Art and Architecture in Ladakh: Cross-Cultural Transmissions in the Himalayas and Karakoram [Illustrated] 9004271783, 9789004271784

Art and Architecture in Ladakh shows how the region's cultural development has been influenced by its location acro

577 79 42MB

English Pages 438 [461] Year 2014

Report DMCA / Copyright

DOWNLOAD FILE

Polecaj historie

Art and Architecture in Ladakh: Cross-Cultural Transmissions in the Himalayas and Karakoram [Illustrated]
 9004271783, 9789004271784

Table of contents :
Contents
List of Contributors
List of Illustrations
Introduction
Ancient Petroglyphs of Ladakh: New Discoveries and Documentation
Embedded in Stone—Early Buddhist Rock Art of Ladakh
Historic Ruins in the Gya Valley, Eastern Ladakh, and a Consideration of their Relationship to the History of Ladakh and Maryul
An Archaeological Account of Ten Ancient Painted Chortens in Ladakh and Zanskar
The Chorten (mChod rten) with the Secret Chamber near Nyarma
The Dating of the Sumtsek Temple at Alchi
The Iconography and the Historical Context of the Drinking Scene in the Dukhang at Alchi, Ladakh
The Wood Carvings of Lachuse. A Hidden Jewel of Early Mediaeval Ladakhi Art
The mGon khang of dPe thub (Spituk): A Rare Example of 15th Century Tibetan Painting from Ladakh
Chigtan Castle and Mosque: A Preliminary Historical and Architectural Analysis
Lamayuru (Ladakh)—Chenrezik Lhakhang: The Bar Do Thos Grol Illustrated As A Mural Painting
The Lost Paintings of Kesar
Tshogs zhing: A Wall Painting in the New ’Du khang of Spituk (dPe thub)
From Benaras to Leh—The Trade and Use of Silk-brocade
Conservation of Leh Old Town—Concepts and Challenges
Revealing Τraditions in Εarthen Αrchitecture: Analysis of Εarthen Βuilding Μaterial and Τraditional Constructions in the Western Himalayas
Conservation of Architectural Heritage in Ladakh
Bibliography
Index

Citation preview

Art and Architecture in Ladakh

Brill’s Tibetan Studies Library Edited by Henk Blezer Alex McKay Charles Ramble

VOLUME 35

The titles published in this series are listed at brill.com/btsl

Art and Architecture in Ladakh Cross-Cultural Transmissions in the Himalayas and Karakoram Edited by

Erberto Lo Bue and John Bray

LEIDEN | BOSTON

Cover illustration: Chigtan castle in 1909. Photo: Babu Pindi Lal. Courtesy Kern Institute, Leiden University, ms nr.XI. fol.31. Library of Congress Cataloging-in-Publication Data Art and architecture in Ladakh : cross-cultural transmissions in the Himalayas and Karakoram / edited by Erberto Lo Bue and John Bray.   pages cm. — (Brill’s Tibetan studies library ; Volume 35)  Includes bibliographical references and index.  ISBN 978-90-04-27178-4 (hardback) — ISBN 978-90-04-27180-7 (e-book) 1. Art—India—Ladakh. 2. Architecture—India—Ladakh. 3. Ladakh (India)—Civilization. I. Lo Bue, Erberto F., editor of compilation. II. Bray, John, 1957– editor of compilation.  N7307.L33A78 2014  709.54’6—dc23 2014006983

This publication has been typeset in the multilingual ‘Brill’ typeface. With over 5,100 characters covering Latin, ipa, Greek, and Cyrillic, this typeface is especially suitable for use in the humanities. For more information, please see brill.com/brill-typeface. issn 1568-6183 isbn 978 90 0427178 4 (hardback) isbn 978 90 0427180 7 (e-book) Copyright 2014 by Koninklijke Brill nv, Leiden, The Netherlands. Koninklijke Brill nv incorporates the imprints Brill, Brill Nijhoff, Global Oriental and Hotei Publishing. All rights reserved. No part of this publication may be reproduced, translated, stored in a retrieval system, or transmitted in any form or by any means, electronic, mechanical, photocopying, recording or otherwise, without prior written permission from the publisher. Authorization to photocopy items for internal or personal use is granted by Koninklijke Brill nv provided that the appropriate fees are paid directly to The Copyright Clearance Center, 222 Rosewood Drive, Suite 910, Danvers, ma 01923, usa. Fees are subject to change. This book is printed on acid-free paper.

Contents List of Contributors  vii List of Illustrations  xii Introduction  1 Erberto Lo Bue and John Bray 1 Ancient Petroglyphs of Ladakh: New Discoveries and Documentation  15 Tashi Ldawa Thsangspa 2 Embedded in Stone—Early Buddhist Rock Art of Ladakh  35 Phuntsog Dorjay 3 Historic Ruins in the Gya Valley, Eastern Ladakh, and a Consideration of Their Relationship to the History of Ladakh and Maryul With an Appendix on the War of Tsede (rTse lde) of Guge in 1083 CE by Philip Denwood  68 Neil and Kath Howard 4 An Archaeological Account of Ten Ancient Painted Chortens in Ladakh and Zanskar  100 Quentin Devers, Laurianne Bruneau and Martin Vernier 5 The Chorten (mChod rten