 sanduqs) 1316.01.05 Ilkhānid [715] 1310 [710]

Repairs of Borujerd congregational mosque

Ilkhānid

Described Borujerd & referred – to the old & new mosques (possibly repeated older statements) Aznā Qāsem Jāpalaq – mausoleum w/numerous repairs & renovations: 1405, 1446 & more Silākhor: Ms ~>6.5 ~>6.5,, I ~>VIII+ Shahr-e Zur, Kordestān K Kordest ān

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084761/spe525-05.pdf by University of Wisconsin Madison user



Mostaufi Qazvini (1340)



Meshkāti (1970); Golombek & Wilber (1988)

Destruction of the pre-14thcentury Borujerd congregational mosque III dome chamber & its minaret Shahr-e Zur destroyed, many killed

Kāshāni (1320, p. 179); Sani’ al-Dauleh (1880–1882, II:23); this study Al-‘Umari (1793, fol. 121r); Ambraseys & Melville (1982); Berberian (1994) (Continued )

Evidence of historical seismic activity along the Zāgros Main Recent fault at Borujerd

183

TABLE 3. PERIODS, IMPORTANT EVENTS, AND MAJOR RECORDED HISTORIC EARTHQUAKES RELATED TO THE CONFIGURATION OF THE MAJOR HISTORIC MONUMENTS IN THE CITIES OF BORUJERD (33°53′N, 48°54′E, +1569 m), HAMÉDĀN (34°47′N, 48°30′E, +1819 m), GOLPĀYÉGĀN (33°27′N, 50°17′E, +1817 m), AND AZNĀ (33°27′N, 49°27′E, +1878 m) ORGANIZED IN CHRONOLOGICAL ORDER (SEE FIGS. 2 AND 16 FOR LOCATIONS) (Continued ) Date C.E. Historic Archaeological milestones of Earthquake Earthquake effect Source [Hijra/Persian] period structures (and architectural style) 1309–1316† Ilkhānid Hamédān ‘Alaviān Tomb – – Herzfeld (1922); Pope [709–716†] Shrine (1934); Wilber (1955) 1275 Ilkhānid Described Borujerd – – Zakaryā Qazvini [674] (1275) 13th century Ilkhānid Hamédān Esther and – – Meshkāti (1970) [7th century] Mordékhāi Tomb Shrine 1262 Ilkhānid ‘Emad al-Din Hasan bn ‘Ali – – Hazin Borujerdi [660] Tabari delivered a sermon (1972) at Borujerd congregational mosque III 1225 Ilkhānid Described Borujerd & referred – – Yāqut (1225) [623] to the congregational mosque III 1219–1259 Invasion of the Mongol Hordes. Borujerd was sacked & people were massacred in Rajab 617/September 1220 C.E. (Hazin [616–657] Borujerdi, 1972). 1194.03.00 Ms ?, I ? Widely felt in Mesopotamia Saljuq Ibn al-Athir (1231, [590] (w/some damage in Najaf) & XII:72); Sani’ al-Dauleh Jebāl* (1880–1882, I:212); Exact epicentral location & Ambraseys & Melville seismic parameters unknown (1982); Berberian (1994) 1191–1192 Ms ?, I ? Strongly felt in Hamèdān Abu Hāmed (1203, Saljuq [587–588] & disrupted a 4 d battle in p. 89); Ambraseys city; epicenter location & & Melville (1982); magnitude unknown in Jebāl* Berberian (1994) Exact epicentral location & seismic parameters unknown † Saljuq (?) Hamédān ‘Alaviān Tomb – – Mostafavi (1953, 1147 (?) † [542 (?) ] Shrine 1967); Meshkāti (1970); Hātami (2000) 1139–1145 Saljuq The Kufic inscription on ? ? Moqaddas (1997) [533–539] Borujerd congregational mosque III indicating a major repair Ms ~>6.5, I ~>VIII+ Kordestān; widely felt in Ibn al-Jauzi (1181, 1135.08.13 [529] Saljuq Baghdād X:46); al-Qusi (1907, Exact epicentral location & p. 95); Ambraseys seismic parameters unknown & Melville (1982); Berberian (1994) Ms ~>6.5, I ~>VIII+ Destructive in Kordestān, 1135.07.25 [529] Saljuq Ibn al-Jauzi (1181, widely & strongly felt in X:46); al-Suyuti (1505, Mesopotamia, Jebāl* & p. 37); al-Qusi (1907, Mosul; followed by violent p. 95) aftershocks Exact epicentral location unknown Ms ~>6.5, I ~>VIII+ Destructive in the western Ibn al-Athir (1231, 1130.02.27 [524] Saljuq Zāgros/Kordestān Zāgros/Kor K destān & Jebāl*, X:469); Ibn al-Jauzi w/widespread damage from (1181, X:14); Bar Mesopotamia to Jebāl* Hebraeus (1286, p. Exact epicentral location & 255/289); Ambraseys seismic parameters unknown & Melville (1982); Berberian (1994) 1125 Saljuq Borujerd Ja’far mausoleum w/ Meshkāti (1970) [519] numerous repairs & rebuilding. 1105–1118 Saljuq Golpāyégān congregational Godard (1965); [498–512] mosque w/numerous repairs & Meshkāti (1970); rebuilding O’Kane (1994); Pope, (1965, 1997) Damaging in Hamèdān & 2 Ms ~>6.0, I >VII+ 1087.11.00 [480] Saljuq Ibn al-Jauzi (1181, districts on the outskirts in IX:38); Ambraseys Jebāl*; many killed, followed & Melville (1982); by strong aftershocks. Exact Berberian (1994) epicentral location & seismic parameters unknown. (Continued )

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084761/spe525-05.pdf by University of Wisconsin Madison user

184

Berberian et al.

TABLE 3. PERIODS, IMPORTANT EVENTS, AND MAJOR RECORDED HISTORIC EARTHQUAKES RELATED TO THE CONFIGURATION OF THE MAJOR HISTORIC MONUMENTS IN THE CITIES OF BORUJERD (33°53′N, 48°54′E, +1569 m), HAMÉDĀN (34°47′N, 48°30′E, +1819 m), GOLPĀYÉGĀN (33°27′N, 50°17′E, +1817 m), AND AZNĀ (33°27′N, 49°27′E, +1878 m) ORGANIZED IN CHRONOLOGICAL ORDER (SEE FIGS. 2 AND 16 FOR LOCATIONS) (Continued ) Date C.E. Historic Archaeological milestones of Earthquake Earthquake effect Source [Hijra/Persian] period structures (and architectural style) Ms ~>6.5, I ~>VIII+ Damaging in W. Kordestān, Ibn al-Jauzi (1181, 1058.12.08 [450] Saljuq many killed. Strongly felt VIII:190); Ibn al-Athir in Mosul, Wāsit/Wāseth, (1231, IX:449); alBaghdād & Hamédān. Exact Suyuti (1505, p. 33); epicentral location & seismic Ambraseys & Melville parameters unknown. (1982); Berberian (1994) BORUJERD Ca. post–1090 Late Buyid– – – Meshkāti (1970); CONGREGATIONAL pre–1139 Early Saljuq Mehryār (1985); [483–533] Moqaddas (1997) MOSQUE III [RĀZI STYLE]: W/single isolated minaret constructed on the remains of the earlier mosque ca. 1000 Invasion of the Saljuq Turk Hordes. [ca. 390] 982 [372] Buyids Described Borujerd Anonymous (982) 978 Buyids Described Borujerd & referred Ibn Hawqal (978) [367] to the congregational mosque I Damaging in Hamèdān 956–957 956: Ibn al-Athir (1231, Buyids [102 km NNW of Borujerd], [345] Ms ~>6.0, I >VII+ VIII:388); al-Suyuti ‘Abbāsābād [114 km NW of (1505, p. 29); Sani’ alBorujerd] & their districts in Dauleh (1880–1882, Jebāl*, killing many people; I:131); Ambraseys felt in Qasr-e Shirin [207 km to & Melville (1982); the WSW of Hamédān] Berberian (1994) Exact epicentral location & seismic parameters unknown BORUJERD Ca. 945–971 Buyids – Tilted brick pillars to the SE & E Estakhri (951); Hazin CONGREGATIONAL [333–360] (unknown time & cause) Borujerdi (1972, p. 82); Mehryār (1985); MOSQUE II [RĀZI STYLE]: Moqaddas (1997) traces & debris w/potsherds at –180 to –240 cm below the present floor 951 Buyids Described Borujerd & referred – – Estakhri (951) [350] to the congregational mosque I BORUJERD – ? (Destroyed before 971) Estakhri (951); Ibn 9th century ‘Abbāsid CONGREGATIONAL Hawqal (978, p. 258, [3rd century] 262); Yāqut (1225, MOSQUE I [KHORĀSSĀNI I:596, II:737); Hazin STYLE]: see text Borujerdi (1972, p. 82); Mehryār (1985); Moqaddas (1997) EARLY BORUJERD Ca. 8th century Umayyad– – – Hazin Borujerdi (1972, p. 82); [ca. 2nd century] ‘Abbāsid MOSQUE [KHORĀSSĀNI Mehryār (1985); STYLE]: Moqaddas (1997) at depth of –240 to –420 cm below the present floor of the congregational mosque, above the native soil 636–652 Invasion of the Moslem Arabs: Destruction of infrastructure & denunciation of the Zoroastrianism. [15–32] Ms ~>6.8, I ~>VIII Kāmbakhsh-Fard Sāssānid Kangāvar Kangāvar earthquake: 274–578# [34°30′N, 47°57′E, +1488 m] Destruction, fire fire & reconstruction (1994); Berberian of the Anāhitā Temple T (1994, 2014); Āzarnoush (2009); Berberian & Yeats (2001) 224–642 C.E. Sāssānid Piruzgerd [later Borujerd] fire ? ? Hazin Borujerdi temple. (1972, p. 82); Maulānā Borujerdi (1974); Mehryār (1985); Moqaddas (1997); Rezvāni (1997) 312 B.C.E.–C.E. Seleucid– ? ? ? – 224 Parthians 330 B.C. Invasion of Alexander III of Macedonia (Continued )

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084761/spe525-05.pdf by University of Wisconsin Madison user

Evidence of historical seismic activity along the Zāgros Main Recent fault at Borujerd

185

TABLE 3. PERIODS, IMPORTANT EVENTS, AND MAJOR RECORDED HISTORIC EARTHQUAKES RELATED TO THE CONFIGURATION OF THE MAJOR HISTORIC MONUMENTS IN THE CITIES OF BORUJERD (33°53′N, 48°54′E, +1569 m), HAMÉDĀN (34°47′N, 48°30′E, +1819 m), GOLPĀYÉGĀN (33°27′N, 50°17′E, +1817 m), AND AZNĀ (33°27′N, 49°27′E, +1878 m) ORGANIZED IN CHRONOLOGICAL ORDER (SEE FIGS. 2 AND 16 FOR LOCATIONS) (Continued ) Date C.E. Historic Archaeological milestones of Earthquake Earthquake effect Source [Hijra/Persian] period structures (and architectural style) Ms ~>7.0, I ~>VIII Godin earthquake: Young (1968, 1969); 1650–1600 B.C.E. Prehistoric Godin III:2 Complete destruction & Young & Levine Godin (34°31′06.29″N, abandonment of Godin & Giyān (1974); Berberian 48°04′06.52″E, +1485 m), sites (1994, 2014); Giyān (34°10′53.34″N, Berberian & Yeats 48°14′37.83″E, +1563 m), (2001); Berberian et NW Zāgros al. (2014) Note: Recorded earthquakes are shown with bold letters and shaded in gray, and the main rebuilding phases of the Borujerd congregational mosque are in uppercase. Dates in brackets are Hijra Arabic lunar calendar before 1900 C.E., and Persian solar calendar after 1900 C.E. I—epicentral intensity (modified Mercalli intensity scale); Ms—surface-wave magnitude; Mw—moment magnitude; Mb—body-wave magnitude. *Jebāl, misspelled in literature as “Jibāl” (Arabic term for “mountains”) from the 7th to the 12th century A.D.; later called “Erāq ‘Ajami” (misspelled as “Irāk ‘Ajami”; to be distinguished from the Arabic Irāq, i.e., Mesopotamia). It covered the broad mountainous region of Iran (ancient “Media/Māda” province) stretching across SE Āzarbāijān to Hamédān, Borujerd, Golpāyégān, Esfahān, Kāshān, Ray, and the western Zagros (Khorramābād, Saimareh) regions (see Le Strange, 1905). † Controversy on the exact date. There have been some differences of opinion as to whether the Hamédān ‘Alaviān Tomb Shrine (Gonbad ‘Alaviān) was erected in the Saljuq period (before 1200 B.C.E.; Mostafavi, 1953, 1967; Meshkati, 1970; Hātami, 1379/2000), or whether it is a monument of the Mongol period (ca. 1309–1316 C.E.: Wilber, 1955). Herzfeld (1922) dated the monument A.D. 1309–1316, whereas Wilber (1955) gave ca. 1315. It is probable that the monument was rebuilt after the 1316 C.E. earthquake. § Periā (in Armenian; home of the Safavid-era Armenian settlement; Pereidani, Martqopi, in Georgian); the modern Faridan District with central town of Feraidunshahr (32°56′N, 50°07″E, +2548 m). # The earlier date of A.D. 224–549 (Berberian and Yeats, 2001) was bracketed on data by Kāmbakhsh-Fard (1994). The present dates are bracketed from new data by Āzarnoush (2009).

Māda/Mād) province, Qāsem ben ‘Ajali Isā Abu-Dolaf (767–841 C.E.). In order to visit the construction site of the future Borujerd congregational mosque I, both Hamuleh and Abu-Dolaf went to the site sometime before 841 C.E., which was located adjacent to the destroyed Sāssānid Zoroastrian fire temple (destroyed during the invasion or immediately after occupation of the country by the Moslem Arabs) and referred to the existing “early mosque” as “hāzā masjed al-‘atiq” (lit. in Arabic: “this is the old mosque”), possibly already in a dilapidated state (Estakhri, 951, p. 165; Ibn Hawqal, 978, p. 258, 262; Yāqut, 1225, I:596, II:737; Hazin Borujerdi, 1972, p. 82; Table 2). As with the “early mosque,” the Borujerd congregational mosque I was a “Khorāssāni-style architecture” (Pirniā, 2003), void of ornaments and luxury, and many exertions were made to avoid vanity. For an unknown reason, the ninth-century Borujerd congregational mosque I was destroyed prior to ca. 971 C.E. (?) when the Buyid-era (ca. 945–1055 C.E.) congregational mosque II was built on the site. Cause of Destruction of Congregational Mosque I (Pre–971 C.E.) No information regarding the cause of destruction of the congregational mosque I is provided by the few surviving chronicles. Paradoxically, around the same time, there was a damaging earthquake in 956–957 C.E. in Hamédān (102 km to the NNW of Borujerd; Fig. 6, inset top right), ‘Abbāsābād (114 km NW of Borujerd), and their districts, where many people perished beneath the debris; the walls of Qasr-e Shirin (270 km WSW of Hamédān, 302 km WNW of Borujerd) fissured (Ibn al-Athir, 1231, VIII:388; al-Suyuti, 1505, p. 29; Sani’al-Dauleh, 1880–

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084761/spe525-05.pdf by University of Wisconsin Madison user

1882, I:131; Ambraseys and Melville, 1982; Berberian, 1994; Tables 2 and 3). Unfortunately, the epicentral area of this historic earthquake is not known. It is probable that the earthquake took place along the Zāgros Main Recent fault (see following) located to the SW of Hamédān (Figs. 2 and 6); however, there are not enough data at this stage to prove the link. It is tempting to think that possibly either (1) this distant event (with unknown seismic parameters and location), or (2) an earthquake that originated in the neighboring High Zāgros Mountains to the SW and S of Borujerd (such as the source region of the 4 December 1955 Ms 6.0 and 17 December 1955 Mb 5.5 Razan earthquakes, ~30 km to the S of Borujerd; see Fig. 6) damaged the Borujerd congregational mosque I (ninth century C.E.). Whether or not this was the effect of an earthquake or natural wear and tear of the structure remains a matter for speculation until further careful investigation is carried out. Borujerd Buyid-Era Congregational Mosque II (Ca. 945–971 C.E.) The floor of the Buyid-era Borujerd congregational mosque II (945–971 C.E.) was discovered at –180 cm below the present floor of the mosque (Fig. 5; Table 2). The floor was composed of a 10 cm gypsum plaster layer (from –180 to –190 cm) constructed upon 40 cm of black debris with potsherds of unknown age (from –190 to –230 cm). The Sāssānid (224–642 C.E.) bricks of the Zoroastrian fire temple destroyed by the Arab invaders were reused in irregular brick work in construction of the Borujerd congregational mosque II (Lotfizādeh, 1980; Mehryār, 1985; Moqaddas, 1997).

186

Berberian et al.

Figure 5. Discovered older floor levels in a small sounding underneath the present floor level (0.0 cm) of the Borujerd congregational mosque: The Safavid period floor (1658 C.E.) was at –20 cm, floors for mosque II (ca. 945–971 C.E.) and mosque III (ca. post–1000/pre– 1139 C.E.) were at –180 cm, an older lower floor was at –230 cm, and virgin soil was detected at –420 cm. Note the deep stone mortar foundation (–190 to –405 cm) constructed for the columns of mosque III (ca. post–1090/pre–1139 C.E.) into the remains of mosque I (the ninth century, destroyed before 971 C.E.) to the lower left. Tr. Tiles— Saljuq-period (ca. 1000–1090 C.E.) turquoise decorative tiles. Figure is modified after Mehryār (1985).

Approximately 2.0–2.5 m of the four peripheral Buyid-era walls of the dome chamber were preserved and detected in the excavations that were constructed at elevation –180 cm below the present floor of the mosque (Fig. 5). Remnants of (1) the Buyidera main brick piers, one with brick work decorative inscription (Figs. 5 and 7); (2) the lower sections of the round-based pseudocolumns with decorative brick motif (Fig. 7, lower right; Figs. 8

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084761/spe525-05.pdf by University of Wisconsin Madison user

and 9); (3) a damaged ribbed vault and collapsed arch (Figs. 5, 10, and 11); (4) a collapsed arch (Fig. 12); (5) a tilted column base (Fig. 13); (6) the brick decorations; (7) a mehrāb (with no gypsum stucco decorations, at elevation lower than the twelfth– thirteenth-century mehrābs with gypsum stucco); and (8) a Buyid-era golden Dinār coin dated 360 H/971 C.E. (Fig. 5) were unearthed as elements of the Buyid-era Borujerd congregational

Evidence of historical seismic activity along the Zāgros Main Recent fault at Borujerd

Figure 6. Recorded seismic history of the Borujerd area in the vicinity of the Zāgros Main Recent fault (ZMRF), with Nahāvand and Dorud fault segments to the northwest and southeast of Borujerd (segment names in parentheses). Wellconstrained meizoseismal areas of the earthquakes are shown with ellipses. Focal mechanisms of the 2006 Mw 6.1 Chālānchulān earthquake constrained by body wave modeling with centroid depth of 6 km (Peyret et al., 2008) and 2005 South Borujerd earthquake (best-double-couple Harvard Centroid Moment Tensor [CMT] solution) were also added. The 28 October 1961 M 5.0 and 2005 Mw 5.0 earthquakes occurred almost in the same area and were followed by the 2006 Mw 6.1 earthquake. Inset top right: Regional historical (pre-1900) earthquakes (956–957, 1087, 1135, 1191, 1310, 1316, 1430, and 1495 C.E.; discussed in the text) with unknown epicentral area in the medieval provinces of Jebāl (Media/ Māda) and Kordestān in western Iran. Historical cities of Hamédān and Golpāyégān mentioned in the contemporary earthquake report are underlined. Figure is modified from Berberian (2005, 2014). AZ—Āzarbāijān; KP—Kopeh Dāgh; M—Makrān; MZRF—Main Zāgros reverse fault; S—Sistān suture zone; TP—Turān plate.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084761/spe525-05.pdf by University of Wisconsin Madison user

187

188

Berberian et al.

Figure 7. Structural remnants of the Borujerd congregational mosque II (ca. 945–971 C.E.; with –180 cm floor) and mosque III (ca. post–1000/ pre–1139 C.E.; at minus –120 cm) that collapsed during the early fourteenth-century earthquake underneath the present structure of the Borujerd congregational mosque. Lower-right panel: N–S profile in trench 1. Upper-right inset: E–W profile in trench 1. Location of trench 1 is shown by a filled area on the lower-left plan. 0.0—the 1795–present floor. Note the remnants of the destroyed mosque II (ca. 945–971 C.E.) column base (lower right), collapsed brick vault with missing arch (possibly mosque III?), as well as scattered collapsed bricks amidst debris material (lower and upper right). The collapsed structures are covered by mixed fill and debris with the 1615 minted coin. Figure modified from Mehryār (1985).

mosque II (Figs. 5, 7, and 13). The discovered coin may represent a terminus post quem date possibly close to the occurrence of the destruction of mosque II. As with other structures of the “Rāzi architectural style” in Iran (Pirniā, 2003), the Buyid-era congregational mosque II (ca. 945–971 C.E.) was much larger with a higher dome (exaggerating the height of the building) than those of the previous and/ or the present mosque, with employment of arches and domes. The structural elements and architectural style of the destroyed

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084761/spe525-05.pdf by University of Wisconsin Madison user

quadrangular Borujerd Buyid-era congregational mosque II resembled those of the 907 C.E. extant Amir Esmā’il Sāmānid mausoleum in Bokhārā (Pope, 1965; misspelled Bukhara in the literature) of the Iranian Soghd (ancient Soghdiānā) province, northeast of the Greater Khorāsān province, now in Uzbekistan (Mehryār, 1985; Moqaddas, 1997). Apparently, sometime between the Buyid (945–1055) and the Saljuq (Seljuk; ca. 1000–1218) eras, the Borujerd congregational mosque II structure collapsed, i.e., around pre–971 C.E.

Evidence of historical seismic activity along the Zāgros Main Recent fault at Borujerd

Figure 8. Experimental trench dug underneath the floor of the eastern ayvān (porch; –20 cm) of the present Borujerd congregational mosque (tiles at the top of the picture). Note the remnants of the mosque II and III collapsed dome chamber underneath the present mosque decorated with colored tiles (upper section of the figure): (i) parts of the mosque III collapsed brick arch (left), and (ii) basal section of the mosque II round-based pseudocolumn (right). The latter was built upon the ruins of an earlier mosque I. See Figure 7 for cross section. Photograph was scanned from a print in Mehryār (1985). The dimensions of the bricks are 20 × 20 × 4 cm and the outer diameter of the destroyed round-based pseudocolumn to the right (mosque II) is ~70 cm. Looking to the east.

Figure 9. Tilted and bulged basal section of the collapsed round-based pseudocolumn of the Buyid mosque II (ca. post–945/pre–971 C.E.) (right), and broken brick arch of mosque III (left) beneath the present congregational mosque. Piles of bricks to the left underneath the damaged arch were added during excavation to support the destroyed structure. Photograph was scanned from a print in Mehryār (1985). The dimensions of the bricks are ~20 × 20 × 4 cm and the outer diameter of the destroyed round-based pseudocolumn column to the right (mosque II) is ~70 cm. Looking to the east.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084761/spe525-05.pdf by University of Wisconsin Madison user

189

Figure 10. Remnants of the collapsed load-bearing brick ribbed vault of mosque III (ca. post–1090/pre–1139 C.E.) of the dome chamber underneath the present mosque (the colored tiles in the upper portion of the figure). Photograph was scanned from a print in Mehryār (1985). The dimensions of the bricks are ~20 × 20 × 4 cm. Looking to the west.

Figure 11. Remnants of the Borujerd congregational mosque II (ca. 945–971 C.E.) brick tilework (“decorative brick motif”; see Fig. 5) with the mosque III (ca. post–1090/pre–1139 C.E.) collapsed brick arch and ribbed vaults underneath the present mosque (colored tiles). Photograph was scanned from a print in Mehryār (1985). The dimensions of the bricks are ~20 × 20 × 4 cm. Looking to the south.

190

Berberian et al. of a massive and very deep foundation (–180 to –405 cm) with the local black phyllite for congregational mosque III (Fig. 5) is also not clear. Was the mosque III (ca. 1055–1218 C.E.) architect aware of stability problems of mosque II and/or the site? We can rule out a site-specific problem, since groundwater level at the site is reported ~15 m below the grades (Mehryār, 1985), and no evidence of sandy deposits and liquefaction potential is found at the city. An earthquake could be responsible for the damages and destruction of the Borujerd Buyid-era congregational mosque II; however, we cannot confidently assign it to a particular earthquake discussed in this paper without further archaeological excavations and paleoseismic trench study across the Zāgros Main Recent fault. Borujerd Saljuq-Era Congregational Mosque III (Ca. Post–1090/Pre–1139 C.E.)

Figure 12. Collapsed and broken arch southeast of the dome chamber of mosque II (ca. 945–971 C.E.) underneath the present congregational mosque (colored tiles). Photograph was scanned from a print in Mehryār (1985). The dimensions of the bricks are ~20 × 20 × 4 cm. Looking to the southeast.

The following structural deformations and remains of congregational mosque II have been detected underneath the present floor of the mosque (Figs. 5, 7, and 13): (1) The Buyid-era eastern brick columns (945–971 C.E.) are tilted toward the south/southeast (Fig. 13); (2) the southern façade and top of the round-based pseudocolumns with “decorative brick motif” and pillar were destroyed (Figs. 5, 7, 8, 9, and 13); (3) the basal section of the round column is bulged (Figs. 7 and 9); and (4) the upper parts of the round-based pseudocolumns with decorative brick motif collapsed (Figs. 5, 7, 8, and 10; Mehryār, 1985). Cause of Destruction of Congregational Mosque II (Ca. 945–971 C.E.) The exact time and cause of tilting of the pillars and destruction of the structure are unknown. The reason for the construction

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084761/spe525-05.pdf by University of Wisconsin Madison user

Atop of the 2.0–2.5 m lower remnants of the four Buyidera walls of the Borujerd congregational mosque II dome chamber (ca. 945–971 C.E.), the Borujerd Saljuq-era (ca. 1055– 1218 C.E.) congregational mosque III, with very thick walls and a mighty dome chamber, was constructed (Fig. 3; Table 2). A deep massive footing, mainly composed of black phyllite with lime mortar, was discovered for the mighty structure (Fig. 5); such material was never used in the earlier or later stages of the structure. The deep massive masonry stone foundations of the pillars of the Borujerd congregational mosque III were dug into the previous phases of –190 to –405 cm construction (Fig. 5, lower-left side). A mehrāb decorated with gypsum stucco at elevation higher than the elevation of the Buyid-era mehrāb (mosque II, referenced earlier), around –150 cm below the present floor (probably the apparent elevation of the Saljuq-era floor), was also discovered. Remnants of the Saljuq-era glazed turquoise decorative tiles (used inside and outside the structure) were also discovered above the Saljuq Kufic inscription band (Figs. 3 and 14) and around the mehrāb arch (Mehryār, 1985; Moqaddas, 1997). The exact date of construction of the congregational mosque III of the Saljuq period is not known. We know that the Esfahān Saljuq congregational mosque (Esfahān became the capital city of the Saljuq dynasty in 1051 C.E.) was built ca. 1075–1088 C.E. during the premiership of Nezām al-Molk (1063–1092 C.E.), the Iranian grand vizier of Malek Shāh I Saljuq (1072–1092 C.E.), whereas the Saljuq congregational mosques in the provincial cities were built decades or nearly a century later: Qazvin (1113– 1115 C.E.), Barsiān (Pārsiān; 1097–1098 C.E.), Golpāyégān (1118 C.E.), Gonābād (1140–1141 C.E.), Zavāreh (1153 C.E.), Ardestān (ca. 1180 C.E.), and others (Pope, 1965). We, therefore, may assume that the Borujerd congregational mosque III was built after completion of construction of the congregational mosque of the capital city of Esfahān, sometime after ca. 1090 C.E. (post–1090/pre–1180 C.E.). The oldest surviving exquisite wide horizontal inscription band with Kufic script composition of chiseled and stamped bricks

Evidence of historical seismic activity along the Zāgros Main Recent fault at Borujerd

191

Figure 13. Structural remnants of the collapsed Borujerd congregational mosque II (ca. 945–971 C.E.) and mosque III (ca. post–1090/pre–1139 C.E.) beneath the present floor of the 1795–present congregational mosque at SE trench (see Fig. 7, lower-left plan for the location and lower-right panel for description). Note the collapsed broken brick vault (mosque III?), and tilted column base (mosque II) with its collapsed upper section surrounded and covered by debris and fill material. Eastern view of the southeastern excavation. Figure is modified from Mehryār (1985).

underneath the octahedral drum transition zone (Fig. 3), on top of the southern square plan wall above the mehrāb (Fig. 14), reads: In the name of Allāh. O Allāh, forgive the commissioner of this lofty dome,… Majd al-Din H … Abul-‘Ez Mohammad, son of Tāher, son of Sa’eed. May Allāh perpetuate his existence?

The middle section of the epigraph is missing (in ellipses above), and there is no date associated with it (Fig. 14). Based on detailed analysis of the inscription by Moqaddas (1997), the inscription could date to ca. 533–539 H/1139–1145 C.E., during the premiership of Majd al-Din ‘Ez al-Molk Abul-‘Ez al-

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084761/spe525-05.pdf by University of Wisconsin Madison user

Borujerdi (Mehryār, 1985). The timing was during the reign of Sultān Sanjar (r. 1118–1153), the last Saljuq ruler in Iran, indicating the Saljuq period of the Borujerd congregational mosque III. It should be noted that the aforementioned inscription was added as a layer to the existing structure during a major renovation phase ca. 1139–1145 C.E. The inscription: (1) supplicates Allāh to forgive the commissioner (the original builder) of the “lofty dome” and (2) praises the present ruler (Abul-‘Ez) for the “repairs” carried out in the twelfth century C.E. It evidently shows that unlike the present congregational mosque (IV; see following section), the Saljuq-era mosque III had a “lofty dome” in the early twelfth century C.E. It is clear that the mosque III

192

Berberian et al.

Figure 14. The oldest Arabic inscription of mosque III (ca. post–1090/pre–1139 C.E.) underneath the early fourteenth-century collapsed octahedral drum transition zone, on the southern square wall of mosque III above the mehrāb with Kufic script composition in chiseled and stamped bricks. The inscription survived the 1316 C.E. earthquake, but the dome chamber structures above the square wall transition zone collapsed. See Figure 3 for the location (Mehryār, 1985; also see Moqaddas, 1997). The vertical length of the largest Arabic letter brick inscription is about 44 cm. For translation and interpretation of the inscription see mosque III section in the text.

was already extant, and the lofty dome above the inscription was repaired or reconstructed by ‘Ez al-Molk. The extreme thickness of the load-bearing square walls below the octahedral drum transition zone (Fig. 3) also indicates that they were constructed in proportion to the height and weight of a “lofty dome.” Basal remnants of a single brick minaret (with inner diameter of 170 cm and brickwork thickness of 80 cm) located outside the mosque to its SW (Fig. 15) belong to the Saljuq-era congregational mosque III (Moqaddas, 1997). Furthermore, it should be noted that, as with other structures of the Rāzi architectural style in Iran (Pirniā, 2003), the Saljuq-era congregational mosque III was characterized by exaggeration of the height of the building with employment of arches and larger and higher domes. Fate of Congregational Mosque III (Ca. Post–1090/Pre–1139 C.E.) Circa 1139–1145 C.E. Renovations The major repair work of ca. 1139–1145 C.E. must have been a major renovation work worthy of dedication by an inscription added to the structure. No data on the cause of damage are recorded in the few surviving old chronicles, and we do not know the exact cause and extent of damage. Interestingly, the major

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084761/spe525-05.pdf by University of Wisconsin Madison user

repair work on the congregational mosque III (1139–1145 C.E.) was carried out a few years after the three regional destructive earthquakes of 1130 and 1135 C.E. (total of three events4; see Fig. 6, inset top right; Table 3), which were strongly felt from Mesopotamia (in the west) to the Jebāl (Media/Māda) province

4 On 27 October 1130 C.E., a destructive earthquake affected Baghdād (ground and buildings swelled like the sea, houses collapsed), Iraq, the Jebāl [Media/ Mād] province, Mosul, al-Jazira (ancient upper Mesopotamia with city of Mosul of Kordestān in the western Zāgros), and Mesopotamia (Ibn al-Athir, 1231, X:469; Ibn al-Jauzi, 1181, X:14; Bar Hebraeus, 1286, p. 255/289; Nabavi, 1972; Ambraseys, 1974; Ambraseys and Melville, 1982; Berberian, 1994). It seems that the widespread damage from Mesopotamia (in the south) to al-Jazira (in the north) and Jebāl (in the east; possibly the western Zāgros/Kordestān; Fig. 2) was caused by more than one earthquake. Nonetheless, exact epicentral locations and seismic parameters are unknown. 25 July 1135 C.E.: A strong earthquake affected Iraq (Baghdād, some walls collapsed) and other places (Mosul, Jebāl), followed by numerous aftershocks; many people were killed (Ibn al-Jauzi, 1181, X:46; Ibn al-Athir, 1231, XI:22; al-Suyuti, 1505, p. 37; al-Qusi, 1907, p. 95; Ambraseys and Melville, 1982; Ambraseys, 2009). The earthquake seems to have been widely and strongly felt in Mesopotamia, al-Jazira (Mosul; Kordestān in the western Zāgros), and Jebāl. Exact epicentral location and magnitude are unknown. 13 August 1135 C.E.: This seems to be the strongest aftershock of the 25 July event, which was widely felt in Baghdād, where some walls collapsed (Ibn al-Jauzi, 1181, X:46; Sani’ al-Dauleh, 1880–2, I:191; al-Qusi, 1907, p. 95; Ambraseys and Melville, 1982; Berberian, 1994).

Evidence of historical seismic activity along the Zāgros Main Recent fault at Borujerd

Figure 15. Invaluable photograph of basal remnants (below the present ground surface) of the collapsed mosque III (ca. post–1090/pre– 1139 C.E.) free-standing minaret, showing deep vertical fractures cutting and displacing the bricks. The minaret remnant was accidentally discovered, and it is not clear if mosque III had the second minaret. Photograph was scanned from an old low-quality print in Moqaddas (1997); the original negative or a proper print was not available and unfortunately has been lost. The dimensions of the bricks are ~20 × 20 × 4 cm. The inner diameter of the basal remnant of the collapsed minaret is ~170 cm, the thickness of the minaret brick wall is ~90 cm, and the outer diameter is estimated to be ~260 cm. Looking to the northeast.

(in the east; Figs. 2 and 6), where Borujerd is located (see relevant references in Table 3). Although it is probable that the Borujerd congregational mosque III (ca. post–1090/pre–1139 C.E.) was damaged by a strong distant earthquake (Tables 1 and 3), we cannot prove it at this stage of our study, and the case is mentioned for further research work at the site. Circa Early Fourteenth-Century Destruction Long after the ca. 1139–1145 C.E. renovations with the Saljuq inscription described above, sometime around the early fourteenth century, the Saljuq-era lofty dome chamber (Fig. 3) and the minaret (Fig. 15) of the Borujerd congregational mosque III simultaneously collapsed (Mehryār, 1985; Moqaddas, 1997). The existing structural record clearly shows the following simultaneous structural damage and destruction: (1) complete collapse of the dome chamber above the square walls with the Saljuq Kufic inscription (ca. 1139–1145), including all the ribbed/transverse-vaults (“tavizeh”), the octahedral and the sixteen-sided drum transition zones with their related arches, and the lofty dome (Fig. 3); (2) destruction of the majority of the Saljuq-era glazed turquoise decorative/inscribed tiles used above the Saljuq Kufic inscription band and above the mehrāb inside the structure, as well as outside the mosque; (3) collapse of the upper parts of the square load-bearing walls above the inscription band (Fig. 3); (4) collapse of the minaret, remnants of which were discovered outside the mosque precinct to its southwest, and

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084761/spe525-05.pdf by University of Wisconsin Madison user

193

development of deep vertical fractures in the remaining basal section of the minaret, indicating the severity of the damage (Fig. 15); (5) collapse of the gypsum stucco around the vaults; and (6) collapse of the vaults of the dome chamber, as indicated by remnants in the basement debris of the dome chamber (with potsherds covering periods from the eighth to twelfth century C.E., Saljuq-era glazed turquoise decorative/inscribed tiles, and Saljuq-Ilkhānid architectural style structures; Figs. 5, 7, and 13). On the bases of numismatic material, potsherds, and structural style, and the absence of any radiometric dating, the timing of sudden destruction of the lofty dome chamber of the Saljuq congregational mosque III and its minaret has been roughly reported as ca. pre–fourteenth century, possibly early fourteenth century (Mehryār, 1985; Moqaddas, 1997). Although both excavators reached the conclusion that the mighty mosque most probably collapsed as a result of a strong earthquake, they could not assign the destruction to any dated earthquake because no earthquake was listed in the Iranian catalogues occurring in the city of Borujerd (Ambraseys and Melville, 1982; Berberian, 1994, 1995, 2014). In his treatise History of Oljāitu, the contemporary Iranian historian and writer of the Court of Sultān Abu Sa’eed BahādorKhān Ilkhānid (r. 1316–1335 C.E.) at Tabriz, Abolqāsem ‘Abdollāh ibn Mohammad al-Qāshāni ([sic] Kāshāni, ca. 1320, p. 179), wrote in Persian that: Events of this year: on Shawwāl first of the year 715 [29 December 1316], in the ‘Siā…’ (?) District of Hamédān and Jorfādqān [Arabicized Golpāyégān; lit. “the Place of Flowers”], continuous rain lasted for seven days and nights. On the eighth day [i.e., 8 Shawwal 715 H/ January 5, 1316 C.E.], there was a severe thunderstorm and lightning, and a great earthquake occurred. More than twenty large and important villages were overwhelmed and left totally ruined, and the land was flattened to the ground level [referring to Qur’anic expression of the Last Day; Āyat 106, Sura Tah 20]. A large number of peasants and farmers were killed under the debris/soil, and their possessions were buried under and covered by the ground. (Kāshāni, ca. 1320, p. 179; translated from the original Persian text; emphasis added; the book covers the history of Iran from 1305 to 1318 and was written during the first years of Sultān Abu Sa’eed, i.e., ca. 1320 C.E.)

This is the only surviving first-hand contemporary account of the 5 January 1316 earthquake. A search in the Iranian geographic directory since 951 C.E. (Estakhri, 951; Ibn Hawqal, 978; Anonymous, 982; Yāqut, 1225; Zakaryā Qazvini, 1275; Mostaufi Qazvini, 1340; Gazetteer of Persia, 1885–1918; Le Strange, 1905; Iranian Gazetteer, 1951) failed to locate the “District of Siā…” in the Hamédān, Lorestān, and/or Ilām provinces, as well as in the Golpāyégān region of western Iran (Fig. 2). There are numerous village names starting with “Siā,” but no “District” name was found starting with the three letters of “Siā” ( ). For example, the village of Siāhkamar is located ~22 km to the ESE of Hamédān (34°44′N, 48°44′E, +1945 m); however, it is not located in a large and populous agricultural area, and it

194

Berberian et al.

is far from any active fault, whereas the village of Siāh Kalleh is located in the fertile Silākhor Valley, ~5 km to the NW of the town of Dorud at 33°31′N, 49°00′E. Nevertheless, the latter is a village name and not a district (Fig. 16). The recorded “Siā…” (?) [… ] District of Hamédān and/ or Jorfādqān is definitely corrupt in this old handwritten manuscript, and some letters are missing from copying the original text of ca. 1320 C.E., and the later copiers replaced the illegible

or missing letters by a few dots. Hence, there is no indication of existence of a “district” name starting with the three letters of “Siā.” Mostaufi Qazvini (1340), finishing his compilation of data ~25 yr after the earthquake, did not mention the event or a district starting with “Siā….” However, he wrote that there are two mosques in Borujerd: “the old” (“atiq” in Arabic: the present congregational mosque belonging to the Shi’a sect) “and the new” (“hadith” in Arabic, belonging to the Sunni sect, upon

Figure 16. Active fault map of the Borujerd study area (filled circle) in the northern part of the Silākhor Valley/Plain (stippled) along the central section of the Zāgros Main Recent fault (ZMRF; the Nahāvand and Dorud segments). Locations addressed in the paper are added. SF—Sang-e Sefid rockfall. Underlined cities of Hamédān and Golpāyégān mentioned in the contemporary description of the 1319 C.E. strong earthquake are highlighted. Underlined town of Aznā and the proposed macroseismic epicenter of the 5 January 1316 earthquake noted by Ambraseys and Melville (1982; indicated by “*A&M (1982)” are also shown (see the text for analysis). The erroneous epicenter is located ~10 km north of the town of Aznā in an area with no active fault and no evidence of destruction/damage. Inset top right: contour map (elevations in m) of the Silākhor area located between the Zāgros fold-and-thrust belt to the SW and Central Iran to the NE. Symbols as in Figure 1. MZRF—Main Zāgros reverse fault.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084761/spe525-05.pdf by University of Wisconsin Madison user

Evidence of historical seismic activity along the Zāgros Main Recent fault at Borujerd which the present Soltāni mosque was built in 1794; Mostaufi Qazvini, 1340, p. 74). It is not known exactly when and how the so-called “new” Sunni mosque was destroyed; Mostaufi Qazvini (1340) repeated the old statements in his compilation without visiting the area or citing his sources. We know that Mostaufi (1340) was a state accountant (hence his name “Mostaufi”), and he compiled his book based on much earlier accounts and was not contemporary or up-to-date in 1340. This is why he missed reporting the 1316 earthquake and its effects. Furthermore, he resided in Qazvin and did not visit the numerous cities, towns, and villages referred in his 1340 chronicle. It is important to note that one of the largest and important agricultural plains in the region (surrounding the towns of Borujerd at its northern edge and Dorud to the SSE) is the “Silākhor” (ā“Sailāb-khor”; lit. “flood-prone area”) Valley (Fig. 16). It is highly probable that the first three surviving Persian letters of this district (“Sil”) could have been corrupted to “Siā.” In this case, the cursive Persian lowercase letter of “lām” (“l”) is corrupted to “alef” (“ā”), and the remaining left segment of the Persian letter “lām” (“l”) as well as the rest of the letters (“ākhor”) of the district name have disappeared in the 700-yr-old manuscript (the cursive Persian lowercase letters “l” and “ā” are very similar to each other). Le Strange (1905, p. 200), quoting Mostaufi Qazvini (1340), stated that the town of Borujerd “was then already falling to ruin.” However, such statement was not mentioned by Mostaufi Qazvini (1340, p. 58, 74, 166) but was addressed by later accounts. Sani’ al-Dauleh (1880–1882, II:23), without quoting his source(s), also briefly referred to an earthquake in Hamédān in 725 H/1325 C.E., which is 10 yr off the date given by contemporary scholar Kāshāni (ca. 1320, p. 179). Eugène Flandin and Pascal Coste, who were in Iran in 1839–1841, wrote that the city of Borujerd had numerous ruined buildings, and the prince did not agree to live there

(Flandin and Coste, 1851). Sani’ al-Dauleh (1881, p. 240), who was in Borujerd in 1859, stated that the houses in Borujerd were in ruins, but the city had a rampart with five gates and a moat. A painting presented by Eugène Flandin (in Flandin and Coste, 1851) shows development of deep fractures in the bastion of the thick city rampart to the right of the city gate (Fig. 17). Quchāni (1988), without referring to his source(s), speculated that: “the 8 Shawwāl 715 H [5 January 1316 C.E.] earthquake, which took place in the area between Hamédān and Golpāyégān, possibly destroyed the town of ‘Karaj [Karah/Qarah Su]-e Abu Dolaf.” The exact location of the latter is unknown; apparently Karaj was located ~60 km ENE of Borujerd (Karaj east of Borujerd in Fig. 2) and was the capital of the “Ighārain” (the two Ighārs in Arabic; lit. “fiefs in perpetuity”) district (see Ibn Hawqal, 978, p. 258, 262; Yāqut Hamavi, 1225, I:420, I:548, III:873, IV:250, IV:270; Mostaufi Qazvini, 1330; Le Strange, 1905, p. 197, 198). The distance between the cities of Hamédān and Golpāyégān is ~225 km (Fig. 16), and the meizoseismal area of the 1316 earthquake (intensities greater than VII on the modified Mercalli intensity [MMI] scale) was definitely much shorter than the mentioned distance between the two cities. Without presenting any supporting macroseismic evidence, Ambraseys and Melville (1982) located the macroseismic epicenter (center of the maximum damage area) of the 5 January 1316 earthquake (with estimated magnitude Ms ~6.2) at 33.5°N, 49.4°E, ~10 km to the north of the town of Aznā (75 km to the SE of Borujerd) and 30 km to the NE of the Zāgros Main Recent fault, in an area where no seismic fault exists (see the asterisk with “A&M [1982]” in Fig. 16). As we will see in this study, their epicentral location as well as the estimated magnitude are not, therefore, warranted anymore. Nonetheless, it is very important to note that the four broad locations given by (1) the contemporary historian Kāshāni (1320, p. 179) in the area between Hamédān and Golpāyégān

Figure 17. Deep fractures developed in the Borujerd city rampart bastion near the Borujerd city entrance gate portrayed by Eugène Flandin ca. 1839–1841 (Flandin and Coste, 1851); arrows were added. Courtesy of wikipedia.org (Wikipedia, 2016).

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084761/spe525-05.pdf by University of Wisconsin Madison user

195

196

Berberian et al.

(Figs. 6 and 16); (2) Sani’ al-Dauleh (1879, II:23) near Hamédān; (3) Ambraseys and Melville (1982) near Aznā; and (4) Quchāni (1988) in the Karaj are all around the Silākhor Valley and the city of Borujerd near the Zāgros Main Recent fault (Fig. 6). It is, therefore, not inappropriate to see if the 1316 earthquake might have happened near Borujerd, especially if we find no evidence of major destruction by the 1316 earthquake in the aforementioned locations or anywhere in Iran. In the mid-1980s, a set of large rockfalls, with undated ancient petroglyphs, was detected in the foothills near Sang-Sefid (also known as Shāhdāgh, Shāh Bulāgh, Shāh Budāgh, located in the Shavdāgh region of Darreh Saydi, Oshtorinān district, ~10 km to the NE of Borujerd), being detached from the mountain and rested on the mountain foot (Fig. 18, SF in Fig. 16). A recent search showed that the rockfalls cover a wider area and were detected at Qapānvari, Tappeh (mound), ChoghāKāfarān (Dehkord), Daudāngeh, Fakhrābād, Bijun Bālā, and Vālānged villages, located to the NE, E, and SE of Borujerd. A single large landslide block that was resting in an open area near the village of Vālāngerd has been stolen. The time of this “possibly earthquake-induced” widespread rockfall is unknown, and it is tempting to link it to the 1316 strong earthquake with an epicenter close to the area; however, we do not have any data to prove it. Perhaps related, legend has it that the Sang-Sefid village (18 km NE of Borujerd in the Darreh Saydi rural district; “SF” in Fig. 16), located in the vicinity of the “old city of Budāgh,” was destroyed by a historical earthquake. (“Budāgh” is the Persian name for “snowball bush” [Viburnum opulus Sterile from the Caprifoliaceae family].) Illegal diggings in the region have revealed numerous old artifacts. The massive structural damage and deformation of the monument were not caused by anthropogenic action, various climate-related effects, or uneven settlement of the structure. All

Figure 18. Large historical rockfall (date unknown) with ancient petroglyph (at the foot of the Sang-Sefid Mountain, 10 km to the NE of the city of Borujerd) (SF in Fig. 16). Each increment of the ranging rod represents 10 cm.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084761/spe525-05.pdf by University of Wisconsin Madison user

the structural failure evidence shows that the structure, located in the Zāgros Main Recent fault zone (Fig. 16), collapsed because of strong earthquake ground motion. Except for alterations such as restorations and architectural modifications and construction of additional wings throughout history, addressed in this paper, anthropogenic damage or destruction of a congregational mosque has rarely been reported, unless during invasion by hordes of different religion and ethnicity. However, even in such a rare circumstance, it is impossible to destroy a properly constructed lofty dome chamber and minaret without leaving any evidence in treatises or on the ground. The approximate estimated time of sudden collapse coincides with the occurrence of the 1316 earthquake near the city (described earlier; Tables 1 and 3); this seems a more reliable hypothesis to explain the observed sudden and simultaneous collapse of the Saljuq-era congregational mosque III (1000–1218 C.E.) lofty dome chamber and its minaret. The survival of the lower parts of the massive structural walls clearly indicates that the structure did not experience a geotechnical foundation problem, such as failure of its foundation due to uneven settlement, sliding, or liquefaction triggered by an earthquake. The structure seems to have suffered intensities of VIII and/or greater generated by an earthquake along the nearby active Zāgros Main Recent fault (Fig. 6). It seems that the earthquake damaged a relatively large area somewhere to the SE of the town of Hamédān and WNW of Golpāyégān (covering the Silākhor district in between; Fig. 2), where no pre–fourteenth-century structure is standing.

Figure 19. Damage caused to the northern portico, minarets, and inner portal on the northern yard (built in 1794 C.E.; see Fig. 4 for location) of the Borujerd congregational mosque by the 31 March 2006 Mw 6.1 Chālānchulān earthquake, located ~20 km south of the city (see Fig. 6). Note the shear fractures developed in the minarets. The top of the left minaret (turnet/goldasteh) collapsed during the earthquake (courtesy of http://en.wikipedia.org/wiki/Jameh_Mosque_of _Borujerd). The height of the minaret is 16.75 m up to the base of the minaret turnets (‘goldasteh’ in Persian; see also Figs. 3 and 4); the outer and inner diameters are 3.5 and 2 m, respectively. Looking to the south.

Evidence of historical seismic activity along the Zāgros Main Recent fault at Borujerd Survival of the earthquake news in the literature alone shows that it was probably a large-magnitude earthquake. Considering the destruction of monuments and the city of Borujerd by the 1316 earthquake, some important notes must be made here: (1) No pre–1316 C.E. standing historical monument exists in the city of Borujerd, the Silākhor Plain, and the surrounding towns of Dorud and Nahāvand or the villages in the area (Figs. 2, 6, and 16; Table 3). (2) As with the Borujerd congregational mosque, both mausoleums of Ja’far of Borujerd and Qāsem of Jāpalaq/ Jāpaleh in Aznā (Fig. 16) underwent numerous periods of repairs, renovations, and rebuilding. However, the causes are not documented, and no excavation or detailed architectural or archaeological investigations have been carried out in the latter two monuments. (3) As with the Borujerd congregational mosque IV (1405– 1446 C.E.), the mausoleums of Qāsem of Jāpalaq in Aznā (ca. 1336 or 1405–1446 C.E.; Golombek and Wilber, 1988; Fig. 16) and Ja’far of Borujerd (ca. 1125 C.E.; Meshkāti, 1970) were (re)built after the 1316 C.E. earthquake. (4) In the case of the Borujerd mausoleum of Ja’far, a contemporary inscribed poem in the mausoleum with a chronogrammatic verse at the last line (Meshkāti, 1970) states that the mausoleum, which “was ruined from damages in this crumbling region [‘Shekasteh bood ze āsib-e in Kharābābād’], was rebuilt in 519” [H/1125 C.E.]. Details of destruction and the post–1316 C.E. repairs have not yet been investigated in this monument. It is probable that this structure was damaged by the distant earthquake of 8 December 1058 (Ibn al-Jauzi, 1181, VIII:190; Ibn al Athir, 1231, IX:449; Abu’l Fida, 1321, II:188; al-Suyuti, 1505, p. 33) and/or the nearby earthquake of November 1087 (Ibn al-Jauzi, 1181, IX:38; Tables 1 and 3; Fig. 6), sometime after which the structure was rebuilt in 1125 C.E. (5) The 31 March 2005 Mw 6.1 Chālānchulān earthquake, with its maximum damage zone along the Zāgros Main Recent fault, located ~20 km to the SSE of Borujerd (Fig. 6; Tables 1 and 3), caused considerable damage (VII– VI+) to the Borujerd congregational (Fig. 19) and Soltāni mosques, as well as to the mausoleum of Ja’far (Table 3) in the city. Peak ground accelerations of 524 cm/s2 (v), 432 cm/s2 (L), and 357 cm/s2 (T) were recorded at the Chālānchulān station (48.913°N, 33.659°E, +1538 m; available at bhrc.ir, 2006), ~29 km to the SSE of Borujerd, on the Zāgros Main Recent fault. Based on the evidence of the collapsed dome chamber and minaret of the Borujerd congregational mosque III (Fig. 15), the 1316 C.E. earthquake seems to have been much stronger than the 2006 event and had stronger ground motion acceleration parameters and an epicenter much closer to the city of Borujerd than the one in 2006 (Fig. 6).

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084761/spe525-05.pdf by University of Wisconsin Madison user

197

Borujerd Shāhrokh Timurid-Era Congregational Mosque IV (Ca. Post–1405 C.E./Pre–1447 C.E.) After simultaneous collapse of the dome chamber, the minaret, as well as other structural elements of mosque III, the congregational mosque IV, with a much smaller and lower dome chamber, was rebuilt on the foundation and remaining square walls of the previous structure (Fig. 3) during the reign of Shāhrokh Timurid (1405–1447 C.E.). Strangely, no minaret was constructed during this major reconstruction phase after destruction of the mosque III minaret (Fig. 15). Archaeological and architectural data (Mehryār, 1985; Moqaddas, 1997) clearly demonstrate that the original design of the collapsed Saljuq-era congregational mosque III (ca. post– 1090/pre–1139 C.E.) was drastically changed during the reconstruction of the Shāhrokh Timurid-era (1405–1447 C.E.) congregational mosque IV by: (1) avoiding grandeur; (2) reducing the size and height of the dome chamber; (3) retrofitting the surviving structural elements such as the thick load-bearing walls; and (4) eliminating tall free-standing minarets. (1) New brick walls were added in front and behind the surviving old load-bearing square walls (Fig. 3) of the Saljuq congregational mosque III (ca. post–1090/pre–1139), increasing the thickness of the walls to 330 cm. (2) A new, smaller dome chamber with much lower conical dome was constructed (Fig. 3). At least 2 m of the dome chamber height was reduced, and the floor was considerably raised as well. (3) All the surviving architectural elements were re-employed and strengthened in the new structure. (4) Large light/ventilation openings were also avoided, and only four very small openings were constructed for lighting the dome chamber (Fig. 3, above the 16-sided transition zone). (5) More timber bracings were added to the structure. (6) Additional arches were added underneath the surviving entrance arches. The newly constructed congregational mosque IV received smaller and shorter 16-sided and octahedral drum transition zones, with a lower and smaller flat conical dome atop a very short cylinder (Fig. 3). The new elements were tied by timber frames to the body of the surviving square wall with the Kufic inscription of congregational mosque III (Mehryār, 1985; Moqaddas, 1997). There seems to have been a hiatus between the estimated time of complete destruction of the Saljuq-era mosque III (ca. pre–early fourteenth century) and reconstruction of the present dome chamber of the Shāhrokh Timurid-era mosque IV (ca. 1405–1447 C.E.). The failure to immediate reconstruct the destroyed Borujerd congregational mosque III might have been the result of major national socioeconomic-political events of the time, and we do not know if the site was abandoned after a strong earthquake. This could be partly due to the fact that construction of a megastructure such as a congregational mosque in a city requires

198

Berberian et al.

a proper budget and availability of artisans during a prolonged peaceful and prosperous period with national and local interests of the authorities. The 1316 C.E. earthquake took place during the year in which Sultān Oljāitu Ilkhānid died (r. 1304– 1316 C.E.). Throughout the reign of his successor Abu Sa’id Ilkhānid (r. 1316–1335 C.E.), the country lost its cohesion, and the Mongol Ilkhānid dynasty collapsed ca. 1077 C.E. This was followed by the rise of the Muzzaffarids (1335–1393 C.E.) and decline of the state during their wars with the remaining Ilkhānid rulers in different provinces. Furthermore, much of the country was in turmoil, undergoing destruction and massacres during the following Timur invasion period (1370–1405 C.E.). Hence, for more than a century, the country was in a state of anarchy and disorder (Boyle, 1968; Jackson and Lockhart, 1986; Eqbal āshtiāni, 1989; Daryāee, 2012), and no major construction activity was initiated throughout the country. It should be also emphasized that memory of a destructive earthquake can survive more than a century (in this case). For example, the modern-day local oral tradition preserves the account of a twelfth-century destructive earthquake in the Kuhbanān district of southeast Iran (Ambraseys and Melville, 1982). Similarly, the local oral tradition calls for a destructive earthquake in 1238 C.E. destroying the city of Gonābād in northeast Iran (Tabāndeh, 1969; Ambraseys and Melville, 1977). Collapse of the Supplementary Crawl Space Mudbrick Ribbed Vaults and Arches in the Dome Chamber Basement (1316? C.E. or Post–1405/1447 to Pre–1658? C.E.) During an undated major construction phase, four supplementary mudbrick load-bearing ribbed vaults and pillars at four corners with arches supporting the ceiling were added on the –180 cm floor in the basement (Fig. 5) for raising the dome

Figure 20. Collapsed supplementary crawl space and mudbrick ribbed vaults and arches in the basement of the Borujerd congregational mosque dome chamber supported by the added brick work during excavation (see also Figs. 5, 7, and 9–13). The dimensions of the bricks are 20 × 20 × 4 cm.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084761/spe525-05.pdf by University of Wisconsin Madison user

chamber floor, releasing the humidity, and naturally ventilating the basement (Figs. 20 and 21). The springing basal section of the supplementary mudbrick vaults was at elevation –135 cm (45 cm above the –180 cm floor), and the top of the vault was ~120 cm above the –180 cm floor, creating a vaulted crawl space (“gorbeh-rau”). Except for a few limited areas, mudbricks were never found to be used in other phases and parts of the dome chamber (Mehryār, 1985; Moqaddas, 1997). Some unidentified time later, the whole supplementary mudbrick basement structure suddenly collapsed, and the vaults fell down (Figs. 5, 7, 9, 10, 12, 20, and 21). Data on timing of construction and destruction of the mudbrick basement structures are clouded and were not bracketed during the excavation by the archaeologists. However, it is clear that the simultaneous collapsed mudbrick structures show the following evidence of strong ground motion due to an earthquake of unknown date: (1) The crown sections of all the basement supplementary mudbrick load-bearing ribbed vaults and ceiling arches holding the ceiling of the basement collapsed simultaneously (Figs. 20 and 21). (2) The springing basal sections of the mudbrick ribbed vaults were detached from their supporting abutment pillars (piers) and slid down to rest on collapsed debris on the basement floor. (3) Mudbricks from the broken arches and ribbed vaults were found scattered amongst collapsed structural elements in the basement (Figs. 5, 7, and 8–13). (4) The crawl space shell collapsed. (5) Numerous eighth- to twelfth-century potsherds, the Saljuq-era glazed turquoise decorative/inscribed tiles (used above the Saljuq Kufic inscription band and around

Figure 21. Another view of the collapsed supplementary crawl space and mudbrick ribbed vaults and arches in the basement of the Borujerd congregational mosque dome chamber supported by the added brick work during excavation (see also Figs. 5, 7, 9–13, and 19). The dimensions of the bricks are 20 × 20 × 4 cm.

Evidence of historical seismic activity along the Zāgros Main Recent fault at Borujerd the top of the mehrāb inside the structure, as well as outside the mosque), bricks, charcoal, and bones of birds and domestic animals were discovered in a highly disturbed context below the imported debris at elevation –65 cm of the basement (Fig. 5; Mehryār, 1985; Moqaddas, 1997). Some unknown time after the simultaneous collapse of the basement structures (–180 cm to –65 cm below the present floor; Figs. 5, 7, and 13), the basement area was filled with imported debris (from –65 to –20 cm) brought from other collapsed parts of the mosque, and then the –20 cm floor was constructed upon the debris (Figs. 5 and 7). Moqaddas (1997) reported that 10 silver coins of the Shāhrokh Timurid period (1405–1447 C.E.) were discovered from the disturbed imported debris, and he bracketed the time of filling the basement with the imported debris to be “around the early ninth to fifteenth centuries.” This long time period covers the period of destruction caused by the 1316 C.E. earthquake and reconstruction by Shāhrokh Timurid in the early fifteenth century discussed already. It is not clear if the debris containing the collapsed pieces of Saljuq-era decorative turquoise tilework found in the imported fill material in the –65 cm to –20 cm level (Fig. 5) was brought to the basement immediately after the 1316 C.E. earthquake and destruction of the dome chamber and collapse of the basement structures or much later. Earlier, Mehryār (1985) reported that near the base of the top imported mixed debris material, glazed and unglazed potsherds, Saljuq-era glazed turquoise decorative/inscribed tiles, and a yellow coin dated 1124 H/1615 C.E. at elevation –60 cm were discovered (Fig. 5). Unfortunately, the imported debris material is highly disturbed, the data are clouded, the reports are not conclusive, and no profile or photograph is available from the excavation activities. We report this controversy hoping it will be resolved in the future excavations. Excavation of trench 2, near the eastern wall of the Western Portico (Shabestān5; Fig. 4), below the –30 cm lime mortar layer (the portico floor was constructed ~30 cm above the dome chamber floor), revealed four silver coins of the Shāhrokh Timurid and Sultān Mohammad Bahādor (851H/1447 CE and 853H/1449 CE) period, indicating that the supplementary Western Portico wing was added sometime after 1449 C.E. (Mehryār, 1985). Usually, debris from a collapsed mosque is considered sacred and is reused during renovation phases or poured into a running river. In this case, the debris from the collapsed sections of the mosque was used to fill the basement. As mentioned earlier, on top of the imported debris, the new –20 cm floor was constructed (Fig. 5). Since the nine-step wooden pulpit (menbar) with inscription date of 1069 H/1658–1659 C.E. rests on this floor (–20 cm below the present floor), the construction date of the –20 cm floor can be broadly dated as 1658 C.E. (Fig. 5). All evidence indicates sudden destruction of the basement structures underneath the dome chamber by an earthquake. The

5

Shabestān (shabistan [sic]): a large nave or hall with many visible piers or columns inside.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084761/spe525-05.pdf by University of Wisconsin Madison user

199

exact date of the sudden destruction of the basement structural elements (below –65 cm; Figs. 5, 7, 8–13, 20, and 21) and accumulation of additional debris (–65 to –20 cm) underneath the dome chamber has not been properly documented by the original excavators. During the 200 yr interval between the construction of the Borujerd Shāhrokh Timurid-era mosque IV (post–1405/ pre–1447 C.E.) and raising the floor to –20 cm below the present floor (ca. 1658 C.E.), two earthquakes were recorded in the region in 1430 and 1495 C.E.6 (Fig. 6; Table 3), and the mosque was repaired in 1613 and 1619 C.E. (Meshkāti, 1970). Unfortunately, due to incomplete and undocumented archaeological investigation, haphazard excavations, and the disturbed state of the material in the basement, the major destruction/renovation phases are clouded, and it is very difficult to determine the exact timing of different destruction/renovation phases and the causes with the existing limited data in this complex structure. The case is discussed and documented here to help further excavations. Major Renovations since the Safavids (1502 C.E.–Present) The Borujerd congregational mosque IV (post–1405/pre– 1447 C.E.) then underwent several additional major phases of rebuilding, repairs, and renovations (Tables 2 and 3) during the Safavid (1502–1722 C.E.), Qājār (1799–1925 C.E.), Pahlavi (1925–1979 C.E.), and recent eras (after the mosque was damaged by the 1987 Iraqi air raids during the Iran-Iraq war). The Safavid dome chamber floor (ca. 1658 C.E.) was constructed at –20 cm (Fig. 5), and during the Qajar era, the present floor (0.0 in Fig. 5) was built (Meshkāti, 1970; Mehryār, 1985; Moqaddas, 1997). The present two minarets with the eastern and western sanctuary ayvāns were constructed in 1794–1795 during a major Qājār-era renovation phase (Figs. 3, 4, and 19). The two dedicated inscriptions clearly mention that the minarets were added after repair works conducted inside the mosque in 1794–1795. The minarets were severely damaged during the 31 March 2006 Mw 6.1 Chālānchulān earthquake, with epicenter located ~25 km to the southeast of Borujerd (Figs. 6 and 19). The northern sanctuary (Fig. 4) was constructed in 1798 (Mehryār, 1985; Moqaddas, 1997).

6 The 1430 C.E. earthquake was reported in Hamédān (103 km NNW of Borujerd), Wāsit (Waseth; in Mesopotamia), and Audar (location unknown; it is not clear if al-‘Umari referred to the modern Auda āl Wahām near the Euphrates River in Mesopotamia, 31°5′N, 46°1′E); “some places were swallowed up and buildings were destroyed and many people perished” (al-‘Umari, 1793, fol:160r; Sani’ al-Dauleh, 1880–2, II:57). It is not clear if al-‘Umari is referring to different earthquakes that happened in the same year (Figs. 2 and 6; Table 3). Wāsit (Wāseth: “Middle Town” in Arabic) was located in between the three cities of Kufa, Basra, and Ahvāz on the west bank of the Tigris River (32°14′N, 46°18′E), about 150 km SSE of Baghdād and 350 km SW of Hamédān. The 1495 C.E. earthquake was reported in Hamédān (103 km NNW of Borujerd), Esfahān (377 km SE of Hamédān and 300 km SE of Borujerd), and Ray (288 km NE of Hamédān), and a mountain in Hamédān was fissured into pieces (al-‘Umari, 1793, fol:174v; Ambraseys, 1979; Ambraseys and Melville, 1982). It is not clear if al-‘Umari referred to a single earthquake or more. Nonetheless, it was associated with landslides and rock avalanches at Hamédān, north of Borujerd (Figs. 2 and 6; Table 3).

200

Berberian et al.

Post–1447 C.E. Damage of the Dome Chamber Evidence of additional post–1447 C.E. damage in the dome chamber with undocumented dates has been recorded: (1) A deep fracture with 25 cm opening was discovered behind the northwestern conical squinch (corner squinch; “filpush” in Persian: a section of vaulted masonry bridging an angle of a rectangular room, often used to create a circular base for a dome). (2) Some bricks of the octagonal drum transition zone, the eight corners of the ribbed vaults above the conical squinches, and parts of the base of the 16-sided drum transition zone (Figs. 3 and 4) either have fallen down or become separated from the rest of the structure, showing an ~30 cm gap. (3) Bricks of the piers of the conical squinches have been buckled, displaced, rotated, and are hanging loosely (see fig. 20 in Moqaddas, 1997). (4) The northeastern ribbed vault above the squinch has collapsed. During this period, four distant earthquakes (1495, 1665, 1853, and 1876 C.E.) and one adjacent earthquake (1909 C.E., Mw 7.4) have taken place (Table 3; Fig. 6). It seems most likely that the mentioned damages were caused by the 1909 C.E. destructive earthquake, which fractured the Borujerd congregational mosque (causing long deep fractures in the main vault of the southern ayvān), damaged other buildings, threw objects from shelves, and caused waves on ponds in the city of Borujerd. POSSIBLE INDIGENOUS PALEO-ARCHITECTURAL INNOVATIVE ATTEMPTS TO INCREASE COHERENCE AND ELASTICITY OF THE RIGID BRICK STRUCTURE TO WITHSTAND EARTHQUAKE SHEAR STRESS The late Renaissance Italian architect Pirro Ligorio (1513– 1583 C.E.), who witnessed the 17 November 1570 Ferrara (the Po Valley, Italy) Ml ~5.5 earthquake and studied the destroyed structures, designed the first recorded earthquake-proof house in Europe, described in his treatise Libro de Deversi terremoti (Book of Several Earthquakes). In the last section of his treatise, “Rimedi Contra Terremoti per la Sicurezza Degli Edifici” (“Remedies against Earthquakes for Building Security”), architect Pirro Ligorio presented his design plans for an earthquake-proof building to withstand strong ground motion of earthquakes. He utilized the correct dimensions for the main load-bearing walls, better and stronger bricks, elastic structural joints, and iron rods. He blamed extensive earthquake damages on the inappropriate techniques and weak building material used in the contemporary buildings (Guidoboni, 1997, 2015; Coffin, 2004). Unlike the documented historic European cases, in the absence of any written historical documents (of architects, architectural plans, treatises dealing with construction details, etc.) during several renovation and rebuilding phases of the Borujerd con-

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084761/spe525-05.pdf by University of Wisconsin Madison user

gregational mosque, a retrospective interpretation of architectural elements may suggest that additional steps were probably taken to increase the stability of the structure throughout its history. Although no planning of the monuments and megastructures has survived, these structures should have been constructed from prepared and calculated plans. The oldest extant treatise dealing with construction and architectural geometry, calculations, plans, and profiles was written by the Iranian polymath, Ghyāth al-Din Kāshāni (Kāshi; 1380–1492 C.E.). In the twelfth chapter of his book Kitāb al-Ahyā wa al-āthār (Book of Animals and Monuments) and chapter nine of Miftāh al-Hisāb (The Key to Arithmetic, 1427 C.E.), Kāshāni applied mathematics to practical architecture and described the architectural and mathematic rules to be followed during construction of arches and buildings (Quatremere, 1836; Wilber, 1955; Afshar, 1979a, 1979b; Jazabi, 1985). In his book Jame’ al-Tavārikh (Compendium of Chronicles), covering the history of the Mongol dynasty in Iran (Hamédāni, 1304; Melville, 2008), the Iranian scholar Rashid al-Din Fazlollāh Tabib Hamédāni (1247–1318 C.E.) wrote that Ghāzān Khān Ilkhānid (1295–1304 C.E.) drew the plans of the High Tomb of Ghāzān Khān. Although the ruler himself did not draw the architectural plans and calculate the structural elements, the phrase shows that architectural plans were prepared prior to construction, and the ruler viewed and accepted the plans (Jahn, 1940; Wilber, 1955; Lewcock, 1984). The extant table of contents of Rashid al-Din Fazlollāh Tabib Hamédāni’s lost book in 24 volumes (al-Ahyā wa al-āthār [Living Things and Monuments]; late thirteenth to early fourteenth century C.E.) includes a chapter on rules to be followed in building houses, religious buildings, and fortresses, and information on the construction of tombs (Lewcock, 1984). After collapse of the dome chamber, minaret, upper parts of the load-bearing walls, and ribbed vaults of the Borujerd Saljuqera congregational mosque III (ca. post–1090/pre–1139 C.E.), most probably by the 1316 C.E. earthquake, in order to enhance the coherency and elasticity of the new mosque (Borujerd congregational mosque IV; ca. post–1405/pre–1447 C.E.) and minimize the future earthquake damage to the structure, major architectural modifications were implemented in the design and construction of the new Shāhrokh Timurid-era mosque IV. The fact is that these indigenous paleo-architectural innovative attempts were never implemented during the earlier or later reconstruction phases (Table 3). Apparently, large safety factors in design and in uniformity with the seismic nature of the region were implemented to resist earthquake forces for a better structural functioning of the building. The use of new proportions in mosque IV indicates that the architects had profound understanding and knowledge of the load, thrust, and shear forces and stresses acting on the structure, structural strength, stiffness, and stability, as well as structural failure due to the strong ground motion caused by strong earthquakes in an earthquake-prone country like Iran. Presumably, the indigenous earthquake engineering steps were taken after studying the structural failure of the previous majestic structure

Evidence of historical seismic activity along the Zāgros Main Recent fault at Borujerd of mosque III (most possibly by the 1316 C.E. earthquake), and gaining knowledge of the fundamentals of how a large brick structure like the Borujerd congregational mosque reacted to earthquake shear stress. Retrofitting the Surviving Load-Bearing Structural Elements Two brick wall layers ~30 cm thick were added to the front and back of the partially surviving original structural square walls of the Borujerd congregational mosque III (Figs. 3 and 4), increasing the total thickness of the load-bearing walls to ~330 cm. The massive fortified load-bearing square walls were then tied together with timbers (below the octahedral as well as the 16-sided transitional substructure) for better support of the new conical dome built upon them, as well as to dissipate the elastic seismic waves. Underneath the major surviving vaults, especially at the entrances, additional “support vaults” or “archivolts” were added, changing them to stronger double archivolts. Avoiding Grandeur and Majesty and Implementing Simplicity by Reducing the Size and Height of the Dome Chamber The dome chamber of the Borujerd congregational mosque IV (ca. post–1405/pre–1447 C.E.) was designed to be much smaller, shorter, and lower in height than the earlier collapsed lofty dome chamber of congregational mosque III (ca. post– 1090/pre–1139 C.E.), and it was built on thicker fortified structural square walls with two additional brick layers (described in previous section). Unlike other Iranian grand and majestic mosque domes, the Borujerd congregational mosque III lofty dome, and the presentday outer conical dome shell, a saucer-like low conical roof (saucer dome, segmental, cloister, calottes) was constructed for the congregational mosque IV (Fig. 3). It had a very low height/span ratio, with no dome curvature, built on a short cylindrical drum. The saucer domes, with a profile of less than half a dome, reduce the dome tension and are stronger than hemispherical domes, despite the fact that they have increased radial thrust. The saucer domes are much thinner than other dome shapes and less unstable (Dodge, 1984; Gye, 1988; Fleming et al., 1991; Pirniā, 1991). By elevating the floor of mosque IV about 2 m, and shortening the height of the dome chamber (estimated to be at least about 2 m), the total height of the structure (from the base of the saucer dome) was reduced at least by about 4 m (Fig. 3). Avoiding Free-Standing Structures The congregational mosque IV did not have free-standing minarets and the present surrounding adjunctions. This might be due to the fact that after collapse of the former tall, slender, solitary free-standing masonry minaret (an inverted pendulum) by

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084761/spe525-05.pdf by University of Wisconsin Madison user

201

the earthquake (Fig. 15), the designers and architects decided to eliminate any free-standing minarets in the earthquake-prone city (this was a peculiar decision taken for a congregational mosque in the history of the country). By eliminating minarets, they avoided additional damage to the structure in case of toppling of free-standing structures on the dome chamber. During the 31 March 2006 Mw 6.1 earthquake along the Zāgros Main Recent fault, with meizoseismal area located ~20 km south of Borujerd (Fig. 6), the 16.75 m minarets of the mosque were badly damaged with development of long shear fractures (Fig. 19), opening brick joints, and collapsing of the top section of the eastern minaret (minaret turnet, “goldasteh”). The present northern sanctuary ayvān, two minarets, and northern portico were built in 1794 as an adjunct to the congregational mosque IV dome chamber (Fig. 3; Table 3). Minimizing the Size and Reducing the Number of the Light/Ventilation Openings No light/ventilation opening was provided in the top or the middle section of the saucer-like conical roof. Only at four points, corresponding to the midpoints of the four sides of the square chamber, is the short cylinder drum above the 16-sided drum transition zone pierced near the base by small rectangular light/ventilation openings with arched tops (Figs. 3 and 4). Utilizing Timber Bracing to Neutralize Strong Ground Motion of Earthquakes Despite the scarcity of strong wood on the Iranian Plateau, wood poles and beams were used as bonding and stiffening structural elements (Wilber, 1955; Wulff, 1966). Wood elements in a structural fashion and timber bracing were utilized to interlock the structural elements of the Borujerd congregational mosque IV, as another major attempt to construct a resilient brick structure. Large expensive timbers were probably transported from a considerable distance to the construction site. The following major timber reinforcement was utilized in different structural elements of the Borujerd congregational mosque IV: (1) Large wooden beams were embedded horizontally along and at right angles to the direction of the structural walls at different intervals. About 11 rows of long timbers along and across the structural elements, from the floor to the dome drum, were utilized throughout the structure, in the bearing walls, above the head of the mehrāb, below the wall arches of the transition zones, and underneath the octahedral and the 16-sided drum transitional zones, forming a chain to bind together the base of dome and vaults, the dome, and lintels (see figs. 12, 13, 15, 17–19 in Moqaddas, 1997). (2) The arched vaults were braced by long horizontal timbers embedded below the springer line of the arches, above the impost block (abacus), to neutralize the arch thrust and additional shear stress of earthquakes.

202

Berberian et al.

The intention of utilizing a broad network of large timber elements in a structural fashion embedded through the brick loadbearing structures was: (1) to interweave the structural fabric of the brick structure closer together; (2) to enhance internal cohesion with the base of the dome; (3) to increase the coherency and elasticity of the rigid brick structure; (4) to prevent uneven differential settlement of the structure; (5) to resist and dampen propagation of shear fractures caused by strong ground motion; (6) to minimize the damage from the strong ground motion of future earthquakes by acting as a resilient bracing; and (7) to dissipate and transmit the imposed shear stress of seismic waves evenly to the whole structure, shaking in harmony with the ground. Similar practices using timber-laced walls, vaults, and dome were later carried out during the design and construction of the Borujerd Soltāni mosque (1794 C.E.), but to a lesser extent, in this earthquake-prone city. Although timbers utilized in the brick structure of the Borujerd congregational mosque IV created a resilient structure by binding the structural elements and distributing load and stress, their gradual deterioration due to the lack of a proper passive ventilation system and the humidity, condensation during the cold season, and termite attacks left long and large hollow tunnels in the structure, resulting in weakening of the brick building through the centuries and leading to shear fracture development and damage of the structure by a medium-magnitude earthquake (see figs. 18 and 19 in Moqaddas, 1997). The aforementioned indigenous old attempts to strengthen the structure resulted in a more compact form to the structure and helped its survival for more than 600 yr with few evidences of development of deep vertical fractures. Despite all these efforts, the 31 March 2006 Mw 6.1 Chālānchulān earthquake, which took place 20 km to the south of Borujerd (Fig. 6), severely damaged both the congregational and Soltāni mosques. The damage was due to: (1) loose joints; (2) weak connections among different structural elements; (3) poor consideration of seismic projections in reconstruction works (Vosoughifar, 2007; Vosoughifar and Davari, 2009); (4) decomposition of the vital large timber elements; and (5) lack of knowledge of the nearby active fault and its maximum credible earthquake. PROBABLE CAUSATIVE SEISMIC FAULT From a review of multidisciplinary data, a reasonable local interpretation can be outlined by putting archaeoseismologic and paleo-architectural indicators within the seismotectonic framework of the region with the ultimate goal of integrating archaeoseismic data within the network of seismic risk analysis. The Zāgros Main Recent fault (Tchalenko and Braud, 1974; Berberian, 1994, 2014; Berberian and Yeats, 2001; Talebian and Jackson, 2002; Alipoor et al., 2012) cuts through the Silākhor Valley to the south and west of the city of Borujerd (located 4 km to the northeast of the fault; Figs. 6 and 22). The fault is the closest major seismogenic fault (with recorded Mw 7.4 1909 C.E. earthquake just south of the city), and unlike the Zāgros fold-and-

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084761/spe525-05.pdf by University of Wisconsin Madison user

thrust belt to the southwest (Berberian, 1995), there is no blind or décollement fault in the vicinity of the city of Borujerd. The Zāgros Main Recent fault is a major NW-SE–trending right-lateral strike-slip seismic fault zone more than 800 km long located between the Zāgros fold-and-thrust belt in the SW and the Central Iranian range and basin in the NE (Fig. 1). The fault more or less follows the trend of the Zāgros Mountains and the Main Zāgros reverse fault (the Neo-Tethyan geosuture). The Zāgros Main Recent fault has been the source of frequent significant earthquakes of up to Mw 7.4 during the twentieth century and should have been such during the earlier historical record (Tables 1 and 2; Figs. 6 and 22). Focal mechanism solutions along the Zāgros Main Recent fault show pure rightlateral strike-slip movement on a high-angle fault plane dipping NE (Fig. 22; Table 1). As discussed earlier, the 1316 C.E. earthquake could have occurred along the northwestern end of the Dorud segment, like the 2006 event, or the southeastern part of the Nahāvand segment of the Zāgros Main Recent fault (Fig. 22). Since no major recent or historical earthquakes are recorded on the southeastern part of the Nahāvand segment of the Zāgros Main Recent fault, and considering the return period of a 1909-type earthquake along the Dorud fault segment, it is most probable that the 1316 C.E. event took place along the Nahāvand segment of the fault (we will return to this issue later). The severity of the 1316 event may be compared with that of the 23 January 1909 Mw 7.4 Silākhor earthquake, which took place along the Dorud segment of the Zāgros Main Recent fault (Berberian, 1995, 2014; Berberian and Yeats, 2001; Figs. 6 and 22; Table 3). The city of Borujerd is located about 4 km to the east of the Zāgros Main Recent fault, near a gap with right step-over along the Dorud and the Nahāvand segments in the northwestern part of the Silākhor Valley (Figs. 6 and 22). It is also located between: (1) the meizoseismal areas of the 23 January 1909 (Mw 7.4, intensity [I] IX+ MMI) Silākhor, 28 October 1961 (Mb 5.0), and 31 March 2006 (Mw 6.1; I VIII–) Chālānchulān earthquakes to the southeast; and (2) the 14 August 1958 (Ms 5.7), 16 August 1958 (Mw 6.5, I VIII+), and 14 October 1961 (I VII) earthquakes to the northwest (Figs. 6 and 22; Table 3). The 1909 C.E. earthquake (Fig. 6) was strong enough to fracture the Borujerd congregational mosque (as indicated by the repaired long deep fractures in the main vault of the southern ayvān), damage other buildings, throw objects from shelves, and cause waves on ponds and pools in the city of Borujerd. The earthquake survivors from the Silākhor villages flooded the city seeking help from the officials. The 14 October 1961 Mogh (VII MMI) and the 28 October 1961 (M 5.0) Haydarābād earthquakes caused slight damage (V+ MMI) at Borujerd (Berberian, 1995; Berberian and Yeats, 2001; Figs. 6 and 22; Tables 1 and 3). During the latter event, one of the minaret turnets (“goldasteh”) of the Borujerd congregational mosque collapsed, the dome of the mosque cracked, and ~40 houses were fissured in Borujerd. During the 3 May 2005 (Mw 4.9) South Borujerd earthquake (15 km to the south of Borujerd), the city was strongly shaken and

Evidence of historical seismic activity along the Zāgros Main Recent fault at Borujerd

203

Figure 22. Meizoseismal area of medium- to large-magnitude earthquakes (top) and space-time diagram (bottom) of recorded seismicity along the central section of the Zāgros Main Recent right-lateral strike-slip fault. Where the date of the earthquake is shown, it is given by year.month. day. Thicker lines show faults with documented surface ruptures. Solid lines in the bottom space-time diagram indicate an earthquake known to have ruptured the ground surface; dashed lines indicate highly probable surface-rupturing events. Distances are along strike. Triangle— archeological mound. Circle—city/town. Focal mechanism solutions: 13 December 1957 (McKenzie, 1972); 14 and 16 August 1958 (Shirokova, 1967); 24 March 1963 (Ni and Barazangi, 1986); 29 May 1987 (Harvard, 2016); 24 April 2002 (Harvard, 2016); 31 March 2006 (Peyret et al., 2008). Figure is modified from Berberian (2014). Inset top right: Map of Iran showing boundary with Arabian plate (line with teeth). AZ— Āzarbāijān; KP—Kopeh Dāgh; M—Makrān; S—Sistān suture zone; TP—Turān plate.

slightly damaged (IV). About 10 suburb villages were damaged (Figs. 6 and 22), one person was killed, 26 people were injured, and 200 villagers of the Silākhor plain demanded tents to live in (irna.ir; Tables 1 and 3). Low-magnitude earthquakes frequently occurring along the Zāgros Main Recent fault are punctuated at long intervals by large-magnitude earthquakes (Figs. 6 and 22). A glance at the recorded seismic history of the Zāgros Main Recent fault (Fig. 22) shows that the Nahāvand segment of the fault (from the city Nahāvand in the NW to Borujerd in the SE) formed a seismic gap in between the temporal clusters of large earthquakes at both its ends. Apparently, the 1316 C.E. strong earthquake filled the gap, and since then, no major earthquake has taken place along this segment (Fig. 22).

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084761/spe525-05.pdf by University of Wisconsin Madison user

Geological investigations showed that the Zāgros Main Recent fault cuts the early Neogene active folded sediments of the Lower Bakhtiāri Formation (Authemayou et al., 2006). A rightlateral offset of a geological marker bed along the Nahāvand and Dorud segments of the Zāgros Main Recent fault was reported by Gidon et al. (1974). Talebian and Jackson (2002) reported a right-lateral offset of 50–70 km and estimated a horizontal slip rate of 10–17 mm/yr along the fault. Vernant et al. (2004), using GPS data, arrived at slip rate of 3 ± 2 mm/yr. Based on the offset of river valleys, Bachmanov et al. (2004) derived a 10 mm/yr slip rate for the Zāgros Main Recent fault around longitude 49°E. Walpersdorf et al. (2006) suggested a cumulative horizontal slip rate of 4–6 mm/yr, accompanied by 3–6 mm/yr perpendicular

204

Berberian et al.

shortening. Authemayou et al. (2009) obtained a minimum slip rate of 3.5–12.5 mm/yr for the southeastern section of the Zāgros Main Recent fault. Later, based on his three-dimensional mechanical modeling using GPS data, Nankeli (2011) arrived at a 2.3 mm/yr horizontal slip rate for the fault. Finally, Alipoor et al. (2012) suggested a cumulative horizontal slip rate of 1.6– 3.2 mm/yr based on 16 km of displacement along the fault. Since 1900, the record shows that earthquakes on the Zāgros Main Recent fault have occurred with greater frequency than on other faults in the Iranian Plateau (Figs. 6 and 22; Table 1), but the pre-1900 earthquake history of the fault is unknown, and a critical and comprehensive account of the historical earthquakes along the fault is still lacking. Paleoseismologic trench studies across the Zāgros Main Recent fault, as well as archaeoseismologic and paleo-architectural studies of the historical monuments along the fault will help understanding of the seismic behavior of the fault. DISCUSSION Evidence of strong ground motion created by large-magnitude earthquakes has already been documented at a small number of archaeological sites on the Iranian Plateau (Berberian and Yeats, 2001). Moreover, there is evidence for site abandonment (Wilkinson, 1986) and shifts in settlement location after some large-magnitude earthquakes, as well as postseismic structural innovations enabling the construction of structures to withstand strong ground motion (Berberian et al., 2012, 2014). It should be reiterated that neither archaeoseismologic indicators at archaeological sites nor the paleoseismological data of the active faults in Iran have been studied to the extent that they have in many other regions (McGuire et al., 2000; Meghraoui et al., 2003; Galadini et al., 2006; Niemi, 2008; Pérez-Lopez et al., 2009; Reicherer et al., 2009; Silva et al., 2011). We therefore have to rely on the limited exposures of archaeological soundings, excavation reports, and photographic documentation taken during excavations. Although the limited and preliminary original investigations referred to in this study were not targeted toward archaeoseismicity, numerous strong ground motion indicators have been detected by scrutinizing the original archaeological reports and historic photographs taken during the excavations. The history of the Borujerd congregational mosque has shown at least seven major damage/destruction phases followed by essential rebuilding and restoration phases, with some uncertainty concerning the exact dates and causes. The cause of destruction is not mentioned or documented in the historic annals that we meticulously scrutinized (Tables 2 and 3). It should be mentioned that the lack of seismic records in the study area in the vicinity of the active Zāgros Main Recent fault since the Sāssānids (224–642 C.E.) is not indicative of lack of seismicity, but simply of nondocumentation of earthquakes (Table 1). The lack of any surviving local history of the city reflects the absence of a continuous scholarly tradition and travelers’ accounts in towns off the main lines of communications (such as

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084761/spe525-05.pdf by University of Wisconsin Madison user

the Khorāsān and Silk Roads) prior to the seventeenth century, with little cultural interest and numerous devastating invasions (Table 3). Archaeoseismologic data defining the cause of the historic grand mosque collapse reported and analyzed here clearly show that the study area was not quiescent but on the contrary suffered large-magnitude earthquakes. It seems that the 1316 C.E. earthquake, which presumably severely damaged the city, contributed to the process of decline and decay caused by over a century of instability, hardship, and destructive sacks by Timur (Tamerlane) hordes, including the demise of the Ilkhānids (1218–1334 C.E.) and internal struggle of the Mozaffarids (1314–1393 C.E.) prior to Timur’s destruction (1385, 1387, 1393, 1400, 1401, 1402 C.E.; Boyle, 1968; Jackson and Lockhart, 1986). Apparently, Borujerd was in a dilapidated and neglected state for a century until the prosperous reign of Amir Shāhrokh Mirzā Timurid (1405–1447 C.E.), when the existing monuments were reconstructed in the region as well as the country at large (Pope, 1938, 1965; Hutt and Harrow, 1977, 1978; Hutt, 1984; Jackson and Lockhart, 1986; O’Kane, 1987). The destruction of the Borujerd monuments and the city is indicative of the close proximity of the city to the epicenter, and the large-magnitude nature, of the 1316 C.E. earthquake. The earthquake must have caused substantial damage and destruction to almost all the buildings in the city and the surroundings, since no pre–fourteenth-century C.E. historical monuments exists in the Silākhor Plain as well as at Nahāvand and other nearby towns and cities. As with the 1909 Mw 7.4 event, the 1316 C.E. earthquake along the Zāgros Main Recent fault should have been associated with surface faulting as well as various ground deformations, and disruption of the surface and underground water supplies in the Silākhor Valley. Estimates of the seismic parameters of the 1316 C.E. earthquake and of the casualties cannot be made at this stage of study. Estakhri (951, p. 165), Ibn Hawqal (978, p. 170), and Zakaryā Qazvini (1275, p. 95) mentioned the city measuring over half a league across. The population of the city during the reign of Nāser al-Din Shāh Qājār (r. 1848–1896 C.E.) was estimated between 20,000 and 50,000 (Dabir Siyāqi, 1962). De Morgan (1894–1905), who was in Borujerd in 1890, mentioned a population of ~20,000 in the city. The population of the city was reported as ~22,000 in the nineteenth century (Ehlers, 1989). The national Iranian census carried out by the Statistical Center of Iran (SCI, 1956–2012) showed the population of Borujerd was 49,186 (in 1956), 71,486 (1996), 101,345 (1976), 183,879 (1986), 201,016 (1991), 217,804 (1996), 229,541 (2006), and 240,654 (2011). The data indicate a 10-fold increase of the Borujerd population during the last century. Bearing in mind that the city is located in the vicinity of the active Zāgros Main Recent fault (Figs. 6 and 22), an increase in destruction and casualties by a future largemagnitude earthquake is expected. From scaling relationships (Wells and Coppersmith, 1994; Scholz, 1982), the Dorud fault segment of the Zāgros Main Recent fault (the source for the 1909 Mw 7.4 earthquake to the southeast of Borujerd; see Fig. 6), with a length >100 km, has

Evidence of historical seismic activity along the Zāgros Main Recent fault at Borujerd the potential to produce earthquakes with magnitudes of >~7.5. Assuming an average right-lateral displacement of ~4.5 m associated with the 1909 event, the cumulative right-lateral slip rate of ~2.4 mm/yr (discussed earlier) will result in an ~2000 yr return period for such a large-magnitude earthquake along the Dorud fault segment of the Zāgros Main Recent fault. Hence, it is more likely that the Nahāvand fault segment was reactivated during the 1316 C.E. earthquake (Figs. 6 and 22). The facts would be revealed by paleoseismic trenching across the Nahāvand and Dorud segments of the Zāgros Main Recent fault. The Nahāvand fault segment of the Zāgros Main Recent fault (the source fault of the 14 August 1958 Ms 5.7 Givaki, 14 August 1958 Ms 5.5 Kalādeh, and 16 August 1958 Mw 6.5 Firuzābād earthquakes, located to the NW of Borujerd; Berberian, 2014; Tables 1 and 3) has a total length of ~100 km (Figs. 6 and 22). The 274–578 C.E. Kangāvar (Kāmbakhsh-Fard, 1994; Berberian, 1994, 2014; Āzarnoush, 2009; Berberian and Yeats, 2001) and 1650–1600 B.C.E. Gowdin-Giyān (Young, 1968, 1969; Young and Levine, 1974; Berberian, 1994, 2014; Berberian and Yeats, 2001; Berberian et al., 2014) earthquakes might have originated on the Nahāvand or Sahneh segments of the Zāgros Main Recent fault (Fig. 22). Regression of fault length, magnitude, and displacement (Wells and Coppersmith, 1994) suggests that the ~100-km-long Nahāvand fault segment of the Zāgros Main Recent fault is capable of generating a maximum credible earthquake of magnitude ~Mw 7.4, with maximum and average right-lateral horizontal displacements of ~3.9 m and ~2.2 m, respectively, if the whole 100 km fault length ruptures during a single event. We think that the 1316 C.E. earthquake apparently took place by reactivation of this fault segment and had a magnitude ~≥7.0, destroying all the monuments and structures in the Borujerd-Nahāvand area (Figs. 6 and 22). The town of Nahāvand (34°11′N, 48°22′E, +1670 m) along the Nahāvand segment of the fault, which is located ~48 km to the NW of Borujerd (Fig. 6), was the decisive site where the Battle of Nahāvand (642 C.E.) almost completed the invasion of the Moslem Arabs and ended the Sassanid Dynasty in Iran (al-Balādhuri, 890; Ibn al-Faqih Hamédāni, 903; Frye, 1975). No intact monument has remained in or around the town of Nahāvand. Limited soundings at the Dau-Khāharān (lit. “Two Sisters”) mound in the city of Nahāvand exposed disturbed contents of the Seleucid finds (312–174 B.C.E.), such as Greek inscriptions (193 B.C.E.), the stone altar, bronze statuettes of the Greek gods, column bases and capitals, and potsherds. Apparently, the findings indicate the most likely site of the Greek temple Laodike at Laodikeia (Nahāvand) built by the Seleucid King Antiochus III (223– 187 B.C.E.) for his wife Queen Laodicea (Rahbar and Alibaigi, 2009, 2011; Rahbar et al., 2014). The cause of destruction of the Nahāvand temple located ~4 km to the north of the Zāgros Main Recent fault is not known. The temple’s columns have been illegally excavated over the years and are currently being used as decorations in Naāvand’s Hājiān Bāzār and several other parts of the city. In some soundings, the excavations continued to a depth of 450 cm from the

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084761/spe525-05.pdf by University of Wisconsin Madison user

205

surface, but all layers were apparently damaged and disturbed because of the Qājār period and recent construction activities and illegal diggings. Although above the Seleucid layers, potsherds of the Parthian (250 B.C.E.–224 C.E.), Sassanid (224–642 C.E.; possible remnants of Yazdgerd Castle; Cais, 2005), Ilkhānid (1218–1334 C.E.), Zand (1750–1779 C.E.), and Qājār (1779– 1925 C.E.) layers have been discovered, no clear archaeoseismic indicators were looked for in the narrow trenches, and no excavation profiles are available (Mehdi Rahbar, 14 November 2012, personal commun. with Berberian,; Sajād ‘Alibaigi, 25 November 2012, personal commun. with Berberian). Tabari (915) wrote that during the 642 C.E. Battle of Nahāvand with the Moslem Arab invaders, Nahāvand had a fire temple. There has been no trace of the fire temple, which was possibly replaced by a mosque after the Muslim Arab invasion. Later, Estakhri (951), Ibn Hawqal (978), Anonymous (982), and Moqaddasi (985) added that Nahāvand had two old and new congregational mosques. The two tenth-century congregational mosques are no longer extant, and we do not know the cause of their destruction. We may speculate that the mosques were possibly damaged by the 1316 C.E. or other earthquakes, but there is no supporting evidence. The present congregational mosque of Nahāvand has a mehrāb stone from an older Saljuq (1000– 1218 C.E.) mosque. It is probable that the Saljuq-period mosque was destroyed by the 1316 C.E. earthquake, and the mehrāb stone was used in construction of the new mosque; however, we have no evidence for this scenario as well. Destruction of the Borujerd congregational mosque III, and the absence of any pre–1316 C.E. monuments in the cities of Borujerd and Nahāvand (along a distance of 48 km) may roughly delineate the maximum damage area of the 1316 C.E. earthquake (Figs. 6 and 22). As discussed earlier, it is probable that the whole 100 km fault length of the Nahāvand segment of the Zāgros Main Recent fault was reactivated during this event. Since the cities of Hamédān (34°47′N, 48°30′E, +1819 m) and Golpāyégān (33°27′N, 50°17′E, +1817 m) were also mentioned in the accounts of the contemporary historian Kāshāni (Kāshāni, ca. 1320, p. 179), and because of recent earthquake reports, we briefly review the status of the extant historical monuments of these cities (Figs. 2, 6, and 16) to see if there is any evidence of the 1316 C.E. earthquake damage in the extant monuments of these cities. There are two historical monuments in the city of Hamédān: (1) the ‘Alaviān Tomb Shrine (National Heritage ID 94), with two different opinions about the date of its construction: (a) ca. 1147 C.E. (Mostafavi, 1953, 1967; Meshkāti, 1970; Hātami, 2000); and (b) ca. 1309–1316 C.E. (Herzfeld, 1922; Pope, 1934; Wilber, 1955). When Herzfeld (1922) studied the monument, the entire dome was missing, walls were preserved to below the dome drum, the entire wall of the mehrāb face was damaged, and portion of the entrance portal was destroyed. The monument was restored in 1938. (2) The second monument in Hamédān is the Jewish Esther and Mordékhāi (Mordechai) Tomb Shrine (National Heritage ID 216), which was apparently

206

Berberian et al.

built in the thirteenth century C.E. and rebuilt later (Meshkāti, 1970). No conclusive study has been carried out regarding the reconstruction phases and the cause of damages to the monument. The painting by Pascal Coste (Flandin and Coste, 1851) shows the structure with a high dome with no visible damage. It should be mentioned that there is no active fault in the Hamédān area, and the closest active fault to the city is the Nahāvand segment of the Zāgros Main Recent fault, located ~60 km to the SW of the city (Figs. 2, 6, and 16). However, the 956–957, November 1087, 1191, 1430, and 1495 C.E. earthquakes either caused damages or were strongly felt at the city (Fig. 6, inset top right; see also relevant references in Table 3). The Golpāyégān Saljuq congregational mosque (National Heritage ID 1757) was either built ca. 1105–1118 C.E. (Godard, 1965; Meshkāti, 1970; O’Kane, 1994), or 1120– 1135 C.E. (Pope, 1965, 1997; Figs. 2, 6, and 16). The monument has been restored several times since its construction, including during the reign of Fath ‘Ali Shāh Qājār (r. 1794–1834). Two grand square ayvāns were added to the NE and SW, and two small ayvāns were added to the NW and NE. The mosque has a single minaret with octahedral base located outside the precinct, behind the southern wall. The top and bottom of the minaret have been repaired. It is topped by a narrow turnet (goldasteh) placed off-center (Meshkāti, 1970; Blair and Bloom, 2012). As with the Hamédān case, no clear evidence of the 1316 C.E. earthquake is documented at Golpāyégān mosque. However, the extant Saljuq minaret at Golpāyégān indicates the absence of large-magnitude earthquakes at least for the past 877 yr at Golpāyégān. This brings us to the earthquake report of the 1316 C.E. earthquake by Kāshāni (ca. 1320, p. 179), who mentioned an event that took place between Hamédān and Golpāyégān (Fig. 16), but that the earthquake was far from both Hamédān and Golpāyégān. Golpāyégān is located ~80 km to the NE of the Dorud segment of the Zāgros Main Recent fault (Fig. 16). Contrary to Hamédān and Golpāyégān, which are far from the Zāgros Main Recent fault, and where no evidence of the 1316 C.E. earthquake damage exists (Fig. 16), the city of Dorud (33°29′N, 49°03′E) is located along the fault (55 km to the SE of Borujerd), and the town of Aznā (33°27′N, 40°27′E; 85 km to the SE of Borujerd and 78 km to the W of Golpāyégān) is situated ~25 km to the NE of the fault (Fig. 16). Dorud, which was completely destroyed during the 1909 Mw 7.4 earthquake (Figs. 6 and 22), has no historical monument, whereas, ~25 km to the NNE of Aznā (3 km east of the Aznā railroad station), there is a Timurid mausoleum of Qāsem and Zaid of Jāpalaq (Arabicized form of the ancient Persian “Gāpaleh”; Fig. 16), which was built in 738H/1337 C.E. (date on a wooden door not on the building) or 808 and 850 H (the latter on the wooden sanduq/chest; i.e., 1405 and 1446 C.E.; Meshkāti, 1970; Golombek and Wilber, 1988; National Heritage ID 1757). In any event, the monument was built (or rebuilt) after the 1316 C.E. earthquake, and there is no active fault near the town. The monument is located ~40 km to the NE of the Zāgros Main Recent fault (Fig. 16).

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084761/spe525-05.pdf by University of Wisconsin Madison user

Meshkāti (1970), followed by Golombek and Wilber (1988), took pictures and described the monument in their catalogues as located in Borujerd, which is not the case. The former mentioned that the monument is located about 4 km to the east of the Borujerd railroad station. The railroad from Arāk to Dorud does not pass through Borujerd; it passes through Aznā. The Timurid mausoleum of Qāsem and Zaid of Aznā was destroyed by the Sāzémān Owqāf (Religious Endowment Organization of Iran) and rebuilt unprofessionally; it has lost its historical and architectural value. CONCLUDING REMARKS No systematic study of archaeoseismology has been carried out in Iran, and the major phases of rebuilding of historical monuments have not been adequately investigated by archaeologists assisted by earthquake geologists (Berberian, 1994, 2014; Berberian and Yeats, 2001; Berberian et al., 2012, 2014). Despite the fact that no particular details of the reconstructions have survived in the literature, the two archaeological reports on the Borujerd congregational mosque (Mehryār, 1985; Moqaddas, 1997) provided an essential basis for initiation of such a study. This is the first attempt made to gather limited archaeological excavations, architectural signatures, historical accounts, and geologic, active faulting, and seismological data from the Borujerd area (in the vicinity of the Zāgros Main Recent fault) that would allow us to better understand the characteristics of hitherto unknown 1316 C.E. earthquake damage in the area, the epicenter of which was erroneously reported in the catalogues in an area ~80 km to the SE of Borujerd, far from any active fault, without any reported macroseismic data. We tried to learn more about the possible devastating effects of earthquakes in the city of Borujerd (Table 3). Our detailed collection of data and critical review of the available sources have produced a data set of information that seems to be consistent with effects from the 1316 C.E. largemagnitude earthquake, as well as other earthquakes, in the Borujerd area along the Zāgros Main Recent fault in western Iran. The available data are represented by stratigraphic layers and architectural elements demonstrating: (1) sudden collapse of edifices over frequented floors; (2) evidence of rebuilding or restoration phases; and (3) indication of abandonment of the collapsed megastructure for about a century after destruction by an earthquake during a century of turmoil in Iran. The date of the event can be defined through several types of archaeological and historic data, and thus the date of major destruction is very close to that of the hitherto unknown 1316 C.E. earthquake. In this case, the earthquake origin might be attributed to the activation of a segment of the Zāgros Main Recent fault located 4 km to the southwest of the city. The Borujerd congregational mosque has undergone numerous vicissitudes and restorations, some of which might have been necessitated by the occurrence of large-magnitude earthquakes (Table 3). Although archaeoseismic data from the study area suffer uncertain attribution due to the equivocal origin of

Evidence of historical seismic activity along the Zāgros Main Recent fault at Borujerd the analyzed architectural and stratigraphic units demonstrating destruction/damage and related effects, the reliability of the pictures obtained appears to be quite high in the near-field, considering the abundance of chronologically constrained sudden collapse phases, and the construction of several floor levels of the structure. Overall, the proposed archaeoseismic framework allows the destruction of the megastructure to be most probably attributed to the 5 January 1316 C.E. earthquake and be defined as a magnitude Mw ~≥7.0 earthquake, based on the intensity of damages documented from the megastructure of the Borujerd congregational mosque III, the length of the Nahāvand segment of the Zāgros Main Recent fault, and the seismic history of the fault (Fig. 22). There is no doubt that the city of Borujerd, built in the close vicinity of the active Zāgros Main Recent fault, was subjected to earthquakes of varying magnitudes throughout its long history, at least since the recorded history of the Sāssānid Dynasty (224–642 C.E.; Fig. 22; Tables 1 and 3). At the moment, all collected data strongly point to sudden collapse of the dome chamber, minaret, and other structures by the strong earthquake of 1316 C.E. Judging from the destruction of the Borujerd congregational mosque III (ca. post–1090/pre– 1139 C.E.) and the lack of other pre–fourteenth-century historical monuments in the area, the city might have been obliterated by this event. No earlier accounts of earthquakes have yet been found in the few surviving chronicles of the time. The 1316 C.E. earthquake possibly gave the coup-de-grâce to the city of Borujerd and destroyed almost all traces of its pre–fourteenth-century historical monuments, reducing the urban and rural buildings to ground level. The lack of data on the historical seismicity of Borujerd and other cities built along the Zāgros Main Recent fault reflects the inadequacy of our historical source material for Iranian rural and urban areas that were outside the main political and economic focal point of the central government. The city contains very few structures of historical interests, which were all built after the fourteenth century C.E. Comparison between the 1316 C.E. earthquake destruction and the damage pattern of the 31 March 2006 Mw 6.1 Chālānchulān earthquake, which struck the area 20 km to the south of the monument (Figs. 6 and 22), shows that the 1316 C.E. event was a large-magnitude earthquake (Mw ~≥7.0) and reveals consistency between the ancient earthquake and the activation of the major adjacent active fault, the Zāgros Main Recent fault (Figs. 6 and 22). Since no other historical events can be attributed to this particular segment of this active fault, we may conclude that the 700 yr time that has elapsed since the last fault activation is much less that the recurrence interval of ~1000 yr along the Nahāvand segment of the Zāgros Main Recent fault. Moreover, considering a regression of fault length, magnitude, displacement, and the GSP slip rates, the causative source is capable of a Mw ~7.4 earthquake with a return period of ~1000 yr. Comparison between the 1316 C.E. earthquake structural damage and the mentioned regression analysis (though we do not put much weight on it) suggests that the magnitude mentioned is consistent with the presumed effects of the 1316 C.E. earthquake on

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084761/spe525-05.pdf by University of Wisconsin Madison user

207

the Borujerd congregational mosque. An in-depth paleoseismic trench study is needed to confirm the seismic history of the fault near Borujerd. Archaeological and architectural data also show additional phases of destruction and damage that may possibly be attributed to distant or nearby earthquakes (Tables 1–3; Fig. 6) for which we do not have any proof but mentioned here for further in-depth analysis. For example: (1) The 956–967 earthquake (Fig. 6) could have been responsible for destruction of Borujerd congregational mosque I (ninth century, destroyed before 971 C.E.); however, this requires further multidisciplinary study. (2) The congregational mosque II (ca. 945–971 C.E.) brick pillars are tilted, but time and cause of damage are not known. (3) The 1139–1145 C.E. major reconstruction phase of the Borujerd congregational mosque III lofty dome chamber addressed on an inscription was carried out shortly after three earthquakes of 27 February 1130, 25 July 1135, and 13 August 1135 C.E. (Table 3; Fig. 6); their connection to this repair has not been investigated. (4) Simultaneous collapse of the supplementary mudbrick ribbed vaults and arches in the dome chamber basement have not been dated, so we cannot address the time and cause of destruction. (5) Additional post–1405–1447 C.E. damage in the dome chamber area above the floor is a promising archaeoseismic case that has not yet been studied or dated. In general, due to the lack of radiometric dating and limited excavation data, the unconstrained destruction-rebuilding dates are still an open debate. Obviously, in order to improve our understanding of earthquake hazards and their effects on the structures in and around the city of Borujerd, future multidisciplinary archaeological excavations should be designed to carefully investigate the archaeoseismologic indicators left in the archaeological records, as well as to conduct paleoseismic trenching across the Zāgros Main Recent fault. This will eventually provide specific insight into a better earthquake history and recurrence period in the study area. No earthquake of any significance has been recorded in Borujerd since the 5 January 1316 C.E. earthquake (not including numerous small- to medium-magnitude earthquakes), justified by: (1) extant historic structures; (2) the accounts of the string of European travelers passing through Borujerd; and (3) later emergence of the Persian newspapers from 7 February 1851. This indicates a minimum 700-yr-long period of relative quiescence of large-magnitude earthquakes in Borujerd along the Nahāvand segment of the Zāgros Main Recent fault (Fig. 22). This quiescence is not surprising. As mentioned earlier, regression analysis indicates that the Nahāvand fault segment of the Zāgros Main Recent fault is capable of generating earthquakes of magnitude ~Mw 7.4 with maximum and average right-lateral displacements of ~3.9 m and ~2.2 m, respectively. Considering the average slip rate of ~2.3 mm/yr of the Zāgros Main Recent fault, such an earthquake is expected to happen every ~1000 yr.

208

Berberian et al.

Finally, the Borujerd monument is located in a highly seismic zone in the vicinity of a major active fault. Throughout its life span of more than ~12 centuries, the structure shows clear damages caused by historic earthquakes. The building has not been properly restored or retrofitted, and the 2006 Mw 6.1 earthquake, which occurred ~20 km to its south, clearly showed that the structure is seismically vulnerable (Figs. 6 and 22). Its past vulnerability should be used as a subject to understand the structural and material behavior of this monument and ultimately properly retrofit it to resist the future earthquakes. This study significantly enhances not only our understanding of the regional earthquake periodicity, but also our ability to reduce seismic hazard in the region, though it raises more questions than it possibly resolved. Vulnerability of historic buildings located near active faults should be taken seriously by the authorities and must be carefully studied, and the monuments need to be protected against future strong ground motions. Paleoseismic fault-trench studies coupled with detailed archaeoseismic investigation and radiometric dating of archaeological stratigraphy and active alluvial/fluvial deposits should constrain earthquake chronology, their source parameters, and periodicity. Details of damage to other historical monuments in the region possibly affected by the 1316 C.E. earthquake, which have been rebuilt since then, have not yet been studied or documented, and we hope this paper will trigger such an interest. The complete picture of destruction and damage pattern and other devastating outcomes of the 1316 C.E. earthquake are not known in the region. Due to lack of literary sources, we have not been able to study the socioeconomic impact of the event and the way in which the inhabitants of the epicentral region reacted to and recovered from this earthquake, and the impact to the water resources is unknown. The ultimate aim of this project was to make a substantial contribution to a better understanding of the earthquake history of this region (in particular, the distribution of seismicity through time and space) by integrating archaeoseismology and paleo-architecture as a complementary source of data in the whole range of approaches. By providing a quantitative assessment of the reliability of the archaeoseismological evidence and trying to translate this evidence into earthquake-related parameters, a significant contribution to seismic hazard studies could be achieved. We emphasize the need for an integrated approach by a multidisciplinary team of archaeologists, earthquake geologists, geomorphologists, geophysicists, seismologists, architects, and historians, ideally all working in concert during the excavation. Even then, it is acknowledged that interpretations will still be confronted with a number of assumptions, much subjectivity, and varying degrees of uncertainty. ACKNOWLEDGMENTS We are grateful to Manuel Sintubin, Stathis C. Stiros, and Rasoul Sorkhābi for their helpful comments and useful recommendations to improve the quality of this work. However, we

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084761/spe525-05.pdf by University of Wisconsin Madison user

remain responsible for the content of the paper. We thank Fereydun Biglari, Mehdi Rahbar, Sajād ‘Alibaigi, Gudarzi (Cultural Heritages, Handicrafts and Tourism Organization, Borujerd), Hossein Jorbozédār (Tehran), and Kāmrān Ansāri (Tehran) for providing additional data. The research for this project was not supported by any grant or organization. This work is dedicated to the memory of archaeological and architectural efforts conducted by Mohammad Mehryār (1939–2004), and to the world-renowned composer and conductor, maestro Loris Tjeknāvoriān, who was born in the city of Borujerd in 1937 and returned to live in Iran in 2000. REFERENCES CITED Abu Hāmed Mohammad ben Ebrāhim, 1203, Zail-e Saljuqnāmeh-ye Zahir alDin Nayshāburi, 599/1203, in Mirzā Esmā’il Afshār (Hamid al-Molk) and Mohammad Ramezāni, eds., Saljuqnāmeh, Ta’lif-e Zahir al-Din Nayshaburi [d. 582 H/1186 CE], be-enzemām-e Zail-e Saljuqnāmeh-ye Abu Hāmed Mohammad ben Ebrāhim: Tehran, Golāleh Khāvar Publishers, 1332/1953, 89 p. [in Persian]. Abu’l Fida [Abolfedā], Isma’il Ibn’Ali, 1321, al-Mukhtasar fi Akhbar alBashar (History of Abu’l Fida [Tārikh-e Abolfedā], Events from the Earliest Times to the Time of the Writer’s Death; 721/1321): Tehran, Library of the University of Tehran, ed. Istanbul 1286 H/1869 CE, Mimeograph, 4 volumes. Also Kitāb al-Mukhtasar fi Akhbār al-Bashar, Tārikh abi al-Fedā, Dhakhā’ir al-Arab, Imāduddin Ismail, ed., 1907: Hossainiah Publishing House, Cairo, Egypt, 2 volumes. Afshar, I., 1979a, Architectural information through the Persian classical texts, in Akten des VII Internationalen Kongress für Iranische Kunst Archäologie, München, 7–10 September 1976: Berlin, D. Reimer Publisher, p. 612–616, ISBN: 3496001038. Afshar, I., 1979b, Moqadamäti dar-Bāreh-ye Tārikh-e Me’māri dar Irān bar Asās-e Motun-e Fārsi (An Introduction to the History of Archaeology in Iran based on the Iranian Treatises): Honar va Mardom, v. CLXXIII, no. 2535, p. 2–5 [in Persian]. al-Balādhuri (Balāzuri), Ahmad ibn Yahyā, 890, Kitab Futuh al-Buldān [Book of the Conquests of Lands], Hitti, P.K., tr. (v. 1), and Murgotten, E.C., tr. (v. 2), The Origins of the Islamic State: New York, 1916–1924. Persian tr. by Āzarmāsh āzarnush: Tehrān Bonyād Farhang Irān Publisher, 1346/1967; Tehran, Soroush Publisher, 1364/1985 [in Persian]. Also, Liber Expugnationis Regionum, ed. De Goeje: Leiden, Netherlands, Brill, 1866. Alipoor, A., Zaré, M., and Ghassemi, M.R., 2012, Inception of activity and slip rate on the Main Recent fault of Zagros Mountains, Iran: Geomorphology, v. 175–176, p. 86–97, doi:10.1016/j.geomorph.2012.06.025. al-Jauzi, I., ‘Abd al-Rahmān ebn ‘Ali Ibn, 1181, al-Montazam fi Tārikh alMoluk val-Omam, 577 H/1181 (Events of the Years 257/870 to 574/1178), ‘Abdul-Qader Atā, M., and ‘Abdul-Qader ‘Atā, M., eds: Dār al-Ketāb al‘Elmiya, Beirut, Lebanon, 1412/1992, 19 vols., https://archive.org/details/ muntazim_tarikh_mlouk_oumm [in Arabic]. Allen, M., Jackson, J., and Walker, R., 2004, Late Cenozoic reorganization of the Arabia-Eurasia collision and the comparison of short-term and long-term deformation rates: Tectonics, v. 23, TC2008, doi:10.1029/2003TC001530. al-Qusi (al-Qowsi), Ahmad ebn al-’Allāma, 1907, Al-Barākin wa’l-Zalāzil, 1325 H [Volcanoes and Earthquakes]: Cairo, Egypt, Cairo National Library Manuscript, Tabi’iyāt no. 114 (Taimur). al-Suyuti, Jalal al-Din, 1505, Kashf al-Salsalah ‘an Wasf al-Zalzalah, 911/1505 [The Transmitted Expositions Concerning the Description of the Earthquake of Doomsday]: ‘Abd al-Latif Sa’adāni, ed., Fez. 1971. French tr., S. al-Nejjār: Rabat, Morocco, Cahiers du Centre Universitaire de la Recherche Scientifique, 1974. al-‘Umari (al-‘Omari), Yāsin ebn Khairallāh al-Khatib, 1793, Al-āthār al-Jaliya fi al-Hawādith al-Arāiya, 1208 H/1793 CE [Obvious Relics of the Earth Related Events]: Najaf al Ashraf, Iraq, Kashf al-Ghota Library Manuscript #157,18,1322 (mimeograph, Tehran). Also Manuscript of the British Library #OR.6300 (an Annualistic Work of the ‘Othmān Period). Also Baghdād, Iraq, Library of Iraq Academy manuscript. Zubdat al-āthār al-Jāliya: Baghdād, Ed. Ra’uf, 1974.

Evidence of historical seismic activity along the Zāgros Main Recent fault at Borujerd Ambraseys, N.N., 1974, Historical seismicity of north-central Iran, in Materials for the Study of Seismotectonics of Iran; North Central Iran: Geological Survey of Iran, v. 29, p. 47–96. Ambraseys, N.N., 1978, The relocation of epicenters in Iran: Geophysical Journal of the Royal Astronomical Society, v. 53, p. 117–121. Ambraseys, N.N., 1979, A test case of historical seismicity: Isfahan and chaharMahal, Iran: The Geographical Journal, v. 145, no. 1, p. 56–71. Ambraseys, N.N., 2005, Archaeoseismology and neo-catastrophism: Seismology, v. 76, p. 560–564. Ambraseys, N.N., 2009, Earthquakes in the Mediterranean and Middle East; A Multidiciplinary Study of Seismicity up to 1900: Cambridge, UK, Cambridge University Press, p. 9–12. Ambraseys, N.N., and Melville, C.P., 1977, The seismicity of Kuhistan, Iran: The Geographical Journal, v. 143, p. 179–199. Ambraseys, N.N., and Melville, C.P., 1982, A History of Persian Earthquakes: Cambridge, UK, Cambridge University Press, 219 p. Ambraseys, N.N., and Moinfar, A.A., 1973, The seismicity of Iran, the Silakhor, Lurestan, earthquake of 23rd January, 1909: Annali di Geofisica, v. XXVI, no. 4, p. 659–678. Ambraseys, N.N., and Moinfar, A.A., 1974, The seismicity of Iran; the Firuzabad (Nehavand) earthquake of 16 August 1958: Annali di Geofisica, v. XXVII, no. 1–2, p. 1–21. Ambraseys, N., and Psycharis, I.N., 2011, Earthquake stability of columns and statues: Journal of Earthquake Engineering, v. 15, p. 685–710. Anonymous, 982, Hudud al-‘Alam men al-Mashreq el al-Maghreb, 372H/982 CE [The Limits of the World from the East to the West], a Persian geography written by an anonymous writer, 372 H/982 CE, Manuchehr Sotudeh, ed.: Tehran, Tāhuri Publishers, 1362/1983 [in Persian]. Also Hudud al-’Alam, trans. and comments by V. Minorsky (2nd ed.): London, Bosworth, 1970. Argyriou, N., Ktenidou, O.J., Manakou, M., Apostolidis, P., Chavez Garcia, F.J, and Pitilakis, K., 2007, Seismic response analysis of ancient columns: 4th International Conference on Earthquake Geotechnical Engineering, Thessaloniki, Greece, June 25–28, 2007, Paper No. 1659, 18 p., https://nikosargyriou.files.wordpress. com/2010/01/1659_arg.pdf. Āshtiāni, E., 1989, Tārik-e Moghol va Avā’el-e Taimuri, 1367/1989 [The History of the Mongols and Ealry Timurids]: Tehran, Nashr Nāmak Publisher [in Persian]. Authemayou, C., Chardon, D., Bellier, O., Malekzade, Z., Shabanian, E., and Abbassi, M., 2006, Late Cenozoic partitioning of oblique plate convergence in the Zagros fold-and-thrust belt (Iran): Tectonics, v. 25, TC3002, p. 1–21, doi:10.1029/2005TCOO1860. Authemayou, C., Bellier, O., Chardan, D., Benedetti, L., Malekzadeh, Z., Claude, C., Angeletti, B., Shahabian, E., and Abbasi, M.R., 2009, Quaternary slip-rates of the Kazerun and the Main Recent faults, active strikeslip partitioning in the Zagros fold-and-thrust belt: Geophysical Journal International, v. 178, p. 524–540. Āzarnoush, M., 2009, New evidence on the chronology of the Anahita temple: Iranica Antiqua, v. XLIV, p. 393–404. Bachmanov, D.M., Trifonov, V.G., Hessami, K.T., Kozhurin, A.I., Ivanova, T.P., Rogozhin, E.A., Hademi, M.C., and Jamali, F.H., 2004, Active faults in the Zagros and central Iran: Tectonophysics, v. 380, p. 221–241. Berberian, M., 1976, Contribution to the Seismotectonics of Iran (part II), in Berberian, M., ed., Geological Survey of Iran, Tehran, v. 39, 518 p. [in English]. Berberian, M., 1977, Contribution to the seismotectonics of Iran (part III), in Berberian, M., ed., Geological and Mining Survey Iran, Tehran, v. 40, 300 p. [in English]. Berberian, M., 1979, Evaluation of the instrumental and relocated epicentres of Iranian earthquakes: Geophysical Journal of the Royal Astronomical Society, London, v. 58, p. 625–630. Berberian, M., 1981, Active faulting and tectonics of Iran, in Gupta, H.K., and Delany, F.M., eds., Zagros–Hindu Kush–Himalaya Geodynamic Evolution: American Geophysical Union Geodynamics Monograph 3, p. 33–69. Berberian, M., 1983, The Southern Caspian, a compressional depression floored by a trapped, modified oceanic crust: Canadian Journal of Earth Sciences, v. 20, no. 2, p. 163–183. Berberian, M., 1994, Natural Hazards and the First Earthquake Catalogue of Iran, Volume 1—Historical Hazards in Iran Prior to 1900: Tehran, International Institute of Earthquake Engineering and Seismology (IIEES), UNESCO/IIEES Publication during United Nations International Decade for Natural Disaster Reduction (IDNDR), 649 p.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084761/spe525-05.pdf by University of Wisconsin Madison user

209

Berberian, M., 1995, Master blind thrust faults hidden under the Zagros folds; active basement tectonics and surface morphotectonics: Tectonophysics, v. 241, p. 193–224. Berberian, M., 1997, Seismic sources of the Transcaucasian historical earthquakes, in Giardini, D., and Balassanian, S., eds, Historical and Prehistorical Earthquakes in the Caucasus: Dordrecht, the Netherlands, Kluwer Academic Publishers, NATO ASI Series 2, p. 233–311. Berberian, M., 2005, The 2003 Bam urban earthquake, a predictable seismotectonic pattern along the western margin of the rigid Lut block, southeast Iran: Earthquake Spectra, v. 21, no. S1, p. S35–S99, doi:10.1193/1.2127909. Berberian, M., 2014, Earthquakes and Coseismic Surface Faulting on the Iranian Plateau; A Historical, Social, and Physical Approach: Amsterdam, Elsevier, Developments in Earth Surface Processes 17, 776 p., ISBN: 9780-444-63292-0. Berberian, M., and King, G.C.P., 1981, Towards a paleogeography and tectonic evolution of Iran: Canadian Journal of Earth Sciences, v. 18, no. 2, p. 210–265. Berberian, M., and Yeats, R.S., 1999, Patterns of historical rupture in the Iranian Plateau: Bulletin of the Seismological Society of America, v. 89, no. 1, p. 120–139. Berberian, M., and Yeats, R., 2001, Contribution of archaeological data to studies of earthquake history in the Iranian Plateau: Paul Hancock Memorial Issue: Journal of Structural Geology, v. 23, p. 536–584. Berberian, M., and Yeats, R.S., 2016, this volume, Tehran: An earthquake time bomb, in Sorkhabi, R., ed., Tectonic Evolution, Collision, and Seismicity of Southwest Asia: In Honor of Manuel Berberian’s Forty-Five Years of Research Contributions: Geological Society of America Special Paper 525, doi:10.1130/2016.2525(04). Berberian, F., Muir, I.D., Pankhurst, R.J., and Berberian, M., 1982, Late Cretaceous and early Miocene Andean-type plutonic activity in northern Makran and Central Iran: Journal of the Geological Society, London, v. 139, no. 5, p. 605–614. Berberian, M., Malek Shahmirzādi, S., Nokandeh, J., and Djamāli, M., 2012, Archaeoseismicity and environmental crises at the Sialk mounds, Central Iranian Plateau, since the Early Neolithic: Journal of Archaeological Science, v. 39, p. 2845–2858. Berberian, M., Petrie, C.A., Potts, D.T., Asgari Chaverdi, A., Dusting, A., Sardari Zarchi, A., Weeks, L., Ghāssemi, P., and Noruzi, R., 2014, Archaeoseismicity of the mounds and monuments along the Kāzerun fault (western Zāgros, SW Iranian Plateau) since the chalcolithic period: Iranica Antiqua, v. 49, p. 1–81, doi:10.2143/1A.49.0.3009238. bhrc.ir, 2006, Building and Housing Research Center, Tehran, Iran: http://www .bhrc.ac.ir/portal/Default.aspx?tabid=543 (accessed February 2014). Blair, D., and Bloom, J., 2012, Golpayegan Jame Mosque: www.ArchNet.org/ library/sites/one-site.jsp?site_id=11201 (accessed December 2012). Boyle, J.A., ed., 1968, The Cambridge History of Iran, Volume 5—The Saljuq and Mongol Periods: Cambridge, UK, Cambridge University Press, 778 p. Cais, 2005, Remnants of Ancient Castle Unveiled in Nahavand: http://www .cais-soas.com/News/2005/July2005/09-07.htm (accessed 9 July 2005). Coffin, D.R., 2004, Pirro Ligorio; The Renaissance Artist, Architect, and Antiquarian Library Building: University Park, Pennsylvania, Pennsylvania State University Press, 242 p. Cultural Heritage, Tourism, and Handicraft Organization of Iran (CHTHO), 2011, Cultural Heritage, Tourism, and Handicraft Organization of Iran: ichhto.org (accessed March 2011). Dabir Siyāqi, M., ed., 1962, Safar-nāmeh-ye Khuzestān-e Haj ‘Abdol Ghaffār Najm al-Molk, 1341/1962 [Travel Notes to Khuzestān]: Tehran, Mo’asseseh Matbu’āti ‘Elmi [in Persian]. Daryāee, T., ed., 2012, The Oxford Handbook of Iranian History: Oxford, UK, Oxford University Press, 432 p. De Morgan, J., 1894–1905, Mission Scientific en Perse: Paris, Orls, 2 volumes on geographical studies. Also tr. into Persian by Kāzem Vadi’ee, Motāle’āt-e Joghrāfi-ye Iran, v. II: Chehr Tabriz Publisher, 155 p. Dodge, H., 1984, Building Material and Techniques in the Eastern Mediterranean from the Hellenic Period to the Fourth Century AD [Ph.D. thesis]: Newcastle upon Tyne, UK, Newcastle University, 263 p. Ehlers, E., 1989, Borujerd: Encyclopedia Iranica, v. IV, fasc. 4, p. 375–376; also http://www.iranicaonline.org/articles/borujerd-town-and-sahrestan -in-lorestan (accessed January 2012). Engdahl, E.R., Jackson, J.A., Mters, S., Bergman, E.A., and Priestley, K., 2006, Relocation and assessment of seismicity in the Iran region: Geophysical Journal International, v. 167, p. 761–778.

210

Berberian et al.

Estakhri, Abu Eshāq Ebrāhim ebn Mohammad Fārsi, 951, Masālek va Mamālek, 350/951, Iraj Afshar, ed.: Tehran, Bongāh Tarjomeh va Nashr Ketāb, 1340/1961, 1347/1968, 165 p. [in Persian]. Flandin, E., and Coste, P., 1851, Voyage en Perse MM. Eugéne Flandin, Peintre, et Pascal Coste, architecte, Pendant les Années 1840 et 1841: Paris, Gide et J. Baury Libraires Editeurs, 1851 (v. I), 1854 (v. II). Also in Persian (Safar-nāmeh Eugene Flandin be Irān), Hosseisn Nur-Sādeqi, tr.: Tehran, Eshrāqi Publisher, 1356/1977. Fleming, J., Honour, H., and Pevsner, N., eds., 1991, Dictionary of Architecture (4th ed.): New York, Penguin Books, ISBN: 9780140512410. Frye, R.N., 1975, The Cambridge History of Iran, Volume 4—The Period from the Arab Invasion to the Saljuqs: Cambridge, UK, Cambridge University Press, 734 p. Galadini, F., Hinzen, K.-G., and Stiros, S., 2006, Archaeoseismology, methodological issues and procedure: Journal of Seismology, v. 10, p. 395–414. Gazetteer of Persia, 1885–1918, Gazetteer of Persia: London, India Office, 4 volumes. Gidon, M., Berthier, F., Billiault, J.P., Halbronn, B., and Maurizot, P., 1974, Sur le caractères et l’ampleur du coulissement de la ‘Main Fault’ dans la region de Borujerd–Dorud, Zagros oriental, Iran: Comptes Rendus de l’Académie des Sciences Paris, ser. D, v. 278, p. 701–704. Godard, A., 1965, The Art of Iran: New York, Frederick A. Praeger Publishers, 358 p. Golombek, L., and Wilber, D., 1988, The Timurid Architecture of Iran and Turan: Princeton, New Jersey, Princeton University Press, 2 volumes; 510 p. (v. 1). Guidoboni, E., 1997, An early project for an antiseismic house in Italy; Pirro Ligorio’s treatise of 1570–1574: European Earthquake Engineering, v. 2, p. 13–20. Guidoboni, E., 2015, When towns collapse; images of earthquakes, floods, and eruptions in Italy in the fifteenth to nineteenth centuries, in Folin, M., and Preti, M., eds., Wounded Cities; The Representation of Urban Disasters in European Art (14th–20th Centuries): Leiden, the Netherlands, Koninklijke Brill, p. 33–58. Gye, D.H., 1988, Arches and domes in Iranian Islamic buildings; an engineering perspective: Iran, British Institute Journal of Persian Studies, v. 26, p. 129–144, doi:10.2307/4298807. Hamédāni, Rashid al-Din Fazlollāh, 1311, Jāme’ al-Tavārikh Rashidi [the History of Rashidi], Karimi, B., ed.: Tehran, Eqbāl Publisher, 1338/1995 [in Persian]. Also Rowshan, M., ed., Jāme’ al-Tavārikh (Tārikh Iran va Eslam): Tehran, Markaz Pazhuhesh Mirāth Maktub, 1392/2013, 3 volumes [in Persian]. Harvard, 2016, Global Centroid Moment Tensor Catalog: http://www .globalcmt.org/CMTsearch.html (accessed January 2016). Hātami, Gh. ‘A., 2000, Me’māri-ye Eslāmi-yeh Irān dar Daureh-ye Saljuqiān, 2000 [Islamic Architecture of Iran during the Saljuqs]: Tehran, Enteshārāt Majed; Mo’assesseh-ye Enteshārāt-e Jahād-e Dāneshgāhi, v. 39, 1379 [in Persian]. Hazin Borujerdi, H., 1972, Tazkareh-ye Hossein Hazin, yā Durnamā’ee az Shahrestān-e Borujerd [Hossein Hazin’s Notes, or, Perspective of the Borujerd County]: Khorramabad, Dānesh Publisher, 1351 [in Persian]. Hebraeus, B., 1286, The Chronography of Gregory Abu’l Faraj, 1225–1286, the Son of Aaron, the Hebrew Physician, Commonly Known as Bar Hebraeus; Being the First Part of His Political History of the World, Translated from Syriac by E.A. Willis Budge, 2 volumes: London, Oxford University Press, 1932. Reprinted: Amsterdam, Apa-Philo Press, 1976. Reprinted: Gorgias Press, LLC, 2003. Also: http://rbedrosian.com/BH/BH.html. Herzfeld, E.E., 1922, Die Gumbadh-I ‘Alawiyyān und die Baukunst der Ilkhane in Iran: Oriental studies, in Walker Arnold, T., and Alleyne, R., eds., A Volume of Oriental Studies Presented to Edward G. Browne on His 60th Birthday: Cambridge, UK, Cambridge University Press, p. 186–199. Hinzen, K.-G., 2005, The use of engineering models to interpret archaeoseismological findings in Tolbiacum, Germany; a case study: Bulletin of the Seismological Society of America, v. 95, no. 2, p. 521–539, doi:10.1875/0120040068. Hinzen, K.-G., 2009, Simulation of toppling columns in archaeoseismology: Bulletin of the Seismological Society of America, v. 99, no. 5, p. 2855– 2875, doi:10.1885/0120080241. Hinzen, K.-G., 2011, Sensitivity of earthquake-toppled columns to small changes in ground motion and geometry: Israel Journal of Earth Sciences, v. 58, p. 309–326, doi:10.1560/IJES.58.3-4.309. Hinzen, K.-G., Fleischer, C., Reamer, S.K., Schreiber, S., Schutte, S., and Yerli, B., 2011, Quantitative methods in archaeoseismology: Quaternary International, v. 242, no. 1, p. 31–41.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084761/spe525-05.pdf by University of Wisconsin Madison user

Hinzen, K.G., Kehmiere, H., and Schreiber, S., 2013, Quantitative archaeoseismological study of a Roman mausoleum in Pinaro (Turkey)— Testing sesismogenic and rockfall damage scenarios: Bulletin of the Seismological Society of America, v. 103, no. 2a, p. 1008–1021, doi:10.1785/0120120245. Hutt, A., 1984, Iran, in Michel, G., ed., Architecture of the Islamic World: New York, Thames and Hudson, p. 251–258. Hutt, A., and Harrow, L., 1977, Islamic Architecture, Iran 1: London, UK, Scorpion Publications Ltd., 192 p. Hutt, A., and Harrow, L., 1978, Islamic Architecture, Iran 2: London, UK, Scorpion Publications Ltd., 177 p. Ibn al-Athir, I., ‘Ez al-Din ‘Ali, 1231, Tarik-e-Kamel, 628/1231 [al-Kāmel fi al-Tārikh, The Complete History; 628/1231], Mohammad Hossain Rohāni, tr.: Tehran, Asātir Publisher, 1371/1992, 446 p. Also Kāmel Tārikh Bozorg Eslām va Irān [The Complete History of Islam and Iran], ‘Abbās Khalili, tr., Mahyār Khalili, ed.: Tehran, Sherkat Sahāmi Chāp va Enteshārāt Kotob Irān Publisher. Also Ibn al-Athir, ‘Izz al-Din, al-Kamil fi’l-Tārikh, ed.: Leiden, Netherlands, C.J. Tornberg, 1851–1876. Ibn al-Faqih Hamédani [ebn-e-Faqih Hamédani], Abubakr Ahmad ebn Mohammad, 903, Akhbār al-Boldān, 290 H [News of the Cities; ca. 903], Mas’ud, H., tr.: Tehrān, Manābe’ Tārikh va Joghrāfiyā-ye Irān, no. 35, 1349/1370 [in Persian]. Also Ibn al-Faqih, Ahmad ibn Muhammad, Mukhtasar Kitab al-Buldan, de Goeje, ed.: Leiden, Netherlands, Bibliotheca Geographorum Arabicorum, BGA v, 1885. Also Massé, H., tr., Damascus, Abrégé du Livre des Pays, 1973. Ibn Hawqal Nasibi, Abolqāsem Mohammad Baghdādi, 978, Saurat al-Arz, 367 H [Configuration of the Earth, 978], Ja’far Sho’ār, tr. (Safarnāmeh-ye Ebn-e Huqal ya Surat al-Arz): Tehran, Bonyād Farhang Irān Publisher, no. 2, 1345/1966, 111 p. Also: Tehran, Enteshārāt Amir Kabir Publisher, 2nd ed., 1366/1987 [in Persian]. International Institute of Earthquake Engineering and Seismology (IIEES), 2005, Gozāresh zaminlarzeh 13 Ordibehesht 1384 Tabarijān Borujerd, Ml 4.9 [The May 3, 2005 Tabarijān, Borujerd, Earthquake of Ml 4.9]: Tehran, International Institute of Earthquake Engineering and Seismology, 19 p. [in Persian], http://www.iiees.ac.ir/bank/brujerd/brujerd84213.html. International Institute of Earthquake Engineering and Seismology (IIEES), 2006, Gozāresh Shenāsā’ee-yeh zelzeleh-yeh 1385/1/11 Darb-e Āstāneh Silākhor [Reconnaissance Report of the Farvadrin 11, 1385 Darb-e Āsténeh, Silākhor Earthquake]: Tehran, International Institute of Earthquake Engineering and Seismology, 233 p. [in Persian]. International Seismological Centre (ISC), 2016, International Seismological Centre: http://www.isc.ac.uk/ (accessed January 2016). Iranian Gazetteer, 1330/1951, Ed. Lotfollāh Mafkham Pāyān, 11 volumes: Tehran, Iranian Gazetteer. Also: Tehran, Plan and Budget Organization of Iran (Farhang-e ābādihāyeh Keshvar), and the Iranian National Geographic Organization of Armed Forces (Farhang-e Joghrāfiyā’ee ābādihāyeh Keshvar Jomhuri Eslāmi Iran), 120 volumes [in Persian]. irna.ir, 2005, Islamic Republic News Agency: Tehran, Iran: www.irna.ir (accessed February 2014) [in Persian]. Jackson, J.A., and McKenzie, D.P., 1984, Active tectonics of the AlpineHimalayan belt between western Turkey and Pakistan: Geophysical Journal of the Royal Astronomical Society, v. 77, p. 185–264. Jackson, J.A., and McKenzie, D.P., 1988, The relationship between plate motions and seismic tensors, and the rates of active deformation in the Mediterranean and Middle East: Geophysical Journal of Royal Astronomical Society, London, v. 93, p. 45–73. Jackson, P., and Lockhart, L., 1986, The Cambridge History of Iran, Volume 6—The Timurid and Safavid Periods: Cambridge, UK, Cambridge University Press, 1120 p. Ja’fari Moqaddas, M.J., 1975, Joghrāfiyā-yeh Shahrestān-e Borujerd [Geography of the Borujerd Township]: Tehran, Sāzémān Joghrāfiāi-e Keshvar Publisher, 16 p. Jahn, K., ed., 1940, Gerschiche Gāzān-Hān’s aus dem Tarih-i-Mubarah-iGazani des Rashid al-Din Fadlallah b. ‘Imad al-Daula Abul-Hair: London, Luzac, 387 p. Jazabi, ‘A.R., 1985, The lines of the Iranian arches based on Ghiyath al-Din Jamshid Kashani: Āthār, Iranian Cultural Heritage Organization, no. 10 and 11, p. 166–174 [in Persian]. Jusseret, S., Langohr, C., and Sintubin, M., 2013, Tracking earthquake archaeological evidence in Late Minoan IIIB (~1300–1200 B.C.) Crete (Greece); a proof of concept: Bulletin of the Seismological Society of America, v. 103, no. 6, p. 3026–3043, doi:10.1785/0120130070.

Evidence of historical seismic activity along the Zāgros Main Recent fault at Borujerd Kāmbakhsh-Fard, S., 1994, The Anahita Temple Kangavar; Archaeological Excavations and Survey, The Reconstruction and Architectural Restoration of the Nahid Temple and Taghe-Gera: Tehran, Iranian Cultural Heritage Organization, Publ. no. 23, 405 p. [in Persian]. Kāshāni, Abolqāsem ‘Abdollāh ebn Mohammad al-Qāshāni, ca. 1320, Tārikh-e Oljāitu, Tārikh-e Pādeshāh Ghiyāth al-Din val-donyā Oljāitu Soltān Tayebollāh Marqada [History of Oljāitu] Hambali, M., ed.: Tehran, Bogāh Tarjomeh va Nashr-e Ketāb, no. 313, Majmu’eh-ye Motun-e Fārsi, no. 40, 1348/1969, 295 p. [in Persian]. Le Strange, G., 1905, The Lands of the Eastern Caliphate; Mesopotamia, Persia, and Central Asia from the Moslem Conquest to the Time of Timur: Cambridge, UK, Cambridge University Press, 536 p. Also: Tehran, Enteshārāt Donyā-yeh Ketāb, 1st ed., 1362/1983 [in Persian]. Lewcock, R., 1984, Architects, craftsmen and builders; materials and techniques, in Michell, G., ed., Architecture of the Islamic World: Its History and Social Meaning: London, Thames and Hudson, p. 112–143. Lotfizādeh, Sh., 1980, Shivehhā-ye Me’māri-ye Irān, Shiveh-ye Khorāsāni, 1980 [Architectural Styles of Iran; the Khorāsāni Style]: Tehran, Āthar, Sāzémān Melli Hefāzat Āthār Bāstāni, Spring 1359/1980, v. 1, p. 108– 122, 1359 [in Persian]. Maulānā Borujerdi, Gh., 1974, Tārikh-e Borujerd; Joghrāfiā, Hokkām, Ruidād-hā, 1353 [The History of Borujerd; Geography, Rulers, Events]: Tehran, Sadr Publisher, p. 9 [in Persian]. McGuire, W.J., Griffith, D.R., Hancock, P.L., and Stewart, I.S., eds., 2000, The Archaeology of Geological Catastrophes: Geological Society, London, Special Publication 171, 417 p. McKenzie, D.P., 1972, Active tectonics of the Mediterranean region: Geophysical Journal of the Royal Astronomical Society, London, v. 30, p. 109–185. Meghraoui, M., Gomez, F., Sbeinati, R., Van der Woerd, J., Mounty, M., Darkal, A.N., Raswan, Y., Layyous, I., Al-Najjar, H., Darawcheh, R., Hijazi, F., Al-Ghazzi, R., and Barazangi, M., 2003, Evidence for 830 years of seismic quiescence from paleoseismology, archaeoseismology, and historical seismicity along the Dead Sea fault in Syria: Earth and Planetary Science Letters, v. 201, no. 1–2, p. 35–52. Mehryār, M., 1985, Barresi-yeh moqadamāti-ye Masjed Jāmé-ye Borujerd [Preliminary Study of the Borujerd Jāmé Mosque]: Tehran, Āthar, Sāzémān Melli Hefāzat Āthār Bāstāni, v. 10 and 11, p. 77–165 (Bahman 1363) [in Persian]. Melville, C., 2008, Jame al-Tavarikh: Encyclopedia Iranica, v. XIV, no. 5, p. 462–468; iranicaonline.org (accessed in 2012). Meshkāti, N., 1970, Fehrest-e-Banāhā-ye Tārikhi va Amāken-e-Bāstāni-e-Irān [Catalogue of Historical Buildings and Ancient Places of Iran]: Tehran, Sāzemān Melli Hefāzat Āsār Bāstāni Irān, no. 1, 1349, 370 p. [in Persian]. Moqaddas, M., 1997, Tārikhche-yeh Mokhtasar-e Majed Jame Borujerd [Brief History of the Borujerd Jame Mosque], in Proceedings of the Bam Symposium on Architecture and Urban Planning, Arg-e Bam, Kermān: Tehran, Iranian Cultural Heritage Organization, no. 92, v. 5, p. 135–206, 1376 [in Persian]. Moqaddasi, Shams al-Din Abu ‘Abdollah Mohammad ebn Ahmad ebn Abibakr ebn Shami, 985, Ahsan al-Taqāsim fi Ma’refat al-Aqālim (375), ‘Alinaqi Monzavi, tr.: Tehran, Sherkat Mo’alefān va Motarjemān Irān, 2 volumes, 1361/1982. Mostafavi, M.T., 1953 and 1967, Negāhi be Honar-e Me’māri-ye Irān [A Glance at the Art of Architecture in Iran]: Tehran, Enteshārāt Sherkat Sahāmi Simān Shomāl, 1332 p., 1346 [in Persian]. Nabavi, M.S., 1972, The Seismicity of Iran [M.Phil. thesis]: Imperial College, London University, 273 p. Nankeli, H., 2011, Slip rate of the Kazerun fault and Main Recent fault (Zagros, Iran) from 3D mechanical modeling: Asian Earth Sciences, v. 41, p. 89–98. Ni, J., and Barazangi, M., 1986, Seismotectonics of Zagros continental collision zone and a comparison with the Himalayas: Journal of Geophysical Research, v. 91, no. B8, p. 8205–8218. Niemi, T., 2008, Historical earthquake catalogue and archaeological data, avoiding circular reasoning: Seismological Research Letters, v. 79, p. 289. Nilforoushan, F., Masson, F., Vernant, P., Vigny, C., Martinod, J., Abbassi, M., Nankali, H., Hatzfeld, D., Bayer, R., Tavakoli, F., Ashtiani, A., Doerflinger, E., Daignieres, M., Collard, P., and Chery, J., 2003, GPS network monitors the Arabia-Eurasia collision deformation in Iran: Journal of Geodesy, v. 77, p. 411–422. O’Kane, B., 1987, Timurid Architecture in Khurasan: Costa Mesa, California, Mazda Publishers, 418 p.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084761/spe525-05.pdf by University of Wisconsin Madison user

211

O’Kane, B., 1994, Iran and Central Asia, in Frishman, M., and Khan, H., eds., The Mosque: History, Architecture, Development and Regional Diversity: New York, Thames and Hudson, p. 118–139. Pérez-Lopez, R., Grutzner, C., Lario, J., Reicherter, K., and Silva, P.G., eds., 2009, Archaeosiemology and Palaeoseismology in the Alpine-Himalayan Collision Zone, First INQUA-IGSP-567 International Workshop on Earthquake Archaeology and Palaeoseismicity, 7–13 September 2009, Abstract Volume: Baelo Claudia, Cadiz, Spain, 179 p., ISBN: 978-84-7484-217-3, https://ees.kuleuven.be/igcp567/activities/bc2009/images/Baelo2009 -abstractvolume.pdf. Peyret, M., Rolandone, F., Domingues, S., Djamour, Y., and Meyer, B., 2008, Source model for the Mw 6.1, 31 March 2006, Chalan-Chulan earthquake (Iran) from InSAR: Terra Nova, v. 20, no. 2, p. 126–133, doi:10.1111/ j.1365-3121.2008.00797.x. Pirniā, M.K., 1991, Domes in the Iranian Architecture: Athar, Iranian Cultural Heritage Organization, no. 20, 152 p. [in Persian]. Pirniā, M.K., 2003, Sabk-shenāsi-ye Me’māri-ye Irāni, 2003 [Typology of the Iranian Architecture], Golām Hossein Me’māriān, ed.: Tehran, Nashr-e Pazhuhandeh-Nashr-e Me’mār Publisher, 1382, 371 p. [in Persian]. Pope, A.U., 1934, The historic significance of stucco decoration in Persian architecture: The Art Bulletin, College Art Association, v. 16, no. 4, p. 321–332, http://www.jstor.org/stable/3045552. Pope, A.U., 1938, A Survey of Persian Art from Prehistoric Times to the Present: Oxford, UK, Oxford University Press, 6 volumes. Reprinted by Mazdā Publishers, California, 16 volumes, 3816 p. Pope, A.U., 1965, Persian Architecture, The Triumph of Form and Color: New York, George Braziller, 288 p. Pope, A.U., 1997, A Survey of Persian Art from Prehistoric Times to the Present: Tehran, Sorush Publisher, v. III, 1014 p. [in Persian]. Psycharis, I.N., Papastamatiou, D.Y., and Alexandris, A.P., 2000, Parametric investigation of the stability of classical columns under harmonic and earthquake excitations: Earthquake Engineering & Structural Dynamics, v. 29, no. 8, p. 1093–1109, doi:10.1002/1096-9845. Psycharis, I.N., Fragiadakis, M., and Stefanou, I., 2013, Seismic reliability assessment of classical columns subjected to near-fault ground motions: Earthquake Engineering & Structural Dynamics, v. 42, no. 14, p. 2061–2079. Qazvini, M., Hamdollāh, 1330, Tatikh-e-Gozideh, 730/1330, ‘Abdol Hossain Navā’i, ed., 1387/2008: Tehran, Amir Kabir Publisher, 5th ed., 816 p. [in Persian]. Also Tehran, Amir Kabir Publisher, 1339/1960, 816 p. [in Persian]. Qazvini, M., Hamdollāh, 1340, The Geographical Part of the Nuzhat al-Qulub [Pleasure of the Hearts; 1340 AD], Composed by Hamd-Allah Mustawfi of Qazwin in 740 (1340), Le Strange, G., ed., 1319, E.J.W. Gibb Memorial Series 23: London, E.J. Berill Imprimerimerie Orientale: London, Leyden, Luzac & Co., 1919, 322 p. Also Mustawfi of Qazvin, The Tārikh-I Guzida of Hamd’ullāh of Qazwin: E.J.W. Gibb Memorial Series 14: London, E.J. Berill Imprimerimerie Orientale: London, Leyden, Luzac & Co., 1910, Parts I and II. Quatremere, E.M., 1836, Histoire des Mongols de la Perse Écrite en Persan par Rashid-Eldin, Publiée, Traduite en Français, Accompagnée de Notes et d’un Mémoire sur la vie les Ouvrages de l’Auteur: Paris, Imprimé Royale, p. cxii–cxiii. Quchāni, ‘Abdollāh, 1988, Bāzshenāsi-ye yek masjed-e qadimi dar Qeshm be estenād-e katibeh-yeh chubi mote’aleq be qarn-e hastom-e hejri, 1988 [Reconnaissance of an Old Mosque in Qeshm based on an Eighth Century H. Wooden Inscription]: Tehran, Majaleh Bāstānshenāsi va Tārikh, v. 3, no. 1, p. 22, Autum and Winter, 1367 [in Persian]. Rahbar, M., and Alibaigi, S., 2009, The hunt for Laodicea: A Greek temple in Nahavand, Iran, in Scarre, C., ed., Antiquity: A Review of World Archaeology, Antiquity Project Gallery online, http://www.antiquity.ac.uk/ projgall/alibaigi322/. Rahbar, M., and Alibaigi, S., 1390/2011, Gozaresh pazhohesh-hayeh bastanshenasi be manzor-e makanyabi-ye ma’bad-e Laodisea dar Nahavand [Report of Archaeological Research in Finding the Location of the Laodicea Temple in Nahāvand]: Abhar, Majaleh Payām Bāstān shenāsi, v. 8, no. 15, p. 133–160, 1390/2011, Āzād Eslāmi University Publisher [in Persian]. Rahbar, M., Alibaigi, S., Haerink, E., and Overlaet, B., 2014, In search of the Laodike temple at Laodikeia in Media/Nahavand, Iran: Iranica Antiqua, v. XLIV, p. 301–329, doi:10.2143/LA.49.0.3009244. Reicherer, K., Michetti, A.M., and Silva Barroso, P.G., 2009, Palaeoseismology, historical and prehistorical records of earthquake ground effects

212

Berberian et al.

for seismic hazard assessment, in Reicherer, K., Michetti, A.M., and Silva Barroso, P.G., eds., Palaeoseismology; Historical and Prehistorical Records of Earthquake Ground Effects for Seismic Hazard Assessment: Geological Society, London, Special Publication 316, p. 1–10, doi:10.1144/SP316.1. Rezvāni, ‘A.A., 1997, Pishineh-ye Tārikhi-ye Shahr-e Borujerd, 1376/1997 [History of the Borujerd City], in Proceedings of the Bam Symposium on Architecture and Urban Planning, Arg-e Bam, Kerman: Tehran, Iranian Cultural Heritage Organization, no. 92, v. 5, p. 497–503, 1376 [in Persian]. Rucker, J.D., and Niemi, T.M., 2010, Historical earthquake catalogues and archaeological data; Achieving synthesis without circular reasoning, in Sintubin, M., Stewart, I.S., Niemi, T.M., and Altunel, E., eds., Ancient Earthquakes: Geological Society of America Special Paper 471, p. 97–106, doi:10.1130/2010.247(09). Sani’al-Dauleh, Mohammad Hasan Khān Etemād al-Saltaneh, 1880–1882, Montazam-e Nāseri [Chronicles of the Naser al-Din Shah Period], 3 volumes, 1298–1300/1880–2: Tehran, Lithographic Press, Mohammad Esmā’il Rezvāni Publication, 1363–7/1984–8 [in Persian]. Sani’al-Dauleh, (‘Etemad al-Saltaneh), Mohammad Hassan Khan, 1881, Mer’āt al Boldān-e Naseri [Mirrors of the Cities; a Gazetteer of the Cities], 4 volumes (1293–8/1876–1881): Tehran, Lithographic Press, v. IV, 1296/1879. Also, Parto Nuri ‘Alā and Mohammad ‘Ali Sepānlu, eds.: Tehran, Afsar Publisher, 1364/1985 [in Persian]. Scholz, C., 1982, Scaling laws for large earthquakes, consequences for physical models: Bulletin of the Seismological Society of America, v. 72, p. 1–14. Sella, G.F., Dixon, T.H., and Mao, A., 2002, REVEL, a model for recent plate velocities from space geodesy: Journal of Geophysical Research, v. 107, no. B4, p. 11–32. Shirokova, E.I., 1967, Stresses effective in earthquake foci in the Caucasus and adjacent districts: Bulletin of the Academy of the USSR: Geophysical Series, v. 10, p. 809–815 [translated in English]. Silva, P.G., Sintubin, M., and Reicherter, K., eds., 2011, Earthquake archaeology and paleoseismology: Quaternary International, v. 242, no. 1, p. 1–258. Sintubin, M., 2011, Archaeoseismology: Past, present and future: Quaternary International, v. 242, no. 1, p. 4–10. Sintubin, M., Stewart, I.S., Niemi, T.M., and Altunel, E., eds., 2010, Ancient Earthquakes: Geological Society of America Special Paper 471, 279 p. Statistical Center of Iran (SCI), 1956–2013, The National Iranian Census: Tehran, The Statistical Center of Iran, sci.org.ir (accessed December 2013). Tābandeh, Haji Soltān Hossain, 1969, Tārikh va Joghrāfiya-ye Gonābād [The History and Geography of Gonābād]: Tehrān, University of Tehrān Publisher, 1348/1969 [in Persian]. Tabari, Abu Ja’far Muhammad ibn Jarir abu al-, 915, Tārikh al-Rusul wa’lMuluk [The History of Prophets and Kings], or Tārikh al-Tabari/Tārikh-e Tabari [The History of al-Tabari]: Dār al-Ma’ārif, Cairo, 1960–69, 10 volumes. Persian translation by ‘Abolqāsem Pāyandeh: Tehrān, Bonyad Farhang Publisher, 1354/1975. English translation by C.E. Bosworth, 1999, The History of Al-Tabari, An Annotated Translation: Vol. V—The Sasanids, the Byzantines, the Lakhmidds, and Yemen: Albany, New York, State University of New York Press, 458 p. Also English translation and annotation by Franz Rosenthal, 1989, The History of Al-Tabari Tarikh alRusul Wal-Muluk: Vol. I—General Introduction and From the Creation to the Flood: Albany, New York, State University of New York Press, 458 p. Talebian, M., and Jackson, J., 2002, Offset on the Main Recent fault of NW Iran and implications on the late Cenozoic tectonics of the Arabia-Eurasia collision zone: Geophysical Journal International, v. 150, p. 422–439. Tatar, M., Hatzfeld, D., Martinod, J., Walpersdorf, A., Ghafory-Ashtiany, M., and Chéry, J., 2002, The present-day deformation of the central Zagros from GPS measurements: Geophysical Research Letters, v. 29, no. 19, p. 33-1–33-4, doi:10.1029/2002GL015159.

Tchalenko, J.S., and Braud, J., 1974, Seismicity and structure of the Zagros (Iran), the Main, Recent fault between 33 and 35 degrees N: Philosophical Transactions of the Royal Society of London, v. 277, p. 1–25. Vernant, P., Nilforoushan, F., Hatzfeld, D., Abbassi, M., Vigny, C., Masson, F., Nankali, H., Martinod, J., Ashtiani, A., Bayer, R., Tavakoli, F., and Chéry, J., 2004, Present-day crustal deformation and plate kinematics in the Middle East constrained by GPS measurements in Iran and northern Oman: Geophysical Journal International, v. 157, p. 381–398. Vosoughifar, H.R., 2007, Evaluating retrofitting process for Imam (Soltani) mosque monument after Silakhor Plain earthquake damage (31 March 2006), in Brebbia, C.A., ed., Sixth International Conference on Earthquake Resistance Engineering Structures VI: Southampton, UK, WIT Transactions on The Built Environment, Wit Press, v. 93, p. 387–397, http://www .witpress.com/Secure/elibrary/papers/ERES07/ERES07037FU1.pdf. Vosoughifar, H.R., and Davari, V., 2009, Fracture analysis of retrofitted monuments out of no proper stress distribution at earthquake time, in Mazzolani, ed., Protection of Historical Buildings, PROHITECH-09: London, Taylor and Francis Group, p. 1199–1203. Walpersdorf, A., Hatzfeld, D., Nankali, H.R., Tavakoli, F., Nilforoushan, F., Tatar, M., Vernant, P., Chery, J., and Masson, F., 2006, Difference in the GPS deformation pattern of North and Central Zagros (Iran): Geophysical Journal International, v. 167, p. 1077–1088. Wells, D.L., and Coppersmith, K.J., 1994, New empirical relationships among magnitude, rupture length, rupture width, rupture area and surface displacement: Bulletin of the Seismological Society of America, v. 84, p. 1940–1959. Wikipedia, 2016, List of Paintings and Plots by Pascal Coste and Eugène Flandin: http://en.wikipedia.org/wiki/List_of_paintings_and_plots_by_Pascal _Coste_and_Eug%C3%A8ne_Flandin (accessed June 2015). Wilber, D.N., 1955, The Architecture of Islamic Iran, the Il-Khanid Period: Princeton Monographs in Art and Archaeology XXIX, Oriental Studies XVII, Department of Art and Archaeology and Oriental Languages and Literatures: Princeton, New Jersey, Princeton University Press. Also Persian trans. ‘Abdollah Faryar: Tehran, Bongāh Tarjomeh va Nashr Ketāb Publisher, no. 256, Majmu’eh Ma’aref ‘Omumi-36, 1346/1965, 228 p. Wilkinson, C.K., 1986, Nishapur; Some Early Islamic Buildings and Their Decoration: New York, The Metropolitan Museum of Art, 328 p. Wulff, H.E., 1966, The Traditional Crafts of Persia: Cambridge, Massachusetts, Massachusetts Institute of Technology Press, 404 p. Yāqut Hamavi, ‘Abdollah, 1225, Mo’jam al-Boldān, 623/1225 [Geographic Dictionary]: Beirut, DārSāder al-Tabā’aé val-Nashriya, 1338/1920, p. 404. Mimeograph, Tehran. Also Leipzig, ed. Wustenfeld, 1866–1873. Young, T.C., Jr., 1968, Survey of excavations, Godin Tepe, Iran: Journal of British Institute of Persian Study, London, v. VI, p. 160–161. Young, T.C., Jr., 1969, Excavation at Godin Tepe, First Progress Report: Toronto, Royal Ontario Museum, Art & Archaeology, Occasional Papers 17, 51 p. Young, T.C., Jr., and Levine, L.D., 1974, Excavations of the Godin Project, Second Progress Report: Toronto, Royal Ontario Museum, Art & Archaeology, Occasional Papers 26, 167 p. Zakaryā Qazvini [Zakaryā ebn Mohammad ebn Mahmud al-Makamuni], 1275, Āthār al-Belād va Akhbār al-’Ebād, 674/1275 [Monuments of Places and History of Gods], ‘Abdol Rahmān Sarafkandi, tr.: Tehran, Mo’assesseh ‘Elmi Andisheh Javān Publisher, 1366/1987, p. 95 [in Persian]. Also, Kitāb Āthār al-Bilād wa Akhbār al-Ibād: Gottingen, Germany, ed. Wustenfeld, H.F., 1848.

MANUSCRIPT ACCEPTED BY THE SOCIETY 15 JUNE 2016 MANUSCRIPT PUBLISHED ONLINE 23 SEPTEMBER 2016

Printed in the USA

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084761/spe525-05.pdf by University of Wisconsin Madison user

The Geological Society of America Special Paper 525

Kinematics of the Great Kavir fault inferred from a structural analysis of the Pees Kuh Complex, Jandaq area, central Iran Sasan Bagheri Razieh Madhanifard Foruzan Zahabi Department of Geology, University of Sistan and Baluchestan, Zahedan, Iran

ABSTRACT A significant part of the convergence between the Iranian and Arabian plates since the late Cenozoic has been accommodated by several strike-slip faults, especially in the eastern and central areas of Iran. The Great Kavir fault is one of the cases in which there is little consensus regarding its kinematics, mechanism of development, and tectonic history, mainly due to a lack of detailed studies. Field and satellite image studies of the Pees Kuh Complex, a well-preserved Cenozoic structure that developed upon the Great Kavir fault near Jandaq, in central Iran, suggest a virtually perfect, positive-flower structure. It is argued that the Pees Kuh Complex is the result of a combination of both left strike-slip and reverse dip-slip displacements on the Great Kavir fault. The main structural elements comprising this flower structure are as follows: a Paleogene sedimentary assemblage, composed of an array of thrust faults with NW to EW trends, thrusted upon the Great Kavir block; a few reverse faults with N to NW dips at the southern side of the Great Kavir fault; several synthetic en echelon faults; and a number of antithetic NW-trending en echelon faults. In addition, left-stepping en echelon folds with NW-trending axial planes are recognizable. The Pees Kuh Complex shows a thrust sequence of an upward-verging antiform structure including overturned folds formed of middle to late Eocene marl and sandstone beds. These sheets have a nearly vertical position at their roots, mainly confined to the Great Kavir fault, which changes to a horizontal position further along the fault. Because the thrusts transported middle–late Eocene rocks atop Oligocene–Miocene red beds and are, in turn, covered by Pliocene-age continental beds, the age of the Pees Kuh Complex is inferred to be younger than the Miocene. Considerable leftlateral displacement of the Great Kavir fault in the Jandaq area is confirmed by geometrically measured counterclockwise rotation of ~20° of the faulted blocks around approximately vertical axes relative to the Great Kavir fault in the Godar-e-Siah area. This study, in addition to other previous lithological evidence gathered from the Jandaq area, demonstrates deformation of the Pees Kuh Complex as a reactivation of an older regional fracture, such as a suture zone, as the Paleogene sedimentary rocks were subjected to a different stress field in late Cenozoic times.

Bagheri, S., Madhanifard, R., and Zahabi, F., 2016, Kinematics of the Great Kavir fault inferred from a structural analysis of the Pees Kuh Complex, Jandaq area, central Iran, in Sorkhabi, R., ed., Tectonic Evolution, Collision, and Seismicity of Southwest Asia: In Honor of Manuel Berberian’s Forty-Five Years of Research Contributions: Geological Society of America Special Paper 525, p. 213–227, doi:10.1130/2016.2525(06). © 2016 The Geological Society of America. All rights reserved. For permission to copy, contact [email protected].

213

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084785/spe525-06.pdf by University of Wisconsin Madison user

214

Bagheri et al.

INTRODUCTION Strike-slip faults are classified either as transform faults, which cut the lithosphere as plate boundaries, or as transcurrent faults within the continental crust (Allen, 1965; Wilcox et al., 1973; Sylvester, 1988). The majority of these faults are generally vertical, and they accommodate horizontal shear within the crust and also oblique displacement associated with an assemblage of related structures and physiographic features (ChristieBlick and Biddle, 1985). These geological characteristics could be a useful means for understanding the mechanics and kinematics of strike-slip faults (Garfunkel, 1974; Woodcock, 1986; McClay and Bonora, 2001; Cunningham and Mann, 2007), as well as possibly having some uses in other fields, such as

petroleum geology (e.g., Harding and Lowell, 1979). Recently, attention has also been given to the role of reactivation of old suture and crush zones as one of the causes for the formation of the strike-slip faults (e.g., Bailey et al., 2000; Tikoff et al., 2001; Taylor et al., 2003). The 700-km-long Great Kavir fault, which is the western continuation of the Doruneh fault in central Iran (Stöcklin, 1974), is considered to be a major structure that transects the region with a direction of approximately N50E, from the Afghan border in the east to Nain in central Kavir (Fig. 1). This lineament, indeed, separates the central-east Iranian microcontinent at its southeastern side from several tectonic slivers, including the Great Kavir block to the northwest. The Great Kavir fault must have had an important role in accommodating a large amount of strain in

Figure 1. Map of central Iran with the most important tectono-stratigraphic terranes of central Iran and adjacent tectonic units and some of the important faults; oceanic sutures are shown in the inset figure. The necessary information used for separating the tectonic units was extracted from Stöcklin (1968), Bagheri and Stampfli (2008), Sahandi and Soheili (2011), and Buchs et al. (2013).

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084785/spe525-06.pdf by University of Wisconsin Madison user

Kinematics of the Great Kavir fault inferred from analysis of the Pees Kuh Complex, Jandaq area, Iran central Iran during the Neogene (e.g., Walker and Jackson, 2004; Meyer and Le Dortz, 2007). However, there have been little published data and consensus regarding the kinematics, displacement rate, and convergence mode of the Great Kavir fault. Almost all previous efforts were focused on its neotectonic activities, and in fact, the majority of existing data on this structure are seismological (e.g., McKenzie, 1972; Nowroozi, 1972; Berberian, 1976; Fattahi et al., 2007) or morphotectonic (Wellman, 1966; Tchalenko et al., 1973), which view the Doruneh fault as an active wrench fault with considerable left-lateral and vertical slips. The morphotectonic evidence is based on the displacement of drainage pathways. With most of the inferences coming from satelliteimage interpretation, there is some evidence for left-lateral slip movement on the Great Kavir fault (Walker and Jackson, 2004; Allen et al., 2011). Information obtained directly from geological field observations and mapping is scarce. Limited paleomagnetic data suggest left-lateral displacement of the Great Kavir fault inferred from the rotation of Neogene blocks positioned on both sides of the fault (Mattei et al., 2012). Moreover, Javadi et al. (2013) recently argued for a complex history of the Doruneh fault and changes in its kinematics from a right-lateral to a left-lateral slip during the Pliocene. On the contrary, Nozaem et al. (2013) have shown that the tectonics of Kuh-e-Sarhangi, a magmaticmetamorphic complex situated at the southern margin of the Doruneh fault near Kashmar, demonstrate recent right-lateral slip on the Doruneh fault. While most parts of the Great Kavir fault are covered by Neogene sedimentary rocks with low topography, the Pees Kuh Complex formed on the fault near Jandaq is one of the rare outcrops showing perfect geological contacts (“Pees” in Farsi means bald, and this appellation was given to the area due to lack of vegetative cover). Accordingly, studies of the Pees Kuh Complex not only shed light on its origin but also reveal the kinematics of the Great Kavir fault in the Jandaq area. Moreover, the Great Kavir fault may be the location of a previous regional fracture, such as a suture zone, that was reactivated in the form of its present strike-slip fault. This hypothesis will be tested in this research paper, and pursuant to these aims, we will try to present a tectonic model regarding the structural evolution of the Pees Kuh Complex. In this paper, we characterize the geometry of the Pees Kuh Complex by orientation analysis of the deformation structures. Stretching lineation was used to determine the orientation of the principal strain axes, and mesoscale fault-slip data were recorded on a regional scale. Fault size and attitude, striae orientation, sense of slip, and polyphase slip and its chronology were measured and determined in the field. Riedel shear, steps on the fault surface, and fibers were used for slip-sense determination. We paid special attention to the relationship between faulting and folding; whereas Eocene–Miocene faulting is commonly associated with folding, post-Miocene or late Cenozoic faulting is not.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084785/spe525-06.pdf by University of Wisconsin Madison user

215

GEOLOGICAL SETTING The Pees Kuh Complex crops out in a 30-km-wide range that extends along the Great Kavir fault in the Jandaq area in the heart of the central Iranian desert (Fig. 2A). This rock complex on the Great Kavir fault includes late Paleogene–Neogene sedimentary rocks that are dissected into tectonic slices and sheets with uncertain relations. The Great Kavir fault is the northwestern boundary of the Central-East Iranian microcontinent, a triangular continental realm that includes vast areas of eastern and central Iran in the Iranian Plateau (Takin, 1972). Two different types of outcrops developed along the Great Kavir fault in the Jandaq area (Fig. 2B): a thick pile of Tertiary sedimentary rocks on the northwestern side of the fault with a few limited outcrops of Early Cambrian mylonitic granite; and a thick succession of Jurassic to Paleocene platform-type sedimentary rocks (Aistov et al., 1984) upon a Variscan–Cimmerian metamorphic basement on the southeastern side of the fault (Bagheri and Stampfli, 2008). Pees Kuh Complex The Pees Kuh Complex is a sedimentary stack of middle Eocene to Oligocene deposits (Fig. 3A). The lower part is composed of a thick pile of siliciclastic sediments, marl, and a few basaltic flow layers, which have been distributed over the entire southeastern block of the Great Kavir fault in the Jandaq area (Aistov et al., 1984). The detrital rocks in some portions show turbiditic sequences similar to flysch-like deposits, and its volcanic flows yielded a K/Ar age of 33 Ma (Sharkovski et al., 1984), corresponding to the late Eocene and early Oligocene Epochs. However, in the upper part, carbonates gradually appear among the marls, with some colored sandstones and gypsum beds that extend through whole upper sequence (Babakhani, 1987). This upper part was previously represented as the “Diapir Formation” (Mohafez, 1963). The Oligocene beds belonging to the Pees Kuh Complex are directly and unconformably transgressed with a basal conglomerate on the Godar-e-Siah complex. It is composed of “Variscan” metamorphic rocks, as well as deformed Upper Cretaceous limestone and Lower Eocene volcanic rocks, which indicates a phase of deformation and uplifting during the early Cenozoic. This tectonic event is comparable with the phase that occurred in other parts of the Great Kavir basin (Jackson et al., 1990), and it might be equivalent to the third deformational phase of Rahmati-Ilkhchi et al. (2010). On the basis of the general stratigraphic scheme of central Iran (Stöcklin, 1968), the lower part of the Pis Kuh Formation is almost equivalent to the syntectonic deposits of the Dareh Anjir conglomerate on the Khur platform (Aistov et al., 1984) and also comparable to the Kerman Conglomerate distributed in the southern part of the study area. Subsequent development of detrital sediments with interbedded middle Eocene nummulitic limestone in the Pis Kuh Formation suggests a new stage of marine transgression. At the same time, the deposition of a thick pile

216

Bagheri et al.

Figure 2. (A) Tectono-stratigraphic unit distribution in a map view of the Jandaq area. (B) Tectono-stratigraphic unit situation in a vertical section demonstrating the tectonic relationships between the Pees Kuh Complex, the main suture zones, and adjacent units. Ai—Airekan mylonitic granite; Vs—Variscan accretionary complex; Ci—Cimmerian accretionary complex; Yz—Yazd block; Pkc—Pees Kuh Complex; Nk—Nakhlak Triassic succession; Kp—Khur platform; No—Nain ophiolite, including Ashin-Zavar ophiolitic mélange; Af—Akhoreh flysch; Pa—Pliocene–Pleistocene continental deposits; RF—Red Formation, including Oligocene to Miocene red beds (Lower Red and Upper Red Formations); QF—Qom Formation.

of flysch-type beds, in the form of the Sahlab Formation, north of Anarak along the Great Kavir fault (Sharkovski et al., 1984), may indicate fault reactivation. We consider that the Great Kavir fault commenced in the Late Jurassic; it was reactivated in the late Cenozoic after the ocean closure/suturing in the Eocene. The huge pile of the Sahlab deposits is observed only along the Great Kavir fault in the Anarak area, and their thickness decreases considerably far from the Great Kavir fault. This sedimentary facies, but with a greater abundance of evaporative and volcanic rocks, is distributed through the entire margin of the Great Kavir block. It has been interpreted as deposition in a horst-and-graben basinal pattern (Jackson et al., 1990). At the end of the Eocene, marine sedimentation was terminated by a regional regression, which was related to an intense phase of faulting, tilting, and regional uplifting that affected all of central and eastern Iran (Tirrul et al., 1983; Jackson et al., 1990). This deformational phase is surprisingly absent in the Torud area, less than 100 km to the north of the central Great Kavir desert (Rahmati-Ilkhchi et al., 2010). Therefore, the upper part of the Pees Kuh Complex has been unconformably roofed by the red terrigenous Oligocene beds. A sharp angular unconformity reported in the Rig-e-Jen area

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084785/spe525-06.pdf by University of Wisconsin Madison user

to the northeast of Nakhlak (Aistov et al., 1984) is also confirmed by our recent observations. The northwestern basement of the Pees Kuh Complex is the Qom and Upper Red Formations distributed along the edge of the Great Kavir block. The northern outline of the Pees Kuh Complex is confined by a north-verging thrust fault against the aforementioned formations (Babakhani, 1987). The next significant unconformity around Pees Kuh is between the Pliocene conglomerate and the Oligocene–Miocene beds. These highly unconsolidated Pliocene layers overlie the main northern thrust. STRUCTURAL FRAMEWORK OF THE PEES KUH COMPLEX According to our new observations, the Pees Kuh Complex developed similar to a compressed and pushed-up ridge in which the sedimentary beds were tectonically repeated and accumulated along the Great Kavir fault (Fig. 3B). We think there may not be any stratigraphic relation between the two structurally lower and upper parts of the so-called Pis Kuh Formation (Perfilyev et al., 1978), and these parts may have come in contact with each other during the Great Kavir fault movements. Such an inference

Kinematics of the Great Kavir fault inferred from analysis of the Pees Kuh Complex, Jandaq area, Iran

Figure 3. (A) Geological map of the Jandaq area modified from the 1:100,000 geological maps of Jandaq (Susov et al., 1984) and Chupanan (Perfilyev et la., 1978) with main focus on the Pees Kuh Complex; main fold, fault, and structural lineaments are marked. The diagrams (beach ball) represent the fault exposures and their orientations at selected localities, given for the fault with proper surfaces and slickensides, which yields an age younger than late Eocene. The median has been considered for the numbers of slickensides. Diagrams are lower-hemisphere and equal-area projections used for folds. (B) Cross-section view of the Pees Kuh Complex along the A–B trend marked on the geological map in A. (C) Conceptual diagram of palmtree structure in left-lateral simple shear derived after Lowell (1972).

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084785/spe525-06.pdf by University of Wisconsin Madison user

217

218

Bagheri et al.

was deduced by the absence of any clear stratigraphic contact between the aforementioned parts, as well as several pronounced differences between their lithology and depositional environment. Moreover, the overlapping condition present in the north of the complex reveals the age of the Pees Kuh Complex to most likely be middle Miocene to Pliocene, because the Eocene Pees Kuh beds are overthrust by the early Miocene red beds. Like all types of tectonic structures that form in relation to strike-slip faults, our observations in the Jandaq area show that the Pees Kuh Complex is composed of the following structural elements: strike-slip faults, thrusts, and folds.

with an axial plane parallel to the thrust further confirms the presence of thrusts in the study area (Figs. 4C and 4G). In this section of the Pees Kuh Complex, where the presence of thrusts is more recognizable than in other parts of the complex, sandstone is dispersed in a marl matrix from the Upper Eocene strata, presenting a mixture with an appearance generally similar to a dismembered, or tectonic mélange unit (see Raymond, 1984). The time span considered for thrust creation is most likely post-Oligocene, early Miocene, and pre-Pliocene, as the main thrust of the northern Pees Kuh is concealed unconformably by conglomerate and pebble gravels of Pliocene alluvial deposits.

Strike-Slip Faults

Folds

Generally, two arrays of strike-slip faults in the Pees Kuh Complex can be discussed on the basis of their trends with respect to the Pees Kuh principal displacement zone: NE-trending and NW-trending faults. The NE-trending faults, which are parallel to the main direction of the Pees Kuh outcrop, have left-lateral strike-slip movement combined with some reverse dip-slip component (Figs. 4A, 4B, and 4B-1). These faults are part of the Great Kavir fault system that have N40E strike with a mostly southeastward dip, exposed in the northeast area of Pees Kuh, and sharply separating two different facies of Eocene sedimentary rocks. They may have played a significant role in the deformation of the Jandaq area for the simple reason that all the northwest structural trends become parallel to the Great Kavir fault when approaching the fault (Fig. 4H). The en echelon, NW-trending faults are mainly recognized by their short length, left-stepping, right-lateral slip, and main development in a more rigid part of the complex, in which the sandy to conglomeratic unit of the Pees Kuh Formation can be found (Fig. 4F).

One prominent feature of a number of strike-slip faults is the occurrence of en echelon folds on and adjacent to the principal displacement zone (Christie-Blick and Biddle, 1985; Sylvester, 1988). The folds associated with the Pees Kuh Complex typically appear in two groups, as if they might share the same origin (Figs. 4C and 4G). The first group is the product of thrusting related to the movements of the thrust sheets that originated in the axial part of the Great Kavir fault and traveled away from the strike of the fault. The probable mechanism of this fold generation was fault-propagation folding (Fig. 4G). The folds that are associated with thrusts are often overturned or recumbent synclines and anticlines formed on the northern slope of the Pees

Thrusts Among the tectonic structures that have been clearly identified in the Pees Kuh Complex, there are various thrust faults that caused the dissection and repetition of the Eocene units and the transposition of these units upon the Oligocene and Miocene sequences (Fig. 4D). On the northern slopes of the Pees Kuh Complex, severely crushed masses, including brittle sandstone and other less calcareous rocks, are observed, in which the fault breccia are cemented in a marl and gypsiferous matrix (Figs. 4E, 4E1, and 4E2). Processes such as solution, cementation, and vein injection have occurred intricately in some stages. The abundant presence of such cohesive fault rocks in the cliffs and reliefs is especially noticeable at the boundary of the Eocene and the younger successions, and these are represented as fault scarps. Additionally, abrupt changes in the lithology and layer slopes beside overturned folds and intricate structures are all indicators of subhorizontal faults or thrusts that were reclined during later progressive shortening. The emergence of large overturned folds

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084785/spe525-06.pdf by University of Wisconsin Madison user

Figure 4. Photographs showing various structural elements that indicate the Pees Kuh Complex is a positive flower structure: (A) Landscape of the Great Kavir fault as the principal displacement zone of the main left-stepping strike-slip fault of the Jandaq area. (B) Field view of the cross section of the Great Kavir fault in the location 33°59′24″N, 54°12′46″E. (B-1) Cyclograph showing lower-hemisphere equal-area projections of the faults situated in the main principal displacement zone of the Pees Kuh Complex. The arrows are indicators of the slip vectors measured from related slickensides, which appear as striations or fibers. The red small circles represent the principal stress axes. (C) Southeastward-verging overturned, sheared fold with a thrust plane in its axial plane position. The location is 33°59′47″N, 54°15′18″E. (D) Landscape of the thrust sheets and the klippe of the Pees Kuh Peak. (D-1) Close-up photo of the Pees Kuh klippe showing the contrast between the dip of the beds on both sides of the thrust plane. This large overturned fold rose over the main south-dipping imbricate thrust system during northward squeezing and thrusting of the Eocene sandstone and marl in the location 34°02′46″N, 54°19′08″E. (E) Landscape of a tectonic thrust sheet lying on the steeply dipping beds. (E-1) Photograph showing the escarpment of a thrust sheet on the northern slope of Pees Kuh. (E-2) Photograph zooming in on the thrust sheet’s tectonic breccia. (F) An array of the left-stepping antithetic shear faults and their position relative to one of the main faults parallel to the Great Kavir fault in middle Eocene sandstone and marl in location 33°59′32″N, 54°13′47″E. (G) View of the south-dipping thrust and related detachment folds cutting the en echelon folds with NWtrending axial plane, which was probably formed sooner in location 34°02′49″N, 54°21′18″E. (H) Google Earth image of the northeastern area of Jandaq showing a leftward deflection of the NW-trending lineaments and folds where they approach the Great Kavir fault.

Kinematics of the Great Kavir fault inferred from analysis of the Pees Kuh Complex, Jandaq area, Iran

Figure 4.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084785/spe525-06.pdf by University of Wisconsin Madison user

219

220

Bagheri et al.

Kuh Complex between the thrust sheets. Indeed, Pees Kuh Peak is a large-scale northwest-verging overturned fold made of sandstone that appears as a klippe and is positioned on a thrust surface (Figs. 4D and 4D-1). On the contrary, the most outstanding group of folds (the second group) in the Jandaq area is arranged in an en echelon pattern in a northwest axial plane orientation oblique to the principal direction of the main shear. The trend of the crestal traces of these en echelon folds varies from 30° to parallel to the Great Kavir fault (Fig. 3A). The axial planes of some of these folds display a curved direction when they approach the main shear zone, becoming parallel with the fault (Fig. 4H), and may have been rotated at a later time due to the progressive displacement of the Great Kavir fault (for an analog description, see Harding and Lowell, 1979). This structure may also be due to the helical geometry of the folds in three dimensions that flatten upward and twist away from the strike of the fault. Consequently, the depth of erosion is an important factor in the amount of rotation presented (see Sylvester, 1988). The fold array seems to be a left-stepping, en echelon pattern relative to the Great Kavir fault. Accordingly, such an apparent arrangement in the fold pattern may indicate the presence of a noteworthy left-lateral strike-slip displacement associated with the Great Kavir fault. Some of these folds, which have perfect exposure, show a plunge away from the Great Kavir fault (see the fold cyclographs on Fig. 3A). Positive Flower Structures Flower structures often develop at the jogs of stepping strike-slip faults (Harland, 1971; Harding et al., 1983). In the case of converging strike-slip faults, these structures are associated with prominent antiforms and even uplifts, and they are known as positive flower structures (Harding and Lowell, 1979). In the case of diverging strike-slip faults (Harland, 1971), these structures are associated with synforms and even basins, and they are termed negative flower structures (e.g., Harding, 1983, 1985). The occurrence of upward-diverging fault splays is thus

not due to convergence, but to the propagation of faults upward through the sedimentary cover toward a free surface (Wilcox et al., 1973; Sylvester and Smith, 1976; Christie-Blick and Biddle, 1985). Along with and close to the Great Kavir fault in the Jandaq area, the slope of the sedimentary layers is predominantly steep and lessens to the sides farther away from the main fault. As indicated by the cross section (Fig. 3B), the presence of such layers, fault slopes, and overturned folds with two different vergences on both sides of the main fault (e.g., comparing Figs. 4C and 4D-1) suggests that a positive flower structure has developed on the Great Kavir fault. Field and satellite imagery studies of the Pees Kuh Complex in the Jandaq area show two sets of NW- to E-W–trending thrust sequences with northeast to north vergence in the northwestern portion of the Great Kavir fault (Figs. 4D and 4E), and a set of E-W–trending thrust to reverse faults with south vergence in the southeastern section of the main fault (Fig. 4C). On the other hand, two sets of NE and NW to E-W–trending en echelon faults appear in the middle of the complex, intersecting the main NE-trending beds. The position of these two fault groups relative to the principal displacement zone may reveal them to be the synthetic and antithetic faults of the principal displacement zone of the Pees Kuh Complex, which have a mainly reverse dip-slip vector. It may possible, by a continuation of Great Kavir fault displacement, that the blocks surrounded by these faults underwent a degree of counterclockwise rotation, and accordingly, several of the faults were repositioned into a nearly E-W direction. Relating the NW-trending en echelon folds to the previously explained fault assemblage reveals the character and distribution of the structural constituents of the Pees Kuh Complex, which, in a plan view of geometric relations among structures, are very similar to the angular relations between structures that tend to form in left-lateral simple shear in Wilcox’s model (Fig. 5; Wilcox et al., 1973). On the other hand, the Pees Kuh Complex indicates successive imbricate thrust sheets, each of which (including the Eocene marl and sandstone beds) has a nearly vertical position at its root (mainly confined to the Great

Figure 5. Simple tectonic map of the Jandaq area showing the main structural elements and their angular relationships to the Pees Kuh Complex and surrounding areas. Inset figure: Angular relationships between the structures that tend to form under left-lateral simple shear in ideal conditions. This is presented for comparison. Terminology is largely from Wilcox et al. (1973), superimposed on a strain ellipse for overall deformation.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084785/spe525-06.pdf by University of Wisconsin Madison user

Kinematics of the Great Kavir fault inferred from analysis of the Pees Kuh Complex, Jandaq area, Iran Kavir fault; Fig. 4A) while becoming horizontal far from the fault (Fig. 4E). These sheets perfectly demonstrate an antiformlike structure with an upward-diverging imbricated fault system (Fig. 3B). In general, therefore, it seems that the Pees Kuh Complex has a structure similar to the conceptual diagram of the palm-tree structure with left-lateral simple shear proposed by Lowell (1972) (Fig. 3C). BLOCK ROTATION Tectonic rotation of Earth’s crustal blocks around a vertical axis in simple shear upon or near to main strike-slip faults has been reported from some parts of the world (Freund, 1970; Nur et al., 1986). In this study, three well-exposed blocks located in the Godar-e-Siah area (Fig. 6A), on the southeastern side of the Great Kavir fault, adjacent to the fault, were identified. These blocks rotated counterclockwise around vertical axes a few tens of degrees relative to the present strike of the Great Kavir fault. These blocks are parts of the Khur platform basement, and they are confined to the major NE-trending left-lateral faults belonging to the Great Kavir fault system and a few minor NW-trending dextral faults between the main faults. The key bands of the Upper Cretaceous limestone and Eocene volcanic rocks on the map view clearly show displacement of about 1 km along the minor faults (Fig. 6C). Here, block rotation occurred with the help of dextral

antithetic shear faults. It seems that the rotated blocks may, in fact, be crustal slabs detached from a shallow horizontal shear surface underlying the blocks. It must be taken into account, however, that the deformation mechanism considered here is a situation between pervasive, continuous bulk simple shear and block rotation with internal antithetic shear (Sylvester, 1988, p. 1694). This mechanism provides us with clear clues regarding the amount of displacement, which is more than what we observe in the two-dimensional geometry on the satellite image. Moreover, this block rotation clearly confirms several kilometers of left-lateral movement of the Great Kavir main displacement zone since the Oligocene– Miocene. Unfortunately, the absence of paleomagnetic data does not allow us to accurately determine the amount of rotation. However, in this study, we have tried to geometrically measure the amount of rotation using Ron’s method (Fig. 6B; Ron et al., 1984). Accordingly, the minimum amount of left-lateral slip of the Great Kavir fault was measured mainly for the Neogene, the period after suture zone formation. Calculation and Parameters In the Godar-e-Siah region, two sinistral parallel strike-slip faults developed in Neogene sediments, and they are most probably two constituents of the Great Kavir fault system. Satellite

Figure 6. (A) Google Earth image showing the position of three faulted blocks of the Godar-e-Siah Complex relative to the Great Kavir fault. The effect of left-stepping simple shear on counterclockwise rotation of blocks can be inferred. (B) Inset image showing geometric model presenting the quantitative relation between the strike-slip displacement and the rotation of the faulted blocks after Ron et al. (1984). Below, the effect of this displacement has been accompanied by the shortening and lengthening of the district around the Great Kavir fault. Explanation: α—acute angle of rhomboidshaped fault block (the angle between synthetic and antithetic faults); δ—angle of rotation; lo—length of fault before slip; l—length of fault after slip; d—lateral slip on the antithetic fault; w—width of fault block; Lo—width of terrane between the synthetic faults before rotation; L—width of the terrane between the synthetic faults after rotation; So—length of the terrane before rotation; and S—length of the terrane after rotation. (C) Google Earth image showing the right-lateral displacement of a Cretaceous key bed. Same scale as in A.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084785/spe525-06.pdf by University of Wisconsin Madison user

221

222

Bagheri et al.

image studies and field observations confirm a moderate lateral slip (d) of ~1.12 km, and the average width of the fault blocks (w) located between the antithetic faults is ~3 km, and the amount of α is ~70° as indicated in Figure 6. The amount of rotation, based on the geometric method presented by Ron et al. (1984), is calculated to be ~20° as follows: d sinσ = = cot (α + σ) – cot α , w sinα sin(α – σ) α – σ = 90 sin α = cos (90 + σ), d = tan σ , w σ

20°.

The amount of rotation may even be greater than 30° at center of the middle block as illustrated in Figure 6. Such rotation is the result of at least 600 m of left-lateral displacement on the Great Kavir fault, based on calculation of the difference between L0 (length before rotation) and L (length after rotation). It seems that block rotation during left-lateral strike-slip fault movements gave rise the NNW-SSE–trending structures, which rotated counterclockwise, assuming the present NW-SE direction. This block rotation event was certainly one of the main causes of NE elongation along the fault and NW shortening of the Pees Kuh Complex in the Jandaq area (Fig. 6). GREAT KAVIR SUTURE ZONE According to the stratigraphy and/or tectonic history variations on both sides of the Great Kavir fault presented here, the Great Kavir fault has the possibility of being a former suture zone. Prior to the Tertiary, the present location of the Great Kavir fault was most probably on the northern margin of the Central Iranian continent, a part of the Cimmerian block, facing a peri-Tethyan ocean (Fig. 7A; Bagheri, 2007). The southeastern zone of the Great Kavir fault is characterized by a “Variscan” basement affected by a Cimmerian orogenic event(s) and late Triassic granite intrusion (Bagheri and Stampfli, 2008), which is perhaps a part of the Eurasian basement, and it is covered by several thousand meters of Cretaceous–Paleocene marine deposits referred to as the Khur platform (Jackson et al., 1990; Bagheri, 2007; Wilmsen et al., 2015). This platform cannot be found in other parts of central Iran, but it is similar to the sedimentary successions in the region of the Caspian Basin (Brunet et al., 2003) and the North Afghan platform (Brookfield and Hashmat, 2001). This sequence was unconformably overlain by a large volume of Eocene-age, flysch-like, turbiditic sediments, and fine-grained, volcaniclastic deposits (Bagheri, 2007). It seems these deposits were then covered by a thick pile of marl, sandstone, and gypsum. This portion of the Eocene deposits in the Jandaq area is referred to as the Pis Kuh Formation (Perfilyev et al., 1978; Babakhani, 1987).

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084785/spe525-06.pdf by University of Wisconsin Madison user

On the contrary, the northwestern zone of the Great Kavir fault has different characteristics. Most areas of the Great Kavir block are covered by mainly Tertiary detrital beds. Airakan granite appears at the margin of the block several kilometers toward the northeast. It has been attributed to an Early Cambrian intrusion (Bagheri and Stampfli, 2008) and is similar to other Iranian late Proterozoic–early Paleozoic igneous-metamorphic rocks (Hassanzadeh et al., 2008), a characteristic which indicates the possible existence of a Gondwanan basement. Several tectonic slices with an identical basement, a Cretaceous ophiolitic mélange, as well as Cenozoic calc-alkaline magmatic rocks are scattered to the north of the Doruneh fault in the Khorasan province (e.g., Lindenberg et al., 1984; Horton et al., 2008; Alaminia et al., 2013). At the western termination of the Great Kavir fault, the Nain ophiolitic mélange crops out along with a thick pile of Eoceneage Akhoreh flysch (Aistov et al., 1984). Some of the upper beds and tectonic slices along the Great Kavir fault, such as the late Eocene Rig-e-Jen granodiorite north of Nakhlak (Susov et al., 1979), are pierced by acidic intrusions. The Oligocene-age sediments of the Lower Red Formation are similar to molasse deposits, with pronounced angular unconformities to disconformities that overlie the older rocks in most of the central Iranian realm. Evaporative basins developed on both sides of the Great Kavir fault during the Eocene and most of the Tertiary; e.g., the

Figure 7. Model of tectonic history of the district, including the Great Kavir block as the active margin and the Khur platform as the passive margin of a peri-Tethyan ocean (the Sabzevar Ocean is a good case in point), since late Mesozoic time in a plate-tectonic scenario preserved in central Iran. This model follows general principles of Bagheri (2007) and Bagheri and Stampfli (2008). (A) Late Cretaceous–Paleocene time: Initiation of closure of a peri-Tethyan ocean originated probably in the back-arc of the Neotethys Ocean, associated with: (1) collapse of the Khur platform, including 5000–6000 m of sediment deposited in a limited time span of ~100 m.y. on the southern continental margin of the mentioned back-arc basin; (2) northward subduction of the oceanic crust under an unknown continental plate (by “unknown,” we mean that the evidence of a pronounced magmatic arc cannot be seen in the region where we expect to find it; this active margin can be seen today in the form of several displaced continental slivers that are arranged on the north side of the Great Kavir fault); and (3) formation of the Dareh Anjir flysch-like deposit showing a rigorous change in type of sedimentary environment from pelagic to continental. (B) Eocene–Oligocene time: Ultimate stages of ocean closure and continental collision, which was accompanied by large lateral displacement and deposition of the Pees Kuh Formation. The Khur platform gradually changes in an entrapped intraplate continental margin up to the early Neogene due to probable counterclockwise rotation of the Central-East Iranian microcontinent. The Middle Alpine orogeny caused uplifting of the folded Biabanak and Khur basins and widespread calc-alkaline magmatism. (C) Miocene–Pliocene time: Changing of stress regime, and reactivation of the Great Kavir fault in a different sense to a sinistral strike-slip displacement, and uplift of the Pees Kuh Complex in the form of a positive flower structure. (D) Lithostratigraphy explanation. (E) Summary of the timing of the main stages of deposition, magmatic episodes, and deformational events.

Kinematics of the Great Kavir fault inferred from analysis of the Pees Kuh Complex, Jandaq area, Iran

Figure 7.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084785/spe525-06.pdf by University of Wisconsin Madison user

223

224

Bagheri et al.

Great Kavir salt domes in the north (Jackson et al., 1990), and the Ardekan deposits to the south of the Great Kavir fault (Valeh and Haghipour, 1972). Therefore, the Great Kavir fault, speculatively, the Great Kavir suture zone, separates two different zones that are affiliated with two different tectonic provinces—the southeastern zone originating from Eurasia (Bagheri and Stampfli, 2008), and the northwestern zone having a Gondwana origin (e.g., Hassanzadeh et al., 2008; Horton et al., 2008). Such an arrangement is inverted relative to the normal position of Gondwana-derived blocks that have been accreted to the southern margin of Eurasia through almost the entire length of the Paleotethys suture zone (e.g., Sengör and Natal’in, 1996; Stampfli and Borel, 2002). Accordingly, the Great Kavir fault and its eastern continuation, the Doruneh fault, could be the signature of a peri-Tethyan Cretaceous oceanic crust, probably belonging to the Sabzevar Ocean, which was subducted under the northern blocks (Rosseti et al., 2010) such as the Great Kavir block, that finally closed during the Middle Alpine orogenic event (Fig. 7B). A large rightlateral strike-slip displacement accompanied the final stages of the closure of the Sabzevar Ocean during the counterclockwise rotation of the Central-East Iranian microcontinent around a vertical axis (Bagheri, 2007; Bagheri et al., 2011; Javadi et al., 2013). This might have resulted from the India-Eurasia collision (Ghodsi et al., 2016). Since the late Oligocene, and contemporaneous with the main period of crustal thickening in the Iranian Plateau caused by the collision of Arabia-Eurasia as recorded by hinterland exhumation (Mouthereau et al., 2012), and due to changes in the stress regime, the Great Kavir fault seems to have changed its behavior from a suture zone to left-lateral strike-slip displacement (Bagheri et al., 2011). Tectonic Evolution of the Pees Kuh Complex and Adjacent Areas Part of the complicated tectonic history of the Great Kavir suture zone (Fig. 7E) has been recorded in the Pees Kuh Complex, which experienced a near-complete Wilson tectonic cycle from the Late Jurassic to the Eocene–Oligocene, followed by Neogene reactivation (Bagheri, 2007). Since the distribution of the thick Neogene sediments is confined mainly to the northwestern margin of the fault, it may be concluded that the Great Kavir suture zone was reactivated along the Great Kavir fault sometime after the Oligocene. This tectonic movement resulted in the piling of the Pees Kuh Complex buildup (Fig. 7C). The Arabia-Eurasia collision began in the early Neogene period (e.g., Walker and Jackson, 2004; Mouthereau et al., 2012) and was the main stress source for the youngest deformation and structural shaping of the Iranian Plateau during the Late Alpine orogeny (Allen et al., 2004). The main stress vector was moderately exerted in a NNE direction to the NEE-striking Great Kavir suture zone (Vernant et al., 2004). Consequently, this suture zone was crushed and moved in a left-lateral manner with a considerable inverse compressive component, and a huge

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084785/spe525-06.pdf by University of Wisconsin Madison user

buildup of Eocene sediments, which had already accumulated near the Great Kavir fault, was then uplifted with beds overriding one another, forming the present Pees Kuh Complex. The Eocene beds were consequently overthrusted north and southward onto both sides of the Great Kavir fault over the Oligocene beds, and fault-propagation folds as well as large-scale NW-axial-plane-trending en echelon folds were thus formed. The lithological nature and thickness of sediments observed at various locations along the Great Kavir fault suggest that the first group of folds, on the northwestern block of the fault, was well developed and preserved, while the second group, on the southeastern block, are not as clearly visible. A continuation of this new stress regime subsequently changed the NW-trending folds, which had formed very early, and which were either truncated to several half-folds, and crushed the preexisting structures to disjointed parts, or were twisted into a NE direction parallel to the Great Kavir fault. Consequently, the present form of the Pees Kuh Complex is a broken formation/tectonic mélange (Raymond, 1984). This recent event may have been Pliocene in age. Moreover, younger thrusts have formed, located in the deeper parts on both sides of the fault, and far from the uplifted region, with the characteristics of a fan-shaped thrust, imbricate, or sequential system, becoming older farther up, and sloping toward the axial region of the main shear zone. Coeval with this deformation and as slip continued, dextral antithetic strike-slip faults or folds underwent counterclockwise rotation, and some of them have outcropped parallel to the Great Kavir fault. Another part of the strain was accommodated by the margin of the Great Kavir fault, where the exhumed basement has been involved in the rotation of blocks around vertical axes. Several right-lateral antithetic faults surround these blocks and have propagated upward into the Neogene deposits, which helped them to rotate ~20° in a counterclockwise direction. This is clearly evidenced by a left-lateral displacement of at least a few hundred meters along the Great Kavir fault since the Miocene. Our observation is confirmed by new paleomagnetic studies carried out in the Jandaq area (Mattei et al., 2012), which show a 14° counterclockwise rotation in the blocks situated on the southeastern side of the Great Kavir fault. DISCUSSION The geometry and style of the structures associated with the Great Kavir fault depend greatly on several factors operating at different times and different locations along and within the fault zone (for theoretical considerations, see Sylvester, 1988). These include (1) the nature and thickness of the gypsum and marl beds being deformed in the northwestern block, (2) the configuration of the preexisting structures, especially in the southeastern block, (3) the amount of horizontal slip during and after the Eocene (the approximate age of the suturing event), (4) the contribution of the vertical component of slip, and (5) the distribution of the strain rate in the shear zone.

Kinematics of the Great Kavir fault inferred from analysis of the Pees Kuh Complex, Jandaq area, Iran Perhaps the most important factor governing uplift or subsidence along a strike-slip fault is the bending geometry of the fault surface relative to its slip factor (Christie-Blick and Biddle, 1985; Cunningham and Mann, 2007). This situation determines whether convergence or divergence prevails locally along the strike-slip fault. However, the satellite image does not show significant bending of the Great Kavir fault in the Jandaq area when following the path of the main principal displacement zone; therefore, the structural setup and origin of the Pees Kuh Complex in this area are still a matter of debate. At present, our suggestion of a restraining bend and hence a positive flower structure should be considered as only one possibility, and further studies are required to decipher the nature of the structure in the study area (Cunningham and Mann, 2007). It is presumed the bending in the main strand of the Great Kavir fault to west of Jandaq, however, is insufficient to have generated such a long positive flower structure. Typically, the ratio between the thickness of this sort of structure perpendicular to the strike-slip fault and its length along the fault is much greater than that of the Pees Kuh Complex. When we consider the Great Kavir fault on a larger scale than that shown in Figure 3, this bending is not even distinguishable. Second, and more importantly, the restraining bends in a noticeable strike-slip fault can generate compressional strike-slip duplexes (Woodcock and Fischer, 1986). There are no asymmetrically stacked horses of rock bodies in a left-stepped en-echelon pattern on any map view associated with the Great Kavir fault. Hence, it is considered that the abnormal thickness of the Neogene sediments (including evaporate with a plastic deformation manner during compression), accumulated and confined to the fault in this area, might have been squeezed upward in a transpressional stress setting. This is reasonable cause for the uplifting of the Pees Kuh Complex. However, factors such as overstepping of fault segments and/or the presence of a segmented basement at the fault location should also be taken into account. It seems that the Pees Kuh Complex is the result of convergent strike-slip fault tectonics, in which one feature is a horizontal shortening of ~10% based on the ratio of L0-L of the rotated blocks, perpendicular to the fault zone (Fig. 6), or even greater as estimated by the restoration of the folds in close proximity to the Great Kavir fault (Fig. 3B, see the cross section). The other feature caused by fault activity is the position of the rotated blocks south of the Great Kavir fault in Godar-e-Siah (Fig. 5), and unsurprisingly, an almost equal amount of lengthening parallel to the Great Kavir fault has taken place in that area. Thus, the approximate amount of left-lateral displacement of the Great Kavir fault since the Neogene is at least 4.5 km in order to produce the Pees Kuh Complex, with a length of ~30 km, based on the same method considered for its shortening rate. It should be kept in mind, however, that a considerable part of the strain was accommodated in order to distort the rocks as well as rotating the faulted blocks. Therefore, the rate of displacement along the Great Kavir fault must be greater than previously considered.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084785/spe525-06.pdf by University of Wisconsin Madison user

225

CONCLUSIONS The Pees Kuh Complex is a series of pushed-up, Eocene-age sedimentary rocks that appear in the form of a broken formation inside the Neogene beds along the Great Kavir fault in the Jandaq area, central Iran. It indicates a positive flower structure formed during the Miocene (as the thrusts have been unconformably covered by Pliocene sediments). The most significant features of the Pees Kuh Complex are the E-W– to NW-trending, left-stepping en echelon folds, with right-lateral minor faults parallel to them, several northverging overturned folds on various scales associated with concave-up imbricated thrust sequences at the northwestern side of the principal displacement zone, and a few upright folds with a strike axial plane parallel to the main fault, and confined to a north-dipping reverse fault located at the southeastern side of the fault. Consideration of the geometric relationships between these structural constituents in comparison to Sylvester’s (1988) model clearly demonstrates a flower structure associated with the leftlateral strike-slip movements of the Great Kavir fault. The amount and direction of slip on these structures also provide consistency with the Riedel model of left simple shear for a flower structure. In addition, several fault-bounded blocks are identified in the Godar-e-Siah area that have experienced at least 20° of counterclockwise rotation around vertical axes relative to the main strike of the Great Kavir fault. We rely on the positive flower structure and the faulted block rotation as the most striking evidence showing left-lateral strike-slip displacement of the Great Kavir fault since the Miocene. This movement was accompanied by at least 10% shortening normal to the fault. The Pis Kuh Formation was originally deposited on the margin of the Khur platform and was later intersected by the Great Kavir fault and compressed into an uplifted ridge in the present form of the Pees Kuh Complex. The presence of such a large transcurrent fault at the boundary between two main terranes with different tectono-stratigraphic nature in central Iran leads us to suggest that the Great Kavir fault coincides with an old suture zone, a welded belt formed around the Eocene, which has been reactivated in a strike-slip style since the Neogene due to changes in the regional stress regime. Continuation of Arabia-Eurasia convergence gave rise to several deflections and ruptures during the Pliocene, and the younger sediments in the Jandaq area show a continuation of the left-lateral movement of the Great Kavir fault in recent times. ACKNOWLEDGMENTS This work is dedicated with gratitude to Manuel Berberian in recognition of his decades of research and inspirational teaching. We also gratefully acknowledge the Jandaq Municipality and its kind staff for their all their logistic support and accommodations during our fieldwork. We would like to express our appreciation to three reviewers for their corrections and

226

Bagheri et al.

constructive suggestions, and Andrea Parishani for her help and comments. Any error is, however, ours. REFERENCES CITED Aistov, L., Melanikov, B., Krivyakin, B., Morozov, L., and Kiristaev, V., eds., 1984, Geology of Khur Area (Central Iran), Explanatory Text of the Khur Quadrangle Map 1:250,000: Tehran, Geological Survey of Iran, and Moscow, V/O “Technoexport” USSR Ministry of Geology Report TE/No. 20, scale 1:250,000, 132 p. Alaminia, Z., Karimpour, M.H., Homam, S.M., and Finger, F., 2013, The magmatic record in the Arghash region (NE Iran) and tectonic implications: International Journal of Earth Sciences, v. 102, p. 1603–1625, doi:10.1007/s00531-013-0897-1. Allen, C.R., 1965, Transcurrent faults in continental areas: Royal Society of London Philosophical Transactions, v. 258, p. 82–89, doi:10.1098/ rsta.1965.0023. Allen, M., Jackson, J., and Walker, R., 2004, Late Cenozoic reorganization of the Arabia-Eurasia collision and the comparison of short-term and longterm deformation rates: Tectonics, v. 23, p. 1–16. Allen, M.B., Kheirkhah, M., Emami, M.H., and Jones, S.J., 2011, Right-lateral shear across Iran and kinematic change in the Arabia–Eurasia collision zone: Geophysical Journal International, v. 184, p. 555–574, doi:10.1111/ j.1365-246X.2010.04874.x. Babakhani, A.R., 1987, Geological Quadrangle Map of Jandaq, No. H6: Tehran, Geological Survey of Iran, scale: 1:250,000. Bagheri, S., 2007, The Exotic Paleo-Tethys Terrane in Central Iran: New Geological Data from Anarak, Jandaq and Posht-e-Badam Areas [Ph.D. thesis]: Lausanne, Switzerland, University of Lausanne, 223 p. Bagheri, S., and Stampfli, G.M., 2008, The Anarak, Jandaq and Posht-e-Badam metamorphic complexes in central Iran: New geological data, relationships and tectonic implications: Tectonophysics, v. 451, p. 123–155, doi:10.1016/j.tecto.2007.11.047. Bagheri, S., Madhani-Fard, R., Zahabi, F., and Javadi-Zadeh, L., 2011, Kinematics of Doruneh fault in central Iran: Two different behaviors since Eocene time, in Proceedings, 9th Swiss Geoscience Meeting, Structural Geology, Tectonics and Geodynamics Session [abs.]: Zurich, Switzerland, part 1.E.1, p. 68. Bailey, W.R., Holdsworth, R.E., and Swarbrick, R.E., 2000, Kinematic history of a reactivated oceanic suture: Journal of the Geological Society, London, v. 157, p. 1107–1126, doi:10.1144/jgs.157.6.1107. Berberian, M., 1976, Contribution to the Siesmotectonics of Iran (Part II): Geological Survey of Iran Report 39, 518 p. Brookfield, M.E., and Hashmat, A., 2001, The geology and petroleum potential of the North Afghan platform and adjacent areas (northern Afghanistan, with parts of southern Turkmenistan, Uzbekistan and Tajikistan): EarthScience Reviews, v. 55, p. 41–71. Brunet, M., Korotaev, M.V., Ershov, A.V., and Nikishin, A.M., 2003, The South Caspian Basin: A review of its evolution from subsidence modelling: Sedimentary Geology, v. 156, p. 119–148, doi:10.1016/S0037 -0738(02)00285-3. Buchs, D., Bagheri, S., Martin, L., Hermann, J., and Arculus, R., 2013, Paleozoic to Triassic ocean opening and closure preserved in central Iran: Constraints from the geochemistry of meta-igneous rocks of the Anarak area: Lithos, v. 172–173, p. 267–287, doi:10.1016/j.lithos.2013.02.009. Christie-Blick, N., and Biddle, K.T., 1985, Deformation and basin formation along strike-slip faults, in Biddle, K.T., and Christie-Blick, N., eds., Strike-Slip Deformation, Basin Formation, and Sedimentation: Society of Economic Paleontologists and Mineralogists Special Publication 37, p. 1–34, doi:10.2110/pec.85.37.0001. Cunningham, W.D., and Mann, P., 2007, Tectonics of strike-slip restraining and releasing bends, in Cunningham, W.D., and Mann, P., eds., Tectonics of Strike-Slip Restraining and Releasing Bends: Geological Society, London, Special Publication 290, p. 1–12, doi:10.1144/SP290.1. Fattahi, M., Walker, R.T., Khatib, M.M., Dolati, A., and Bahroudi, A., 2007, Slip-rate estimate and past earthquakes on the Doruneh fault, eastern Iran: International Journal of Geophysics, v. 168, p. 691–709, doi:10.1111/ j.1365-246X.2006.03248.x. Freund, R., 1970, Rotation of strike slip faults in Sistan, southeast Iran: The Journal of Geology, v. 78, p. 188–200, doi:10.1086/627500.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084785/spe525-06.pdf by University of Wisconsin Madison user

Garfunkel, Z., 1974, Model for the late Cenozoic tectonic history of the Mojave Desert, California, and for its relation to adjacent regions: Geological Society of America Bulletin, v. 85, p. 1931–1944, doi:10.1130/0016 -7606(1974)852.0.CO;2. Ghodsi, M.R., Boomeri, M., Bagheri, S., Ishiyamab, D., and Corfu F., 2016, Geochemistry, zircon U-Pb age and tectonic constraints on the Bazman Granitoid Complex, southeast Iran: Turkish Journal of Earth Sciences, doi:10.3906/yer-1509-3. Harding, T.P., 1983, Divergent wrench fault and negative flower structure, Andaman Sea, in Bally, A.W., ed., Seismic Expression of Structural Styles: American Association of Petroleum Geologists Studies in Geology Series 15, Volume 3, p. 4.2–1 to 4.2–8. Harding, T.P., 1985, Seismic characteristics and identification of negative flower structures, positive flower structures, and positive structural inversion: American Association of Petroleum Geologists Bulletin, v. 69, p. 582–600. Harding, T.P., and Lowell, J.D., 1979, Structural styles, their plate-tectonic habitats, and hydrocarbon traps in petroleum provinces: American Association of Petroleum Geologists Bulletin, v. 63, p. 1016–1058. Harding, T.P., Gregory, R.F., and Stephens, L.H., 1983, Convergent wrench fault and positive flower structure, Ardmore Basin, Oklahoma, in Bally, A.W., ed., Seismic Expression of Structural Styles: American Association of Petroleum Geologists Studies in Geology, Series 15, Volume 3, p. 4.2–13 to 4.2–17. Harland, W.B., 1971, Tectonic transpression in Caledonian Spitsbergen: Geological Magazine, v. 108, p. 27–42, doi:10.1017/S0016756800050937. Hassanzadeh, J., Stockli, D.F., Horton, B.K., Axen, G.J., Stockli, L.D., Grove, M., Schmitt, A.K., and Walker, J.D., 2008, U-Pb zircon geochronology of late Neoproterozoic–early Cambrian granitoids in Iran: Implications for paleogeography, magmatism, and exhumation history of Iranian basement: Tectonophysics, v. 451, p. 71–96, doi:10.1016/j .tecto.2007.11.062. Horton, B.K., Hassanzadeh, J., Stockli, D.F., Axen, G.J., Gillis, R.J., Guest, B., Amini, A., Fakhari, M.D., Zamanzadeh, S.M., and Grove, M., 2008, Detrital zircon provenance of Neoproterozoic to Cenozoic deposits in Iran: Implications for chronostratigraphy and collisional tectonics: Tectonophysics, v. 451, p. 97–122, doi:10.1016/j.tecto.2007.11.063. Jackson, M.P.A., Cornelius, R.R., Craig, C.H., Gansser, A., Stocklin, J., and Talbot, C.J., 1990, Salt Diapirs of the Great Kavir, Central Iran: Geological Society of America Memoir 177, 139 p., doi:10.1130/MEM177-p1. Javadi, H.R., Ghassemi, M.R., Shahpasandzadeh, M., Guest, B., Esterabi Ashtiani, M., Yasaghi, A., and Kouhpeyma, M., 2013, History of faulting on the Doruneh fault system: Implications for the kinematic changes of the Central Iranian microplate: Geological Magazine, v. 150, no. 4, p. 651–672, doi:10.1017/S0016756812000751. Lindenberg, H.G., Görler, K., Jacobshagen, V., and Ibbeken, H., 1984, PostPaleozoic stratigraphy, structure and orogenic evolution of the Southern Sabzevar zone and the Taknar block (Khorasan, NE Iran): Neues Jahrbuch für Geologie und Palaontologie, Abhandlungen, v. 168, no. 2/3, p. 287–326. Lowell, J.D., 1972, Spitsbergen Tertiary orogenic belt and the Spitsbergen fracture zone: Geological Society of America Bulletin, v. 83, p. 3091–3102, doi:10.1130/0016-7606(1972)83[3091:STOBAT]2.0.CO;2. Mattei, M., Cifelli, F., Muttoni, G., Zanchi, A., Berra, F., Mossavvari, F., and Eshraghi, S.A., 2012, Neogene block rotation in central Iran: Evidence from paleomagnetic data: Geological Society of America Bulletin, v. 124, no. 5–6, p. 937–956. McClay, K., and Bonora, M., 2001, Analog models of restraining stepovers in strike-slip fault systems: American Association of Petroleum Geologists Bulletin, v. 85, no. 2, p. 233–260. McKenzie, D., 1972, Active tectonic of the Mediterranean region: Geophysical Journal of the Royal Astronomical Society, v. 30, p. 109–185, doi:10.1111/j.1365-246X.1972.tb02351.x. Meyer, B., and Le Dortz, K., 2007, Strike-slip kinematics in central and eastern Iran: Estimating fault slip-rates averaged over the Holocene: Tectonics, v. 26, TC5009, doi:10.1029/2006TC002073. Mohafez, S., 1963, Geology and Oil Possibilities of Khur-Jandaq-Biabanak Area: Tehran, National Iranian Oil Company Geological Report, 250 p. Mouthereau, F., Lacombe, O., and Verges, J., 2012, Building the Zagros collisional orogeny: Timing, strain distribution and the dynamics of Arabia/ Eurasia plate convergence: Tectonophysics, v. 532–535, p. 27–60, doi:10.1016/j.tecto.2012.01.022.

Kinematics of the Great Kavir fault inferred from analysis of the Pees Kuh Complex, Jandaq area, Iran Nowroozi, C.A.F., 1972, Focal mechanism of earthquakes in Persia, Turkey, West Pakistan, and Afghanistan and plate tectonics of the Middle East: Bulletin of the Seismological Society of America, v. 62, no. 3, p. 823–850. Nozaem, R., Mohajjel, M., Rossetti, F., Della Seta, M., Vignaroli, G., Yassaghi, A., Salvini, F., and Eliassi, M., 2013, Post-Neogene right-lateral strikeslip tectonics at the north-western edge of the Lut block (Kuh-e-Sarhangi fault), central Iran: Tectonophysics, v. 589, p. 220–233, doi:10.1016/j .tecto.2013.01.001. Nur, A., Ron, H., and Scotti, O., 1986, Fault mechanics and kinematics of block rotation: Geology, v. 14, p. 746–749, doi:10.1130/0091 -7613(1986)142.0.CO;2. Perfilyev, Y.S., Aistov, L., Selivanov, E., and Melnikov, B., 1978, Geological Quadrangle Map of Chupanan: Tehran, Geological Survey of Iran, scale 1:100,000. Rahmati-Ilkhchi, M., Jerabek, P., Faryad, S., and Koyi, H., 2010, MidCimmerian, Early Alpine and late Cenozoic orogenic events in the Shotur Kuh metamorphic complex, Great Kavir block, NE Iran: Tectonophysics, v. 494, p. 101–117, doi:10.1016/j.tecto.2010.09.005. Raymond, L.A., 1984, Classification of mélanges, in Raymond, L.A., ed., Mélanges: Their Nature, Origin, and Significances: Geological Society of America Special Paper 198, p. 7–20, doi:10.1130/SPE198-p7. Ron, H., Freund, R., and Garfunkel, Z., 1984, Block rotation by strike-slip faulting: Structural and paleomagnetic evidence: Journal of Geophysical Research, v. 89, no. B7, p. 6256–6270, doi:10.1029/JB089iB07p06256. Rossetti, F., Nasrabady, M., Vignaroli, G., Theye, T., Gerdes, A., Razavi, S.M.H., and Moin, V.H., 2010, Early Cretaceous migmatitic mafic granulites from the Sabzevar range (NE Iran): Implications for the closure of the Mesozoic peri-Tethyan oceans in central Iran: Terra Nova, v. 22, p. 26–34, doi:10.1111/j.1365-3121.2009.00912.x. Sahandi, M.R., and Soheili, M., 2011, Geological Map of Iran: Tehran, Geological Survey of Iran, scale 1:1,000,000. Sengör, A.M.C., and Natal’in, B.A., 1996, Paleotectonics of Asia: Fragments of a synthesis, in Yin, A., and Harrison, T.M., eds., The Tectonic Evolution of Asia: Cambridge, UK, Cambridge University Press, p. 486–640. Sharkovski, M., Susov, M., and Krivyakin, B., 1984, Geology of the Anarak Area (Central Iran). Explanatory Text of the Anarak Quadrangle Map 1:250,000: Tehran, Geological Survey of Iran, and Moscow, V/O “Technoexport” Report 19, scale 1:250,000, 143 p. Stampfli, G.M., and Borel, G.D., 2002, A plate tectonic model for the Paleozoic and Mesozoic constrained by dynamic plate boundaries and restored synthetic oceanic isochrones: Earth and Planetary Science Letters, v. 196, p. 17–33, doi:10.1016/S0012-821X(01)00588-X. Stöcklin, J., 1968, Structural history and tectonics of Iran: A review: American Association of Petroleum Geologists Bulletin, v. 52, no. 7, p. 1229–1258. Stöcklin, J., 1974, Possible ancient continental margins in Iran, in Burk, C.A., and Drake, C.L., eds., The Geology of Continental Margins: Berlin, Springer-Verlag, p. 873–887, doi:10.1007/978-3-662-01141-6_64. Susov, M., Dvoryankin, A., and Selivanov, E., 1979, Geological Quadrangle Map of Nakhlak, Sheet 6757: Tehran, Geological Survey of Iran, scale 1:100,000. Susov, M., Dvoryankin, A., Selivanov, E., and Desyaterik, N., 1984, Geological Quadrangle Map of Jandaq: Tehran, Geological Survey of Iran, scale 1:100,000.

227

Sylvester, A.G., 1988, Strike-slip faults: Geological Society of America Bulletin, v. 100, p. 1666–1703, doi:10.1130/0016-7606(1988)1002.3.CO;2. Sylvester, A.G., and Smith, R.R., 1976, Tectonic transpression and basementcontrolled deformation in San Andreas fault zone, Salton Trough, California: American Association of Petroleum Geologists Bulletin, v. 60, no. 12, p. 2081–2102. Takin, M., 1972, Iranian geology and continental drift in the Middle East: Nature, v. 235, p. 147–150, doi:10.1038/235147a0. Taylor, M., Yin, A., Ryerson, F.J., Kapp, P., and Ding, L., 2003, Conjugate strike-slip faulting along the Bangong-Nujiang suture zone accommodates coeval east-west extension and north-south shortening in the interior of the Tibetan Plateau: Tectonics, v. 22, no. 4, p. 18-1–18-21. Tchalenko, J., Berberian, M., and Behzadi, H., 1973, Geomorphic and seismic evidence for recent activity of the Doruneh fault, Iran: Tectonophysics, v. 19, p. 333–341, doi:10.1016/0040-1951(73)90027-9. Tikoff, B., Kelso, P., Manduca, C., Markley, M.J., and Gillaspy, J., 2001, Lithospheric and crustal reactivation of an ancient plate boundary: The assembly and disassembly of the Salmon River suture zone, Idaho, USA, in Holdsworth, R.E., Strachan, R.A., Magloughlin, J.F., and Knipe, R.J., eds., The Nature and Tectonic Significance of Fault Zone Weakening: Geological Society, London, Special Publication 186, p. 213–231, doi:10.1144/GSL.SP.2001.186.01.13. Tirrul, R., Bell, I.R., Griffis, R.J., and Camp, V.E., 1983, The Sistan suture zone of eastern Iran: Geological Society of America Bulletin, v. 94, p. 134– 150, doi:10.1130/0016-7606(1983)942.0.CO;2. Valeh, N., and Haghipour, A., 1972, Geological Quadrangle Map of Ardekan: Tehran, Geological Survey of Iran, scale: 1:250,000. Vernant, Ph., Nilforoushan, F., Hatzfeld, D., Abbassi, M.R., Vigny, C., Masson, F., Nankali, H., Martinod, J., Ashtiani, A., Bayer, R., Tavakoli, F., and Chery, J., 2004, Present-day crustal deformation and plate kinematics in the Middle East constrained by GPS measurements in Iran and northern Oman: International Journal of Geophysics, v. 157, p. 381–398, doi:10.1111/j.1365-246X.2004.02222.x. Walker, R., and Jackson, J., 2004, Active tectonics and late Cenozoic strain distribution in central and eastern Iran: Tectonics, v. 23, TC5010, doi:10.1029/2003TC001529. Wellman, H.W., 1966, Active wrench faults of Iran, Afghanistan, and Pakistan: Geologische Rundschau, v. 55, p. 716–735, doi:10.1007/BF02029650. Wilcox, R.E., Harding, T.P., and Seely, D.R., 1973, Basic wrench tectonics: American Association of Petroleum Geologists Bulletin, v. 57, p. 74–96. Wilmsen, M., Fürsich, F.T., and Majidifard, J., 2015, An overview of the Cretaceous stratigraphy and facies development of the Yazd block, western central Iran: Journal of Asian Earth Sciences, v. 102, p. 73–91, doi:10.1016/j .jseaes.2014.07.015. Woodcock, N., 1986, The role of strike-slip fault systems at plate boundaries: Royal Society of London Philosophical Transactions, ser. A, v. 317, p. 13–29. Woodcock, N., and Fischer, M., 1986, Strike-slip duplexes: Journal of Structural Geology, v. 8, p. 725–735, doi:10.1016/0191-8141(86)90021-0. MANUSCRIPT ACCEPTED BY THE SOCIETY 15 JUNE 2016 MANUSCRIPT PUBLISHED ONLINE 26 OCTOBER 2016

Printed in the USA

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084785/spe525-06.pdf by University of Wisconsin Madison user

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084785/spe525-06.pdf by University of Wisconsin Madison user

The Geological Society of America Special Paper 525

Mid-ocean-ridge to suprasubduction geochemical transition in the hypabyssal and extrusive sequences of major Upper Cretaceous ophiolites of Iran Morteza Khalatbari Jafari* Research Institute for Earth Sciences, Geological Survey of Iran, Box 13185-1494, Tehran, Iran Hassan A. Babaie* Department of Geosciences, Georgia State University, Atlanta, Georgia 30302, USA Mohammad Elyas Moslempour* Research Center for Earth Science, Department of Geology, Zahedan Branch, Islamic Azad University, Zahedan, Iran

ABSTRACT We discuss geochemical, chronological, and field data from the extrusive sequence and individual diabase and sheeted dikes in the Upper Cretaceous Khoy, Kermanshah, Fannuj, Nosratabad, south Fariman, northwest Fariman, Dehshir, and Sabzevar ophiolite massifs of Iran. The extrusive sequences include pillow lava, sheet flow, hyaloclastite, hyaloclastic breccia, and interbeds of chert and pelagic limestone with Late Cretaceous microfauna. The Khoy, northwest Fariman, and Sabzevar massifs also include Upper Cretaceous–Lower Paleocene supra-ophiolitic volcanic and volcanic-sedimentary rocks that formed in a trough near the basin where the extrusive sequence formed. The Khoy pillow lava displays transitional (T) mid-ocean-ridge basalt (MORB) characteristics but no chemical contribution from the components released from the subducted slab. On the other hand, the diabase dikes that cut the Khoy extrusive sequence show signatures of subduction-zone magmatism and contribution from the melt released through the partial melting of the subducted slab. While lava in the Harsin (Kermanshah) extrusive sequence in west Iran displays enriched (E) MORB and plume (P) MORB characteristics, the pillows in the Fannuj, northwest Fariman, Dehshir, and Sabzevar extrusive sequences indicate the contribution of both fluids and melt from the subducted slab. The Nosratabad and south Fariman ophiolites also show evidence for either melt or fluids, respectively. Partial melting of the subducted slab sedimentary cover may have formed the acidic pillow lava and sheet flow in the Fannuj and Nosratabad extrusive sequence, respectively. Some pillows in the Nosratabad, Sabzevar, northwest Fariman, and to a lesser extent, Dehshir

*E-mails: [email protected] (corresponding author); [email protected]; [email protected]. Khalatbari Jafari, M., Babaie, H.A., and Moslempour, M.E., 2016, Mid-ocean-ridge to suprasubduction geochemical transition in the hypabyssal and extrusive sequences of major Upper Cretaceous ophiolites of Iran, in Sorkhabi, R., ed., Tectonic Evolution, Collision, and Seismicity of Southwest Asia: In Honor of Manuel Berberian’s Forty-Five Years of Research Contributions: Geological Society of America Special Paper 525, p. 229–289, doi:10.1130/2016.2525(07). © 2016 The Geological Society of America. All rights reserved. For permission to copy, contact [email protected].

229

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084793/spe525-07.pdf by University of Wisconsin Madison user

230

Khalatbari Jafari et al. extrusive sequence display oceanic-island basalt (OIB) geochemical characteristics. Mantle plumes or asthenospheric flow that probably moved up through weak zones of the subducted slab may have affected the partial melting of the mantle wedge above the slab. The combined OIB and suprasubduction characteristics suggest the role of rollback of the subducted slab in the magmatism of the northeast Iranian ophiolites. The clear MORB-like geochemical characteristics in the extrusive sequence of the ophiolites in northwest and west Iran, and the suprasubduction-zone characteristics in the diabase and extrusive sequence of the ophiolites in southeast and east Iran, west of the Lut block, and northeast Iran may represent a major Late Cretaceous transition from a MORB to a suprasubduction-zone setting.

INTRODUCTION As a part of the Tethyan ophiolites, which were emplaced during the Alpine-Himalayan orogeny (Dilek and Flower, 2003), Iranian ophiolites connect to the Pakistan-Himalayan ophiolites to the east and those in Turkey to the west. Knipper et al. (1986) classified the Tethyan ophiolites into 10 groups (Fig. 1), with Iranian ophiolites constituting the Van, Central Iran, MakranZahedan, and peri-Arabic (ophiolitic crescent) groups. The Tethyan ophiolites and suture zones of the Alpine-Himalayan orogen are shown schematically in Figure 2 (Dilek and Furnes, 2009). The positions of the Iranian ophiolites were revised on this map based on new data from the distribution of ophiolites in Iran (National Geosciences Database of Iran, 2005). The Iranian ophiolitic massifs mainly occur along major structural zones of Iran, such as the Zagros, Alborz, Central Iran, and Eastern suture zones, and Makran accretionary prism (Fig. 3). In this paper, we compare and contrast the lithology and geochemistry of the Upper Cretaceous Iranian ophiolitic massifs in northwest and west Iran with coeval ophiolites in southeast, east, and northeast Iran. The ophiolite massifs studied in this paper include: (1) Khoy; (2) Ker-

manshah; (3) Fannuj; (4) Nosratabad; (5) Dehshir; (6) south Fariman; (7) NW Fariman; and (8) Sabzevar. Field studies on these eight ophiolite outcrops started in 1999. METHODOLOGY AND DATA In this paper, except for the data from the Kermanshah ophiolites, which were locally published in Persian with English abstracts (Allahyari et al., 2012), the geochemical data used in this comparative study are original and unpublished. Field mapping was done using 1:25,000 and 1:50,000 scale Advanced Spaceborne Thermal Emission and Reflection Radiometer (ASTER) and Satellite Pour l’Observation de la Terre (SPOT) satellite imagery and 1:50,000 topographic maps and aerial photographs. Samples from the Khoy ophiolite were analyzed in the geochemical laboratory at Université de Bretagne Occidentale, Brest, France, with the laboratory techniques described in Khalatbari Jafari et al. (2006). Samples from the southeast, east, and northeast Iranian ophiolites were analyzed with the inductively coupled plasma–atomic emission spectrometry (ICP-AES) and inductively coupled plasma–mass spectrometry (ICP-MS)

Figure 1. Map showing the distribution of the 10 groups of the Tethyan ophiolites classified by Knipper et al. (1986): (1) Ligurian, (2) Dinaride-Hellenics, (3) Carpathian, (4) peri-Arabic (ophiolitic crescent), (5) Pontic-Lesser Caucasus, (6) Van, (7) Central Iran, (8) Makran-Zahedan, (9) Pakistan-Kabul, (10) War. The Iranian ophiolites fall under groups 4, 6, 7, and 8.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084793/spe525-07.pdf by University of Wisconsin Madison user

Figure 2. Distribution of major Tethyan ophiolites and suture zones in the Alpine-Himalayan orogenic system (modified from Dilek and Flower, 2003). The locations of major Iranian ophiolites are based on the map published by the Geological Survey of Iran (2005).

Mid-ocean-ridge to suprasubduction geochemical transition in hypabyssal and extrusive sequences

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084793/spe525-07.pdf by University of Wisconsin Madison user

231

232

Khalatbari Jafari et al.

Figure 3. Distribution of the major ophiolites in Iran discussed in this paper (from the 1:100,000 magmatic map of Iran; Emami et al., 1993). The major geological units are from the sedimentary-structural map of Iran (Aghanabati, 2004). Modifications are based on authors’ geochemical data. OIB—oceanicisland basalt; MORB—mid-ocean-ridge basalt; N—normal; T—transitional; P—plume.

methods at the SGS laboratory in Toronto, Canada. After preparing thin sections and their petrographic study, samples from the ophiolitic rocks were selected to determine their geochemical characteristics and tectonomagmatic setting. The samples were sent to the SGS laboratory in Toronto, Ontario, Canada, where they were analyzed for major and some trace elements (Ba, Sr, Y, Zr, and Zn) with ICP-AES, and other trace elements (including rare earth elements [REEs]) by ICP-MS. The ICP-AES analysis used ~0.1 g of rock powder that was first mixed with 0.9 g of lithium borate in a carbon crucible and then melted in a furnace at 1100 °C for 30 min. The fused glass samples were dissolved (using a magnetic stirring device) in 100 mL of 1% HNO3 solution with Ge as an internal standard. The ICP-MS analysis of the trace elements used 0.1 g of rock powder that was dissolved in a mixture of HF, HCl, and HNO3 in screw-top Teflon vials.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084793/spe525-07.pdf by University of Wisconsin Madison user

An internal standard solution containing indium was then added, and the spiked sample dissolution was diluted with 1% HNO3. The internal standard was used for monitoring drift in mass response during measurement. The geochemical results are given in Tables 1–5. Table 6 synthesizes paleontological and geochronological data for the eight Upper Cretaceous ophiolites from Iran discussed in this paper. Table 7 gives a tectonic comparison of these ophiolitic massifs. PRE–LATE CRETACEOUS OPHIOLITES Significant numbers of Upper Cretaceous Iranian ophiolitic complexes occur in fault contact next to remnants of ophiolitic bodies that are older than Late Cretaceous. This suggests that the

Mid-ocean-ridge to suprasubduction geochemical transition in hypabyssal and extrusive sequences fate of the Upper Cretaceous ophiolites may somehow be intertwined with that of the pre–Late Cretaceous ophiolites, including those that are remnants of the Paleotethys Ocean. For this reason, we first provide a discussion of these older ophiolites. The pre–Late Cretaceous ophiolitic remnants in Iran are mostly related to Paleozoic, Late Triassic, and Jurassic events. These ophiolites occur as scattered, tectonized, and more or less metamorphosed exposures in the following areas: Soghan in the extreme southeast end of the Esfandagheh region of the Sanandaj-Sirjan zone (Sabzehei, 1974, 1998; Ahmadipour et al., 2003), Jandag-Anarak along the northwestern margin of Central Iran (Sharkovski et al., 1984; Bagheri and Stampfli, 2008; Torabi, 2009, 2011, 2012; Buchs et al., 2013), south of Mashhad (Alavi, 1979; Majidi, 1981; Alavi, 1991, 1992; Eftekharnezhad and Behroozi, 1991; Ghazi et al., 2001), Asalem-Shanderman in the Talesh Mountains, southeast of the Caspian Sea (Şengör, 1984; Alavi, 1996; Eftekharnezhad et al., 1992), and Misho Mountain, northwest of Tabriz (Saccani et al., 2013a). In the following section, we briefly present the pre–Late Cretaceous ophiolites as Paleotethys remnants and early–middle Mesozoic ophiolites as part of Neotethys. Paleotethys Remnants The Soghan ultramafic-mafic complex in the Esfandagheh area at the southeastern end of the Sanandaj-Sirjan zone, in southeast Iran, includes dunite, harzburgite, and chromitite that grade into lherzolite and wehrlite, which in turn grade upward into layered gabbro. The complex is covered by the Sargaz-Abshur metamorphic complex of marble and amphibolite (Sabzehi, 1974, 1998; Ahmadipour et al., 2003). It is believed that the Soghan complex formed as a result of the differentiation of komatiitic magma during the late Proterozoic–early Paleozoic as mantle diapirs in an intercontinental rift basin (Sabzehi, 1974, 1998). It is probable that the whole Soghan sequence was metamorphosed into the amphibolite facies due to an early Cimmerian orogenic phase (Ahmadipour et al., 2003). Based on their study of the platinum group elements (PGEs) in 20 chromitite mines in the Esfandagheh, Sikhoran, Faryab, and Neyriz ophiolites, Jannessary et al. (2012) concluded that large-scale, two-stage partial melting in a suprasubduction zone probably produced boninitic melts that crystallized refractory Cr-Mg–rich and Ti-poor chromitites. The Dehshaikh peridotitic massif in the Esfandagheh area includes harzburgite, dunite, and several deposits of chromitite that are cut by pyroxenitic dikes. Peighambari et al. (2011) attributed the Dehshaikh massif to a suprasubduction zone based on the mineralogy of the peridotitic massif and their associated chromitites. One of the most controversial remnants of old Iranian ophiolites northwest of the Central Iranian microcontinent is exposed around Anarak and Jandaq. This complex includes tectonized remnants of serpentinized peridotite and scattered masses of gabbro, diabase, and plagiogranite, with basalt and tuff metamorphosed to the greenschist facies, and interlayers of shale, chert, and carbonate at the top (Davoudzadeh et al., 1981; Alma-

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084793/spe525-07.pdf by University of Wisconsin Madison user

233

sian, 1997). The lower contact of the Anarak and Jandaq ophiolite is not seen, but at its top, the ophiolite contains algal carbonate known as Lak marble. The Lak marble may correlate with the Kabodan marble in the Khor area (northeast of Nain), which has an Early Cambrian age based on archaeocyathids. Davoudzadeh et al. (1981) attributed the Anarak ophiolite to the Carboniferous Paleotethys and considered it to be a continuation of the Herat ophiolite in Afghanistan, which was displaced to its position in Anarak due to the rotation of the Central Iran microcontinent. The contact between the Anarak metamorphic complex and the Lak marble is a fault (Buchs et al., 2013). Sharkovski et al. (1984) determined the wide range of 400–120 Ma (K/Ar) for the age of the Anarak metamorphic complex and attributed it to retrograde metamorphism. Bagheri and Stampfli (2008) named the metamorphic rock complex lying to the northwest of the CentralEast Iranian microcontinent as a Variscan accretionary complex that includes the Jandaq and Anarak ophiolite complexes and the marginal sea ophiolitic remnants in the Kabodan area. Serpentinized peridotite occurs as large slivers in the middle part of the Anarak metamorphosed complex (Davoudzadeh et al., 1981). The middle part of the Anarak metamorphic complex locally includes pillow lava, gabbro, trondhjemite, and mafic dikes that have escaped metamorphism (Bagheri and Stampfli, 2008; Torabi, 2011, 2012). Bagheri and Stampfli (2008) provided a 333–320 Ma 40Ar/39Ar radiometric age for large muscovite that represents an older metamorphic event in the complex. These authors reported a 215 Ma zircon age for a granite pegmatite that cuts the northern part of the Variscan complex in the Jandaq area. Bagheri and Stampfli (2008) also reported 40Ar/39Ar radiometric muscovite and hornblende ages of 163 Ma and 156 Ma, respectively, for the metamorphic rocks on the northern flank of Rashid Kuh and interpreted these ages to the cooling and exhumation of the Jandaq complex during the Late Triassic or earlier. Isotopic and paleontologic data attribute the age of the deposition of the Variscan accretionary complex to the Early Devonian–Carboniferous (Bagheri and Stampfli, 2008). Torabi (2009) considered mantle peridotite to be the main rock component of the Jandaq ophiolite that has gone through several metamorphic events. He interpreted the absence of chromitite to indicate a low-Cr and high-Al mantle origin and attributed the composition to the lherzolitic peridotite. Buchs et al. (2013) identified geochemical characteristics of subduction zone, mid-ocean ridge, oceanic interplate, and continental rift settings for the meta-igneous rocks of the metamorphic complex in the Anarak and Meraji areas and interpreted these rocks to represent the evolution of the Paleotethys during a whole Wilson cycle between 450 and 225 Ma. Scattered remnants of the Paleotethys are exposed in northeast Iran in the area between Mashhad and north of Fariman, and to the east to Torbat-e Jam. Ultramafic rocks are also exposed in the southwest around the periphery of the Agh-Darband erosional window in Darreh-Anjir. The contact between these and the neighboring rocks is a fault. Northeast of Fariman and Sefid Sang, the fault zone contains serpentinized peridotite (lherzolite and harzburgite), cumulate gabbro, pyroxenite, and wehrlite,

234

Khalatbari Jafari et al.

and the extrusive sequence contains extensive vesicular pillow lava, hyaloclastic breccia, and minor sheet flow (Zanchetta et al., 2013). South of the village of Abgarm and north of Sefid Sang, the extrusive sequence of mostly sheet flow is thrust over the conglomerate of the Middle Jurassic Kashf Rud Formation. In Mashhad, south of the village of Virani, volcanic-sedimentary rocks of the upper part of the extrusive sequence represent the remnants of the Paleotethys and are metamorphosed to the greenschist-amphibolite facies around the village of Khalaj. South of Mashhad, a flysch sequence, which is faulted at its base, contains lherzolite pseudosills at the base, which grade into diabase-microgabbro and andesite-basalt toward the top. Most of the sills have a hyaloclastic margins with hydrated green minerals, suggesting their interaction with water. A complex of slivers of metamorphosed volcanic-sedimentary rocks with turbidite features, such as meta-chert, radiolarite, recrystallized pelagic chert, and highly altered andesitic-basaltic lava, extends from north of Fariman to east of Sefid Sang. These rocks may have formed in a trough along the northern margin of the Paleotethys Ocean. According to Stöcklin (1974), it is possible that the ultrabasic rocks around Mashhad are relicts of the Paleotethys that separated Gondwana from Agh Darband during the Paleozoic. The meta-ophiolites of Mashhad may have been thrusted onto the flysch as allochthonous sheets (Alavi, 1979, 1991, 1992). The Late Devonian–Carboniferous rocks in the Agh Darband area have lithological similarities to coeval formations in Central Iran and east of Alborz (Eftekharnezhad and Behroozi, 1991). These authors interpreted the central and northern Alborz, including the Kopet Dagh, to represent the narrow shelf margin of the epi-Baikalian Afro-Arabian platform and saw no evidence for the Hercynian orogeny in this area. Eftekharnezhad and Behroozi (1991) inferred that tectonic extension led to the eruption of ultrabasic and basic magma in the area north of Fariman– Torbat-e-Jam, along the northernmost Permian Afro-Arabian margin platform. The 40Ar/39Ar and K/Ar dating of hornblende gabbro (Ghazi et al., 2001) in the Paleotethys remnants yielded Pennsylvanian–Early Permian ages of 281.4 and 277 Ma, and 273 and 265 Ma, respectively. Remnants of Paleotethys are cut by granitoid masses in Mashhad. The Late Triassic (Norian) 217 Ma and 215 Ma zircon U-Pb isotopic ages in the Kuhsangi granodiorite and Dehnow diorite, respectively (Karimpour et al., 2010), yielded evidence for pre–Late Triassic metamorphism of the remnants of the Paleotethys Ocean. Rocks from the Fariman volcanic-sedimentary units and Darreh Anjir complex may have been deposited in a Permian subsiding basin, possibly in a suprasubduction setting, in which siliciclastic turbidites, representing debris from the magmatic arc and its basement, were intercalated with carbonate and transitional and calc-alkaline basaltic lava (Zanchetta et al., 2013). Zanchetta et al. (2013) attributed the rocks of the Fariman and Darreh Anjir complexes to the magmatic arc and related basins that existed before the collision of the Iranian and Eurasian plates between the southern margin of Eurasia and the northwardsubducting Paleotethys oceanic plate.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084793/spe525-07.pdf by University of Wisconsin Madison user

The Shanderman-Asalem metamorphic complex, exposed in the Talesh Mountains, southwest of the Caspian Sea, includes greenschist-facies schist, gneiss, mica schist, and basic volcanic rocks. The metamorphism and the scarcity of paleontologic data make it difficult to date the complex; the age has been approximated to be Precambrian (Clark et al., 1975). However, Şengör (1984) and Alavi (1996) related the Shanderman complex to the Paleotethys ophiolitic suture that extends from the Talesh Mountains in the west to the exposed remnants of the Paleotethys Ocean in the Mashad-Binaloud area in east Iran. Eftekharnezhad et al. (1992) attributed the volcaniclastic sediments of the Shanderman-Asalem metamorphic complex to the continental margin. These authors reported a Permian–Carboniferous paleontologic age for the Shanderman-Asalem limestone and correlated the metamorphic complex to the Mashhad-Binaloud remnants of Paleotethys. Zanchetta et al. (2009) surmised that the Shanderman metamorphic complex included metabasite, with a well-preserved eclogite-phase assemblage that was later intruded by midcrustal intermediate-basic intrusive bodies. Paragonitic white mica, in equilibrium with high-pressure assemblages, give a Pennsylvanian (315 Ma) Ar/Ar age (Zanchetta et al., 2009). These authors suggested that the metamorphic basement of the complex was a part of a Paleozoic orogenic belt, and that the complex represents a fragment of the Upper Paleozoic European continental crust. Remnants of the Paleotethys Ocean in the East Azerbaijan province, in northwest Iran, can be traced from the east of Misho to southwest of Marand as gabbro and dolerite dikes that cut the continental basement (Asadian et al., 1994). The Misho Mountain is cut, in the north and south, by oblique strike-slip faults. The northern fault is the continuation of the Tabriz fault, which extends from north of the Central Iran tectonic province through Azerbaijan to Turkey. These faults were responsible for the uplift of the Precambrian–Jurassic units by major positive flower structures (Eftekharnezhad et al., 1989). Saccani et al. (2013a) reported an isotopic zircon (U-Pb) age of 356 Ma ([Mississippian) for a leucogabbroic dike from the Misho mafic complex in Misho Mountain. Some of the gabbro and basaltic dikes display normal (N) mid-ocean-ridge basalt (MORB) geochemical characteristics, while others display characteristics of plume (P) MORB (Saccani et al., 2013a). According to Saccani et al. (2013a), the Misho mafic complex may have formed through the interaction of a depleted MORB-like asthenosphere and mantle plume. These authors suggested that the mafic-ultramafic rocks in the Misho mafic complex may have formed through a Carboniferous magmatic event during the breakup of the northern edge of Gondwana, which led to the opening of the Paleotethys, which in turn was originally initiated by a mantle plume. EARLY–MIDDLE MESOZOIC OPHIOLITES Mesozoic ophiolites are distributed along the borders of main geological provinces such as Zagros and the Central Iran microcontinent, and in Makran and Khoy. The early to middle Meso-

Mid-ocean-ridge to suprasubduction geochemical transition in hypabyssal and extrusive sequences zoic ophiolites are metamorphosed to greenschist-amphibolite facies, with local signs of polymetamorphism. For example, the Khoy meta-ophiolite, tectonically included in a metamorphic subduction complex, yielded an Early Jurassic apparent age for its metamorphic amphiboles, while its primary magmatic age may be pre-Jurassic (Khalatbari Jafari, 2002; Khalatbari Jafari et al., 2003, 2004). In the Sikhoran complex in central Iran, multiple gabbroic dikes, veins, and plutons cut through a mafic/ ultramafic complex and its metamorphic cover. Several basins, characterized by abundant submarine basaltic extrusion, developed during the Jurassic, and the corresponding feeding dikes may be represented by the diabasic dike swarms that intrude the whole Sikhoran complex and its metamorphic cover (Ghasemi et al., 2002). The ophiolite massifs of Kahnuj, western Makran, and the southern Central Iran microcontinent contain a variety of cumulate to massive gabbro, diabase sheeted dike complexes, and basaltic pillow lavas in which amphiboles have yielded a Jurassic apparent 40Ar/40Ar age between 150 and 139 Ma (Kananian et al., 2001). In northeastern Iran, around Soltanabad, Rossetti et al. (2009) assigned an Early Cretaceous (Albian) isotopic age for the peak metamorphism. The metamorphic rocks of Soltanabad are considered to represent the remnants of the Neotethys Ocean in northeastern Iran. According to Ghazi et al. (2004), the Makran accretionary prism in southeast Iran contains intact Mesozoic ophiolite and ophiolitic mélange. The Band-e-Zeyarat/Dar-Anar complexes, located to the northwest of the accretionary prism, represent the extrusive crustal ophiolite sequence. The Band-e-Zeyarat/ Dar-Anar complex contains cumulate layered gabbro, isotropic gabbro, diorite, trondhjemite, diabase sheeted dike, and pillow lava with pelagic sediment interlayers. There is no peridotite in these complexes. Three 40Ar/39Ar plateau ages in the 140–142 Ma range, and five K/Ar ages in the 146–121 Ma range, from hornblende gabbro, indicate that the Band-e Zeyarat/Dar-Anar complex formed during Late Jurassic–Early Cretaceous time (Ghazi et al., 2004). The Band-e-Zeyarat/Dar-Anar igneous ophiolitic rocks were derived by fractionation from melts with a composition similar to average enriched (E) MORB (Ghazi et al., 2004). The Birjand ophiolite complex, north of the Sistan suture zone, displays the characteristics of a MORB-type ophiolite. Two U-Pb zircon ages from the leucogabbro of the Birjand ophiolite yielded Middle Cretaceous ages of 113 ± 1 Ma and 107 ± 1 Ma (Zarrinkoub et al., 2012). GEOLOGICAL SETTINGS OF UPPER CRETACEOUS OPHIOLITES Figures 1 and 2 and Tables 6 and 7 show the classifications (Knipper et al., 1986) and locations of the ophiolitic massifs of Iran, respectively. These massifs align with the main structural zones in Iran (Fig. 3). The Upper Cretaceous massifs are among the most extensive Neotethyan ophiolites in Iran. The largest Neotethyan ophiolitic exposures (Fig. 3) in northwest Iran are near the border with Turkey and include the metamor-

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084793/spe525-07.pdf by University of Wisconsin Madison user

235

phosed and nonmetamorphosed sections of the Khoy ophiolite. The Kermanshah ophiolites (Fig. 3) in west Iran and Neyriz ophiolite in south-central Iran lie along the northern edge of the Zagros fold belt (Fig. 3). The ophiolites in the Makran zone are exposed in east Iran between the area east of Iranshahr to around Bandar-e-Abbas (Fig. 3). These are known as Fannuj, Rameshk, Mokhtarabad, Kahnuj, and Band-e-Zeyarat/Dar-Anar (McCall, 1985, 1997). The ophiolites in the east Iran suture zone, which extend from north of Birjand to south of Zahedan along two tectonized belts, are known with the following local names: Birjand, Sarbisheh, Nehbandan, Techelkureh, Nosratabad, and Zahedan (Fig. 3). The ophiolites along the interior margin of the Central Iran microcontinent are known as: Shahrbabak, DehshirKahdoiyeh, Marvast, and Nain (Fig. 3). The south Fariman and Torbat-e-Heydarieh ophiolites lie to the north of the Central Iran microcontinent. The northwest Fariman and Sabzevar ophiolites lie between the northern margin of the Central Iran microcontinent and Alborz-Binaloud (Fig. 3). In this paper, we discuss the following eight Upper Cretaceous Iranian ophiolites: Khoy in northwest Iran, Kermanshah in west Iran, Fannuj in the Makran area of southeast Iran, Nosratabad in the east Iran suture zone, Dehshir in the area west of the Lut block, and northwest Fariman, south Fariman, and Sabzevar in northeast Iran. Paleontological studies of pelagic limestone in the extrusive sequences of the studied ophiolitic massifs yielded a Late Cretaceous age. The list of the microfauna for each massif is given in its corresponding section below. Khoy Ophiolite The Khoy ophiolite (Figs. 1, 2, and 3) falls under group 6 (Van Group) of Knipper et al. (1986; see also Fig. 1 herein), which is distinguished from group 4 (peri-Arabic) ophiolites that include the Kermanshah ophiolites. Kamineni and Mortimer (1975) provided the earliest description of the Khoy area. Ghorashi and Arshadi (1978) mapped the geological units (scale: 1:250,000), and Radfar et al. (1993) and Amini et al. (1993) (with contribution of the first author) provided the first detailed description of the Upper Cretaceous Khoy ophiolite complex in the geological maps of Khoy and Dizaj with a scale of 1:100,000. Hassanipak and Ghazi (2000) reported the E-MORB and N-MORB REE characteristics for the pillow lavas, and Ghazi et al. (2001) defined a basal metamorphic zone at the base of the Khoy ophiolite. Khalatbari Jafari et al. (2003, 2004, 2006) indicated that there are two ophiolite complexes in the Khoy area: an older (112–194 Ma), multiple-metamorphosed complex, and a younger (Upper Cretaceous) nonmetamorphosed complex (Fig. 4). The older complex occurs as tectonic slices of rocks that were metamorphosed during similar Early Jurassic, Middle Jurassic, Early Cretaceous, and Late Cretaceous metamorphic conditions. Ghazi et al. (2001, 2003) proposed a suprasubduction-zone model for the Khoy ophiolite based on the 118–111 Ma K-Ar ages that they determined from gabbro and amphibolite in the ophiolite. Pessagno et al. (2005) proposed the existence of three ophiolites in the

236

Khalatbari Jafari et al.

Figure 4. Simplified geological map showing the main geological units in the Khoy region (after Khalatbari-Jafari et al., 2004).

Khoy area: a Late Jurassic, a Late Cretaceous (early Coniacian) with N-MORB characteristics, and a Late Cretaceous (latest Campanian) with E-MORB characteristics. Aftabi et al. (2006) reported Cyprus-type characteristics for the copper massive sulfide lenses in the Upper Cretaceous, volcanogenic-exhalative units of Zurabad (Sekmanabad) near Khoy. Azizi et al. (2006), studying the pressure-temperature-time (P-T-t) path of the metamorphic rocks in Khoy, attributed them to a Late Cretaceous– Tertiary continental collision. Azizi et al. (2011) attributed the U-Pb zircon age of 595–566 Ma from two meta-granite samples, and a 550 Ma age from one amphibolite sample to late Proterozoic magmatic zircon crystallization ages.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084793/spe525-07.pdf by University of Wisconsin Madison user

There are five major NW-SE–trending geologic units in the Khoy area, which, from northeast to southwest, include (Fig. 4): (1) the southwestern margin of Central Iran; (2) the Eastern metamorphic complex, containing the meta-ophiolite complex; (3) Upper Cretaceous–Lower Paleocene volcanic-sedimentary and turbiditic supra-ophiolite series; (4) Upper Cretaceous nonmetamorphic Khoy ophiolite; and (5) the western metamorphic complex. The vergence of thrusting of these units is to the southwest, such that the Central Iran units lie above the Eastern metamorphic complex. The Upper Cretaceous Khoy ophiolite (Figs. 4 and 5A) includes large plutonic and extrusive sequences. The plutonic

Mid-ocean-ridge to suprasubduction geochemical transition in hypabyssal and extrusive sequences sequence includes lherzolite, clinopyroxene-bearing harzburgite, and serpentinized harzburgite. It contains chromitite lenses and extensive layered and cumulate gabbro. The gabbroic bodies include olivine gabbro, troctolite, pyroxene gabbro, ferro-gabbro, and anorthosite, which are cut by sills and wehrlitic intrusions, pegmatite gabbro, and individual diabase dikes. The extrusive sequence of the Khoy ophiolite is exposed in two northern and southern massifs in the Khoy area, and it forms

prominent volcanic piles in the southern massif in Jehnnem dere, and the Goldag highs in the northern massif. The thickest part of the extrusive sequence (>1000 m) in the Jehnnem dere (Fig. 5A) mostly contains volcanic rocks with minor volcanic-sedimentary rocks. This makes the southern massif the thickest extrusive sequence among Iranian ophiolites. Figure 5A depicts a schematic stratigraphic column of the Upper Cretaceous extrusive sequence of the Khoy ophiolite (located in Jehnnem dere), and

Figure 5. Schematic stratigraphic columns showing the main units of the (A) Upper Cretaceous extrusive sequence of the Khoy ophiolite in Jehnnem dere, and (B) supra-ophiolitic series in the Jehnnem dere-Sekmanabad valley.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084793/spe525-07.pdf by University of Wisconsin Madison user

237

238

Khalatbari Jafari et al.

Figure 5B depicts a schematic stratigraphic column of the Upper Cretaceous–Lower Paleocene supra-ophiolite series in the Khoy area. The extrusive sequence begins with sheet flows in the Jehnnem dere (Fig. 5A). Successive phyric and aphyric pillow lavas have interlayers of sheet flow, hyaloclastite, and pelagic limestone with Senonian-Santonian-Campanian microfauna (Khalatbari Jafari et al., 2004). Fossil lava lake and N-S–striking oceanic normal faults are seen in the northern massif. Volcanic blockyclast breccia and deep-slope debris avalanche breccia occur in the upper part of the extrusive sequence in the southern massif (Fig. 5A). Porphyritic diabase sills and intrusions, N-S–striking epidote and calcite veins, and sporadic sulfide minerals (pyrite and chalcopyrite) occur in the southern massif. The southern massif is thrust upon serpentinized peridotite in its southern contact and is covered by the turbidites of the supra-ophiolitic series in its northern and eastern side. The Upper Cretaceous–Lower Paleocene supra-ophiolitic volcanic-sedimentary series is exposed northwest of KhoySekmanabad valley along a NW-SE–striking belt (Fig. 5B). The supra-ophiolitic series highlights the vertical and lateral lithological variations. The supra-ophiolitic series is made up of four different sections with different lithologies (Fig. 5B): a lower section of turbidite, a second section of mostly pillow lava, a third section of epiclastic breccia and ankaramitic lava, and a top section of volcanic-sedimentary rocks. Paleontological studies of the microfauna in the pelagic limestone interbeds indicate a Campanian–Maastrichtian age for the lower three sections, and an early Paleocene age for the top section (Khalatbari Jafari et al., 2004), which is covered by Paleocene conglomerate. It seems that the supra-ophiolitic series is a remnant of a trough that formed near the Khoy Neotethyan Ocean at the end of the Cretaceous during the last stages of the formation of the extrusive ophiolite sequence. Twenty samples taken from the extrusive ophiolite sequence were analyzed (Khalatbari-Jafari et al., 2006), applying classification, spider, multi-elemental REE, and tectonomagmatic diagrams. The pillow lava, sheet flow, picrite, and diabase plot in the andesite/basalt field of the Nb/Y versus Zr/TiO2 classification diagram (Winchester and Floyd, 1976), except for one sample that plots in the andesite field (Fig. 6A). Most samples plot in the tholeiitic to calc-alkaline basalt field in the Th versus Co classification and magmatic trend identification diagram proposed by Hastie et al. (2007) for altered rocks, with the diabase samples plotting near the border of the high-K shoshonite field (Fig. 6B). The different position of the diabase samples compared to other rocks of the extrusive sequence on this diagram may reflect their different source. The chondritenormalized (Sun and McDonough, 1989) multi-elemental REE patterns of pillow lava, sheet flow, and picrite are systematic and flat, with the pillow samples plotting in the average T-MORB field (Fig. 6C; Khalatbari Jafari et al., 2006). The picrite patterns that plot lower than the average T-MORB suggest a depleted mantle source. The greater enrichment of the pattern of some pillow lava samples, and their distance from the field

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084793/spe525-07.pdf by University of Wisconsin Madison user

of the mean N-MORB (Fig. 6C) may reflect the presence of a hotspot component in the mantle source. No depletion of the high field strength elements (HFSEs) is apparent on the primitive mantle–normalized (Sun and McDonough, 1989) spider diagrams of the extrusive sequence of the Khoy ophiolite (Fig. 6D; Khalatbari Jafari et al., 2006). Except for the lithophile elements such as Rb, Ba, Th, and K, the trace elements are systematically distributed on these patterns. Chondrite-normalized (Sun and McDonough, 1989) multielemental REE patterns of the diabase samples (Fig. 6C) show a significant decrease from the light (L) REEs toward the heavy (H) REEs that correlates with that in the magmatic calc-alkaline patterns of the suprasubduction zones. The primitive mantle– normalized (Sun and McDonough, 1989) spider diagrams of the samples of diabase dikes that cut the lava of the northern massif display depletion of Nb, Ti, and Zr, and enrichment of Th and Ba (Fig. 6D). The higher Th and SiO2 contents in the diabase that cuts the northern massif, compared to those in the lavas of the extrusive sequence, suggest partial melting of the subducted slab (Parlak et al., 2009; Pearce, 2003). Compared to the samples taken from the diabase that plot in the arc field of the Nb/Th versus Y diagram (Jenner et al., 1991), the majority of the samples taken from the pillow lava, sheet flow, and picrite of the nonmetamorphic extrusive sequence of the Khoy ophiolite plot in the nonarc field (Fig. 7A). One pillow lava and one sheet flow sample plot in the transitional zone. All samples plot in the ocean floor field on the Ti/100-Zr-Y diagram (Fig. 7B; Pearce and Cann, 1973; Khalatbari Jafari et al., 2006). Samples from the diabase dikes that cut the extrusive sequence plot in the arc field, while lavas in the extrusive sequence plot in the MORB field of the V versus Ti/1000 diagram (Fig. 7C; Shervais, 1982; Khalatbari Jafari et al., 2006). All samples from the Khoy extrusive sequence plot in the “subduction-unrelated” field with MORB-array affinity on the Th/Yb versus Nb/Yb diagram (Fig. 7D; Shervais, 1982; Dilek and Furnes, 2011). The field for the continental crust is also delineated on this diagram to reveal the potential effect of partial melting of the subducted slab sediments and crustal contamination on magma generation. All samples plot in the MORB field with oceanic-island basalt (OIB mantle source) affinity on the Sm/Yb versus Nb/Yb diagram (Fig. 7E; Green, 2006). Figure 6F plots the samples on the (La/Sm)N against Nb/Yb diagram, normalized to chondrite (Sun and McDonough, 1989; Tian et al., 2011, and references therein). The Indian MORB and the Tonga arc are also shown to represent mid-ocean ridges and arcs, respectively. The field representing partial melting of pelagic sediments of the subducted slab is also shown. The short, dashed arrows indicate the effect of subduction components released as melt from the subducted slab. The samples from the Khoy extrusive sequence plot in the Indian MORB with OIB enrichment, away from the fields that indicate the effects of fluids or melt released from the subducted slab. The diabase dikes, on the other hand, plot in the Tonga field and clearly show the effect of melt released from the subducted slab.

Mid-ocean-ridge to suprasubduction geochemical transition in hypabyssal and extrusive sequences

239

Figure 6. Composition and geochemistry of pillow lava, sheet flow, and picrite in the extrusive sequence and crosscutting individual diabase dikes of the Khoy ophiolite. (A) Zr/TiO2 vs. Nb/Y classification diagram (Winchester and Floyd, 1976); (B) Th vs. Co magmatic trend diagram (Hastie et al., 2007); (C) chondrite-normalized rare earth element (REE) diagram showing the fields for the normal (N) and transitional (T) mid-ocean-ridge basalt (MORB; Sun and McDonough, 1989); (D) primitive mantle–normalized (Sun and McDonough, 1989) spider diagram (data from Khalatbari Jafari et al., 2006).

In summary, geochemical analyses indicate that the Upper Cretaceous extrusive ophiolite sequence of Khoy displays MORB characteristics, with REE patterns that represent transitional MORB (T-MORB). On the other hand, the patterns in the samples from the diabase dikes that cut the extrusive sequence show depletion in the HFSEs and subduction-zone geochemical characteristics (Arculus, 1994; Pefander et al., 2002; Juteau and Maury, 2012). Layered gabbro is cut by diabase dikes with T-MORB characteristics that are similar to those of the lavas in the extrusive sequence (Khalatbari Jafari et al., 2006). The crosscutting diabase dikes in the northern massif, on the other hand, display suprasubduction-zone characteristics. This indicates that, before the injection of the diabase dikes that cut the extrusive sequence, the Khoy ophiolite went through a major geodynamic transition from a MORB to a suprasubduction-zone environment.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084793/spe525-07.pdf by University of Wisconsin Madison user

Kermanshah Ophiolites The Kermanshah ophiolites (Figs. 1, 2, 3, and 8) are exposed along the northeastern edge of the 2000-km-long Zagros foldand-thrust belt, which stretches from Turkey to Hormuz Island in the Persian Gulf, and they are related to the geodynamic evolution (opening and closing) of the Neotethys Ocean (Leturmy and Robin, 2010). As remnants of the Neotethys, the ophiolites occur along the Zagros crush zone south of Sahneh and southeast-east of Harsin (Fig. 8) and are attributed to the peri-Arabic ophiolite belt (Ricou, 1994). The 1:250,000 scale Kermanshah (Braud, 1978, 1987) and 1:100,000 scale Harsin (Shahidi and Nazari, 1997) and Kermanshah (Karimi Bavandpour, 1999) geological maps show the Sahneh ophiolite in the north of the Kermanshah crush zone separated from the Harsin ophiolite to the south

240

Khalatbari Jafari et al.

Figure 7. Tectonomagmatic setting of pillow lava, sheet flow, and picrite in the extrusive sequence and crosscutting individual dikes of the Khoy ophiolite. (A) Nb/Th vs. Y diagram (Jenner et al., 1991); (B) ternary Zr-Ti/100-Y*3 diagram (Pearce and Cann, 1973); (C) V vs. Ti/1000 (Shervais, 1982); (D) Th/Yb vs. Nb/Yb diagram (Shervais, 1982) showing the normal (N) and enriched (E) mid-ocean-ridge basalt (MORB) and suprasubduction fields (Dilek and Furnes, 2011); (E) Sm/Yb vs. Nb/Yb diagram (Green, 2006) showing the mantle array, oceanic-island basalt (OIB), and MORB fields (Green, 2006); (F) chondrite-normalized (Sun and McDonough, 1989) (La/Sm)N vs. Nb/Yb diagram (Tian et al., 2011, and references therein; data from Khalatbari Jafari et al., 2006).

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084793/spe525-07.pdf by University of Wisconsin Madison user

Mid-ocean-ridge to suprasubduction geochemical transition in hypabyssal and extrusive sequences

Figure 8. Simplified map of the Kermanshah ophiolites (after 1:100,000 Harsin geological map, Shahidi and Nazari, 1997). Modifications are based on authors’ fieldwork.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084793/spe525-07.pdf by University of Wisconsin Madison user

241

242

Khalatbari Jafari et al.

through the carbonate Bisotun unit. Other exposures of ophiolitic massifs along the Kermanshah ophiolite occur in Miyanrahan, Kamyaran, Piyazeh (south Marivan), Sardasht, and Piranshar near the Iraqi border in Iran. Equivalent exposures occur around Mavat, Daraban, Penjwin, Bulfat, and Rayat across the border in Iraq, which discontinuously join the ophiolite belt in southern Turkey. The Neyriz ophiolite (Babaie et al., 2001, 2005, 2006) occurs in the southern part of the high Zagros in line with the Kermanshah ophiolites. The NW-SE–trending Zagros fold-and-thrust belt, covering the northeastern edge of the Arabian plate, involves Permian–Triassic to Upper Cretaceous and Paleocene shallow-marine sedimentary rocks, mostly limestone, that are covered by Paleocene–Pliocene sedimentary rocks (Stöcklin, 1968; Berberian and King, 1981). The Kermanshah and Neyriz are believed to be the most important ophiolites that occur along the suture zone between the Arabian plate and the Sanandaj-Sirjan zone (Stöcklin, 1968). While the Zagros fold-and-thrust belt lies along the passive margin of the Arabian plate, the Sanandaj-Sirjan zone lies on the active margin of the Iranian plate (Ricou, 1994). Rifting and drifting separated the Cimmerian blocks (including the Sanandaj-Sirjan zone) from Gondwana and assisted their accretion onto Eurasia (Şengör, 1990). Continued rifting during the Permian led to the spreading of the Neotethys Ocean in Oman and along the eastern margin of the Mediterranean Sea (Stampfli and Borel, 2002; Chauvet et al., 2009). The rifting and spreading of the Neotethys Ocean between the Arabian plate and Eurasia may have occurred during the Late Permian and Early Triassic (Robertson et al., 2007). The NE-dipping subduction of Neotethys under the Sanandaj-Sirjan zone could have started in the Late Jurassic (Stampfli and Borel, 2002), in the Jurassic, synchronous with the convergence of Africa and Eurasia (Omrani et al., 2008), or in the Mesozoic between 120 and 83 Ma (Rosenbaum et al., 2002). The Zagros crush zone in Kermanshah includes Tethyan ophiolitic units and thrust sheets between the Zagros fold-andthrust belt in the southwest and the Sanandaj-Sirjan zone in the northeast (Braud, 1978, 1987). In addition to the ophiolitic sheets, the crush zone includes other units, such as radiolarian chert, the Bisotun limestone, volcanic-sedimentary rocks, and postophiolitic intrusive bodies. The highly folded radiolarian cherts, interpreted to be continental margin sediments (Ricou et al., 1977; Braud, 1987), occur at the base of the ophiolitic and volcanic-sedimentary units. Paleontological studies of the radiolarian chert around the village of Gamasiab yielded early Pliensbachian to early Turonian ages (Gharib and De Wever, 2010). The igneous rocks in the Kermanshah area can be divided into two distinct zones (Braud, 1987). The first includes the Harsin ophiolitic units (Fig. 8) that extend to the southeast of Sahneh. This zone is dominated by peridotite that is intruded by gabbroic isolated dikes and dike swarms. Sheeted dikes are absent. The ophiolitic units in this zone were obducted during the Late Cretaceous onto the passive margin of the Arabian plate (Braud, 1987). The second zone extends from east of Sahneh to the north

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084793/spe525-07.pdf by University of Wisconsin Madison user

of Kamyaran, and it includes several sheared serpentinite thrust sheets. Lava flows occur with interlayers of Paleocene–middle Miocene sedimentary rocks. These units are covered by the Paleozoic and Mesozoic thrust sheets of the Sanandaj-Sirjan zone (Braud, 1987; Agard et al., 2011; Whitechurch et al., 2013). The Kermanshah ophiolites may be the remnants of a slow-spreading ocean that existed before the collision of the Arabian plate with Eurasia that were cut by diabase dikes with arc geochemical characteristics (Braud and Bellon, 1974; Braud, 1987). Ricou et al., (1977) inferred the Kermanshah and Neyriz ophiolites to be remnants of a mid-ocean ridge. Delaloye and Desmons (1980) reported an 81.4 Ma (Campanian) whole-rock K/Ar age for a sample from a dolerite dike with leucodiorite composition. These authors interpreted the Late Cretaceous age as the age of intrusion of the Kermanshah ophiolite by the dikes. Leterrier (1985) reported oceanic-island arc geochemical characteristics for the remnants of the ophiolite north of Kamyaran. They reported Paleocene–Eocene volcanic and volcanic-sedimentary units around Kamyaran that are cut by gabbroic to dioritic masses with Oligocene Rb-Sr isotopic ages. Whitechurch et al. (2013) also reported an Oligocene (26.5 Ma) K/Ar age for gabbro. The volcanic rocks of the Kermanshah ophiolite have subalkaline to alkaline compositions (Ghazi and Hassanipak, 1999). The gabbroic rocks of this ophiolite display E-MORB characteristics and were probably formed during the early stages of spreading of a mid-ocean ridge (Allahyari et al., 2010). Allahyari et al. (2010) also described lherzolite and residual harzburgite with enriched LREEs due to fluids in a suprasubduction zone. Wrobel-Daveau et al. (2010) documented evidence for detachment faults and mantle exhumation in the Zagros suture zone in the Kermanshah area. Allahyari et al. (2012) interpreted pillow lavas around the village of Tamerk, south of Harsin, to have E-MORB characteristics, and those in the Gashor Pass south of Harsin to have within-plate basalt characteristics. The basalt in these two localities is mantle derived and enriched in the large ion lithophile elements (LILEs) and LREEs. The model for the evolution of the Kermanshah ophiolite, presented by Saccani et al. (2013b), is based on the rifting model of Robertson et al. (2007) for the peri-Arabic ophiolite belt given for the southern Tethyan ophiolites, including Oman. According to Whitechurch et al. (2013), the Kermanshah ophiolites represent a Late Cretaceous–Paleocene backarc basin, and they were cut by Eocene arc rocks. U-Pb dating of the intermediate to felsic volcanic and granitic rocks in the Harsin area yielded a 95–94.6 Ma age, which is attributed to local collision of the Arabian plate and BisotunAvaraman block in the Late Cretaceous (Nouri et al., 2015). Figure 8 presents a simplified 1:100,000 geological map of Harsin (modified after Shahidi and Nazari, 1997) and a schematic stratigraphic column of the Kermanshah ophiolites in the metamorphic (Sahneh) and nonmetamorphic (Harsin) sections. The Sahneh ophiolite includes meta-peridotite of lherzolite type and clinopyroxene-bearing harzburgite, and it contains thin thrust sheets of meta-gabbro–amphibolite intrusions and meta-diabase to meta-gabbro dikes (Fig. 8). Petrographic analysis of the

Mid-ocean-ridge to suprasubduction geochemical transition in hypabyssal and extrusive sequences meta-peridotite revealed evidence for mantle deformation such as kink-banded olivine, stretched orthopyroxene, and wavy extinction in orthopyroxene and clinopyroxene. Undeformed orthopyroxene and clinopyroxene within the deformed minerals may be the products of transition melting. Evidence for the melt can be found in the melting and removal of the edges of the deformed orthopyroxene phenocrysts, which are filled with fine-grained, unkinked olivine neocrystals. Fine-grained neocrystals of orthopyroxene and clinopyroxene also represent crystallization from the melt. Layering in the clinopyroxene-bearing meta-harzburgite is oriented about N-S in the area northeast of the village of Sarassiab. A thrust sheet of meta-gabbro–amphibolite, bounded by serpentinite, occurs along the Gamasiab River around Khalagrud (Fig. 8). Both contacts of this thrust sheet are tectonized and contain slices of Triassic marble, dark shale, and Jurassic–Cretaceous metavolcanic rocks. Although the thrust sheet is ductilely deformed, evidence for original magmatic layering can be found as ribbons of dark and light minerals, superposed by a metamorphic foliation. Rotated clinopyroxene porphyroclasts and zoned plagioclase represent original igneous minerals. Except for remnants of clinopyroxene, most dark bands are green neoformed hornblende, tremolite-actinolite, and epidote. In addition to sparse plagioclase, which is mostly altered, the light bands include clay minerals, epidote hydrogrossular after Ca-rich plagioclase, and albite neocrystals. Rare sphene also occurs along the foliation. The mineralogy suggests metamorphism of the original gabbro to the greenschist to amphibolite facies. Light bands of rodingite along the foliation may represent original anorthosite. Isolated rodingitized metadiabase dikes cut across the meta-peridotite. The serpentinite has a tectonized contact with the meta-gabbro–amphibolite and is cut by individual, more or less rodingitized diabase dikes. Small sills and dikes of plagiogranite also occur in the meta-gabbro and amphibolite. South and north of Dehlor, bodies of meta-andesite have tectonized contacts with the Triassic marble and Cretaceous carbonate, similar to the metagabbro-amphibolite thrust sheets (Fig. 8). Determining whether they belong to the Sahneh ophiolite requires further fieldwork. The nonmetamorphic Harsin ophiolite extends as a thrust sheet from around Harsin to the village of Noorabad. At the base, the thrust sheet includes tectonized ophiolitic units and radiolarian chert, and at the top, above the hyaloclastic and epiclastic breccia, it is covered by Miocene conglomerate and limestone (Fig. 9). The ophiolite occurs as serpentinite peridotite in the core of a tectonic window near Gashor Pass, where the peridotite is unconformably overlain by Miocene limestone with a basal conglomerate that is sheared along with the ophiolitic units at its lower few meters. The ophiolitic units include serpentinized peridotite, cumulate layered gabbro, and the extrusive sequence (Figs. 8 and 9). The mantle sequence of the Harsin ophiolite includes harzburgite and serpentinized clinopyroxene-bearing harzburgite cut by individual diabase dikes. Compared to the Sahneh ophiolite, evidence for mantle deformation is scarce in the Harsin ophiolite, where foliation and lineation are absent, and cumu-

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084793/spe525-07.pdf by University of Wisconsin Madison user

243

late fabric and composition correlate with the upper mantle and asthenosphere. The cumulate gabbro is exposed near the villages of Paien ab and Shah-abad (Fig. 8) and includes olivine gabbro, anorthosite, coarse-grained pyroxene gabbro, and fine layers of wehrlite, all cut by plagiogranite (Fig. 8). The extrusive sequence of the Harsin ophiolite is exposed east of the village of Tamerk in the Gashor Pass, and sporadically, in the Harsin area (Fig. 8). The base of this sequence and its contact with peridotite are tectonized and start with porphyritic pillow lava and sheet flow, which grade upward to vesicular, phyric and aphyric lava (Fig. 9). Paleontological analysis of the sponge and spicules of the radiolarids (Spumellaria and Nessellaria) yielded a late Mesozoic (probably Cretaceous) age (Geological Survey and Mining Exploration of Iran, unpublished internal report, Sara Soleimani, 2014, personal commun.). U-Pb dating of the core of the zircons from rudingitized gabbro within the serpentinized peridotite of Shahabad village, southeast Harsin, revealed a 79.3 Ma age (Ao et al., 2015, with contribution of the first author). This age is interpreted as the best estimate of the crystallization age of the gabbro in the Harsin ophiolite (Ao et al., 2015). Hyaloclastic and epiclastic breccia, overthrust by serpentinite, is exposed southeast of the village of Tamerk. Figure 10 displays the classification, magmatic series discrimination, REE patterns, and spider diagrams for the extrusive sequence of Harsin (chemical results from Allahyari et al., 2012). Figure 10A shows the Zr/TiO2 versus Nb/Y classification diagram (Winchester and Floyd, 1976) of the samples taken from the Kermanshah-Harsin ophiolite. Samples from the pillow lava of the Tamerk and Gashor Pass areas plot in the andesitebasalt and alkali basalt fields, respectively. The Tamerk and Gashor Pass samples plot in the calc-alkaline and high-K shoshonite fields, respectively, on the Th versus Co classification and magmatic series discrimination diagram (Fig. 10B; Hastie et al., 2007). The Gashor Pass samples are more enriched in the LREEs compared to those in the Tamerk samples, and they display P-MORB patterns, compared to the E-MORB patterns of the Tamerk pillow lava (Fig. 10C). The primitive mantle– normalized (Sun and McDonough, 1989) spider diagrams of the REEs and trace elements of the samples taken from the extrusive sequence of the Kermanshah ophiolite display no depletion of Nb, Zr, and Ti in the HFSEs (Fig. 10D). The enrichment of the Gashor Pass pillow lava may be attributed to an enriched and garnet-bearing mantle source. These patterns indicate that the extrusive sequence of the Kermanshah ophiolite does not have the suprasubduction-zone characteristics, and their magmatic evolution relates to mantle plumes. The Tamerk pillow lava samples, taken from the extrusive sequence of the Harsin ophiolite, plot in the shoshonite with enriched mantle source field on the Th/Yb versus Ta/Yb tectonomagmatic diagram (Fig. 11A; Pearce, 1982, 2003; Faustino et al., 2006). The pillow lavas plot in the ocean floor basalt field in the Zr-Ti/100-Y*3 ternary diagram (Fig. 11B) of Pearce and Cann (1973). The pillow lavas of the Tamerk and Gashor Pass areas plot in the MORB and the more-enriched ocean-island and

244

Khalatbari Jafari et al.

Figure 9. Schematic stratigraphic column of the Kermanshah ophiolites reconstructed from the remnants of the Harsin and Sahneh metamorphosed ophiolites (cpx—clinopyroxene).

alkali basalt fields, respectively, on the V versus Ti/1000 diagram (Fig. 11C) of Shervais (1982). Samples from the Tamerk pillow lava plot in the field that is unrelated to subduction, and those from Gashor Pass plot in the more-enriched field of OIBs in the Th/Yb versus Nb/Yb diagram (Fig. 11D; Shervais, 1982; Dilek and Furnes, 2011). The Tamerk and Gashor Pass pillow samples plot far from the average MORB near the OIB field on the Sm/ Yb versus Nb/Yb diagram (Fig. 11E; Green, 2006), and in the Indian MORB and OIB enrichment fields, respectively, on the chondrite-normalized (Sun and McDonough, 1989) (La/Sm)N

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084793/spe525-07.pdf by University of Wisconsin Madison user

versus Nb/Yb diagram (Fig. 11F; Tian et al., 2011, and references therein). No evidence of the effect of subduction components such as melt or fluid is observed for these samples. Ophiolitic units that occur along the Zagros thrust zone in the Piyazeh (Sawlava, south Marivan) area in Iranian Kurdistan include cumulate lherzolite with dunite and chromitite lenses (Sabzehei et al., 2009) that are assumed to have formed as a result of crystallization from a boninitic melt and partial melting of depleted peridotite (Allahyari et a., 2014). The Mavat ophiolite along the Zagros trend in Iraq is assumed to represent

Mid-ocean-ridge to suprasubduction geochemical transition in hypabyssal and extrusive sequences

245

Figure 10. Composition and geochemistry of pillow lava in the extrusive sequence of the Harsin (Kermanshah) ophiolite. (A) Zr/TiO2 vs. Nb/Y classification diagram (Winchester and Floyd, 1976); (B) Th vs. Co magmatic trend diagram (Hastie et al., 2007); (C) chondritenormalized rare earth element (REE) diagram showing the fields for the normal (N) and transitional (T) mid-ocean-ridge basalt (MORB; Sun and McDonough, 1989); (D) primitive mantle–normalized (Sun and McDonough, 1989) spider diagram.

a suprasubduction zone (Ismail et al., 2010). Leucogranitic dikes that cut the serpentinite Mavat ophiolite are peraluminous and resemble those in collision zones (Mohammad et al., 2013). K-Ar dating of the muscovite from the leucogranitic dikes yielded a 37.57 Ma age, which correlates with the age of the leucogranitic dikes that cut gabbro in the Kamyaran ophiolite (Ao et al., 2015). Mohammad et al. (2013) attributed this age to the collision of the Eurasian and Arabian continental plates during the opening of the Gulf of Eden ca. 37 Ma. Two parallel, NE-SW–trending Late Cretaceous ophiolite belts occur in southeast Turkey. The northern belt correlates with the Khoy ophiolite in northwest Iran, and the southern belt correlates with the Kermanshah and other isolated ophiolites along the northeastern edge of the Zagros fold-and-thrust belt, such as the Kamyaran, Miyanrahan, Piyazeh, Sardasht, and Piranshahr ophiolitic massifs. The southern ophiolite occurs between the BitlisPütürge metamorphic massifs and the Arabian platform (Yilmaz et al., 1993; Yiğtibas and Yilmaz, 1996; Robertson, 2002; Rob-

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084793/spe525-07.pdf by University of Wisconsin Madison user

ertson et al., 2012, and references therein) and includes the Troodos, Baer-Bassit (Syria), and Kizildağ ophiolites and the Koçali mélange, which formed in the southern branch of the Neotethys Ocean (Robertson et al., 2012, and references therein; Parlak et al., 2009; Dilek and Thy, 2009). The formation of the southern belt is attributed to spreading above a north-dipping, intraoceanic subduction zone (Al-Riyami et al., 2002; Bağci et al., 2006, 2008; Rızaoğlu et al., 2006; Parlak et al., 2009). The Harsin and Sahneh ophiolites probably correlate with the southern belt ophiolites in the southeast Anatolian region in Turkey (e.g., Yilmaz et al., 1993; Robertson et al., 2012). Fannuj Ophiolite The Makran ophiolites in the Sistan-Baluchistan province in southeast Iran are exposed along the southern edge of the Jaz Murian depression near Iranshahr in the east, and Bandare-Abbas in the west (Figs. 1, 2, 3, 12, and 13). These ophiolites

246

Khalatbari Jafari et al.

Figure 11. Tectonomagmatic setting of the extrusive sequence of the Harsin (Kermanshah) ophiolite. (A) Th/Yb vs. Ta/Yb diagram (Pearce, 1982, 2003) showing the tholeiitic and calc-alkaline fields (Faustino et al., 2006); (B) ternary Zr-Ti/100-Y*3 diagram (Pearce and Cann, 1973); (C) V vs. Ti/1000 (Shervais, 1982); (D) Th/Yb vs. Nb/Yb diagram (Shervais, 1982) showing the normal (N) and enriched (E) midocean-ridge basalt (MORB) and suprasubduction zone (SSZ) fields (Dilek and Furnes, 2011); (E) Sm/Yb vs. Nb/Yb diagram (Green, 2006) showing the mantle array, oceanic-island basalt (OIB), and MORB fields (Green, 2006); (F) chondrite-normalized (Sun and McDonough, 1989) (La/Sm)N versus Nb/Yb diagram (Tian et al., 2011, and references therein).

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084793/spe525-07.pdf by University of Wisconsin Madison user

Mid-ocean-ridge to suprasubduction geochemical transition in hypabyssal and extrusive sequences are known as Iranshahr, Fannuj-Maskutan, Deyadaer, RemeshkMokhtarabad, Ganj, Dar Anar, and Band-e-Zeyarat complexes (Fig. 12). The elongate, lens-shaped, E-W–trending Fannuj ophiolite complex, which extends over 2800 km2 south of Jaz Murian between Remeshk-Mokhtarabad in the west and Iranshahr ophiolite complex in the east (Fig. 12), is one of the largest and most complete ophiolites in Iran. The ophiolite is part of the Makran-Zahedan group (group 4) defined by Knipper et al. (1986) (Fig. 1) and belongs to the Mesozoic Neotethyan AlpineHimalayan ophiolites (Desmons and Beccaluva, 1983). The Makran includes the Jaz Murian depression in the north as a subsiding backarc basin, and an uplifted block in the south that extends from the Beshagard Mountains to the deepest part of the Gulf of Oman (Farhoudi and Karig, 1977). Arshadi and Forster (1983) defined two northern and southern geotectonic zones in the Makran area that are separated by a narrow continental ridge. These authors presented a stratigraphic column for the Makran ophiolites that includes peridotite-serpentinite,

ultramafic-cumulate mafic, gabbro and its leucocratic associates, diabase sheeted dikes, pillow lava, and pelagic sediments. Paleontological data (Arshadi and Forster, 1983; Eftekharnezhad, 1987) from the pelagic sediments in between the pillows assign an Early Cretaceous–Paleocene age for the Fannuj ophiolite. The Remeshk ophiolite complex includes plutonic ophiolitic rocks and, to a lesser extent, diabase sheeted dikes at the top. The Mokhtarabad ophiolite complex includes pillow lava, pelagic sediment, turbidite, and diabase dikes. Paleontological study of carbonates in the Mokhtarabad ophiolite by McCall (1985) narrows the complex’s age to the Late Cretaceous, and farther west to Paleocene. Glennie et al. (1990) related the Makran to a Late Jurassic–Early Cretaceous intracontinental splitting of the Iranian platform. Based on geochemical data from igneous rocks and samples from the Makran ophiolite, Shahabpour (2010) presented a tectonomagmatic model that relates the ophiolite to a backarc basin. Desmons and Beccaluva (1983) studied the crustal extrusive sequence of the Fannuj-Maskutan and

Figure 12. Generalized map of the Makran ophiolites (Moslempour, 2011).

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084793/spe525-07.pdf by University of Wisconsin Madison user

247

248

Khalatbari Jafari et al.

Figure 13. Simplified geological map of the Fannuj ophiolite (modified from Eftekharnezhad, 1987). Rose diagrams show the trend of the diabase sheeted dikes in five stations where dikes are well exposed.

related their formation to a mid-ocean ridge with MORB affinity, although they did not rule out formation in a backarc basin. Moslempour (2011) associated the peridotite of the Fannuj-Maskutan ophiolite to abyssal and suprasubduction-zone settings. Hunziker et al. (2011) studied the igneous and metamorphic rocks of the Fannuj-Maskutan and Remeshk-Mokhtarabad, and related them to the suprasubduction-zone ophiolites. The Fannuj ophiolite includes crustal and mantle sequences (Figs. 13 and 14). The mantle sequence includes lherzolite and porphyroclastic clinopyroxene-bearing harzburgite, fine-grained lherzolite with neocrystals, and scattered lenses of dunite and

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084793/spe525-07.pdf by University of Wisconsin Madison user

chromitite. Despite serpentinization, these units display evidence for mantle and asthenosphere textures. The peridotite in the Fannuj ophiolite can be divided into two groups. The first group lies at the base of the mantle sequence and extends from north of Fannuj to south of the village of Aparang (Figs. 13 and 14). This group includes porphyroclastic clinopyroxene-bearing harzburgite and porphyroclastic lherzolite and displays textural evidence for mantle deformation. These include high-temperature, plastic porphyroclastic texture and mantle foliation. The mantle deformation is evidenced by kinked porphyroclastic olivine, stretched porphyritic enstatite, and rotated clinopyroxene porphyroclasts

Mid-ocean-ridge to suprasubduction geochemical transition in hypabyssal and extrusive sequences

249

Figure 14. Schematic stratigraphic column showing the main units of the Fannuj ophiolite (cpx—clinopyroxene).

that match those interpreted as evidence for mantle deformation in ophiolites (Nicolas, 1989; Juteau and Maury, 2012). The peridotite of the second group is exposed as small outcrops to the south, east, and west of the town of Fannuj (Fig. 13), and it displays cumulate textural characteristics with scarce evidence for mantle deformation. The second group of peridotite is cut by isolated gabbroic and diabase dikes and features that are common in asthenospheric peridotite, such as boudinaged pyroxenite sills and dikes with varying thicknesses (Marchesi, 2006; Suhr et al., 2008; Dijkstra et al., 2010). Plagioclase, clinopyroxene, and occasionally olivine veins and veinlets also cut the peridotite. The gabbroic-diabase dikes vary in thickness from 1 cm to 1 m and strike about N-S, dipping near vertical north of Fan-

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084793/spe525-07.pdf by University of Wisconsin Madison user

nuj. Small and scattered lenses of chromitite occur, with serpentinite cover, in the residual peridotite south of the village of Aparang. It is possible that the chromitite-bearing serpentinized peridotite were originally dunite. Pyroxenite veins and veinlets cut across both groups of peridotite and cumulate layered gabbro. Extensive exposures of cumulate layered gabbro, as much as 700 m thick, are exposed east of Fannuj, west of the villages of Gormazar and Gogaz (Fig. 13). The cumulate layered gabbro lies above the asthenospheric peridotite, under massive gabbro, and forms the largest bodies of cumulate layered gabbro in Iran. The layered gabbro includes olivine gabbro, troctolite, pyroxene gabbro, microgabbro, leucogabbro, and anorthosite. West of the villages of Gormazar and Gogaz, wehrlitic intrusions cut across

250

Khalatbari Jafari et al.

the cumulate layered gabbro. Isotropic gabbro is exposed east of the Gogaz, Gormazar, and Fannuj ophiolites, around the village of Maskutan (Fig. 13). East of Gogaz, massive gabbro lies gradationally above the cumulate layered gabbro. The massive gabbro is altered and lacks, or has weak, layering. The isotropic gabbro includes pyroxene gabbro, coarse-grained gabbro, pegmatite gabbro, and leucogabbro, and it is cut by veins full of epidote. Leucocratic quartzofeldspathic veins and dikes, and small pockets of plagiogranite-trondhjemite, cut across the isotropic gabbro. Exposures of diabase sheeted dikes as wide as a few hundred meters in the Fannuj area are the thickest among Iranian ophiolites, and they form spectacular exposures along the northnorthwest segments of the Maskutan-Fannuj road (Figs. 13 and 14). As was mentioned already, the contact of the isotropic gabbro with the diabase sheeted dikes to the east and south of Gogaz is gradational. The upper contact of the sheeted dikes with pillow lavas of the extrusive sequence, exposed along the Fannuj-Maskutan road, is also gradational, and it is characterized by a mixture of dike–pillow lava with extensive zones of altered sulfides along the contact. The sheeted dikes strike about N-S (Fig. 13), which may represent the trend of the opening of the Neotethys. The thickness of the dikes varies from a few centimeters to less than a meter, and their margin is finer grained than their cores. The composition of the sheeted dikes is dominantly diabase, with epidote, chlorite, and minor calcite also present due to alteration. Sporadic dikes and veins of epidote, quartz, and calcite cut the sheeted dikes. The extrusive sequence of the Fannuj ophiolite is mostly exposed to the east and south of Maskutan and Petkan (Fig. 13), and it includes a sequence of pillow lava, sheet flow, hyaloclastic breccia, and interlayers of pelagic limestone, radiolarite, and chert that are cut by diabase dikes. On the top, the sequence is covered across an angular unconformity by Paleocene–Eocene flysch (Fig. 13). Pillow lavas are moderately phyric to phyric, with the latter only seen at the base of the sequence. Vesicles are common in the phyric pillows, and they are filled with smectite, chlorite, calcite, and epidote. Sheet flows are more extensively exposed than the pillow lavas in the extrusive sequence. Brass-colored dacitic-rhyolitic pillow lava is exposed along the Maskutan-Fannuj road and forms high weathering outcrops compared to andesitic-basaltic pillow lava. Paleontological study of pelagic limestone interbeds among the pillows yielded Early– Late Cretaceous Textularia sp., Acicularia sp., and Cuneolina sp., which confirm earlier ages suggested by McCall (1985) and Eftekharnezhad (1987). Hyaloclastic breccia occurs as interbeds or interstitial fills among the pillows. Table 1 lists the results of geochemical analyses of selected samples of the extrusive sequence of the Fannuj ophiolite. Samples of diabase sheeted and individual dikes plot in the basaltic and andesitic fields on the Zr/TiO2 versus Nb/Y classification diagram (Fig. 15A) of Winchester and Floyd (1976). The data from the pillow lavas of the extrusive sequence cluster in two separate basalt-andesite and rhyodacite-rhyolite assemblages on this and other types of diagrams. The rhyodacite-rhyolite of the

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2065/chapter-pdf/4084793/spe525-07.pdf by University of Wisconsin Madison user

second group suggests an evolved magma source for their origin. Samples from the diabase sheeted and individual dikes, and some pillow lavas, plot in the tholeiitic field on the Th versus Co magmatic series discrimination diagram (Hastie et al., 2007), with the rest of the pillows plotting in the calc-alkaline high-K shoshonite (Fig. 15B). The chondrite-normalized REE patterns (Sun and McDonough, 1989) also reveal two groups. The first group is less enriched (15–40 times chondrite) compared to the second group, which is 65–100 times chondrite. Sample data for the first and second group plot in the forearc-backarc and the backarc fields, respectively (Fig. 15C), of Nicholson et al. (2000). The wide range of enrichment (15–100 times chondrite) may reflect different rates of partial melting and a heterogeneous mantle source for the Fannuj pillow lavas. The enriched multi-elemental chondrite-normalized REE patterns of the pillow lavas (Fig. 15C) may reflect the role of partial melting in the enrichment. The LREE patterns for the pillow lava, sheeted dikes, and individual diabase dikes, which gently decrease toward the HREEs, may represent the enrichment of the LREEs by fluid components that were derived from the subducted slab. The effect of the subducted slab fluids is more apparent on patterns in the first group of pillow lavas and sheeted and individual dikes. The formation of the rhyodacite-rhyolite (second group) composition in the extrusive sequence of the ophiolite can be attributed to the partial melting of the sedimentary cover of the subducted slab, and the effect of the melt product on the mantle wedge above the subducting slab during the formation of the Fannuj extrusive sequence—a hypothesis that was proposed for the Miridita and Pindos ophiolites in Albania and Greece, respectively (Dilek and Furnes, 2009). The primitive mantle– normalized (Sun and McDonough, 1989) patterns of the Fannuj extrusive sequence samples show depletion of Ta, Nb, and Ti among the HFSEs, and enrichment of Th and U (Fig. 15C), similar to the patterns for subduction zones. Most samples from the diabase sheeted dikes of the extrusive sequence plot in the arc field, with a few in the nonarc field, on the Nb/Th versus Y tectonomagmatic diagram (Fig. 16A; Jenner et al., 1991). The majority of the samples plot in the island-arc tholeiite field, with a few that plot in the N-MORB and calc-alkaline basalt fields on the Th-Hf/3-Nb/16 ternary diagram (Fig. 16B; Wood, 1980). The samples scatter in the MORB field on the V versus Ti/1000 diagram (Fig. 16C; Shervais, 1982) and range between the subduction-related and subduction-unrelated fields on the Th/Yb versus Nb/Yb diagram (Fig. 16D; Shervais, 1982; Dilek and Furnes, 2011). Most samples display a depleted MORB source pattern on the Th/Yb versus Ta/Yb diagram (Pearce, 1982), while some plot in the suprasubduction-zone field between tholeiitic and calc-alkaline series (Fig. 16E). The chondrite-normalized (La/Sm)N versus Nb/Yb diagram (Fig. 16F; Tian et al. 2011, and references therein) shows clustering of the samples away from the Indian MORB field and the effect of subduction components on the melt. It seems that the melt produced by the partial melting of the subducting slab played a significant role in the formation of the extrusive sequence of the Fannuj ophiolite.

Fan 33 48.7 1.79 15.8 11 0.22 7.8 7.09 3.5 0.79 0.23 0.03 2.12 99.07

dia

dia

idd

Fan 86 Fan 128 Fan 160 Fan 130 46.2 47.6 47.8 45.8 1.15 1.51 0.96 1.96 15.9 16.5 19.2 14.6 10.6 10 7.87 12.4 0.18 0.16 0.19 0.15 8.79 6 7.34 6.97 10.7 10.1 12.2 10.9 2.9 2.9 2.4 3.2 0.28 0.76 0.14 0.18 0.1 0.16 0.1 0.22 0.05 0.03 0.05 0.04 1.9 2.71 1.48 1.82 98.7 98.43 99.73 98.24

dia Fan 37 46.3 1.23 15.1 9.83 0.15 11.4 10.6 2.6 0.13 0.11 0.08 0.65 98.18

idd Fan 79 48.3 1.51 15.4 8.6 0.12 4.53 13.3 3.6 0.51 0.21 0.03 2.26 98.37

pl Fan 23 59.3 1.39 14 8.16 0.08 2.16 4.36 5.5 0.61 0.31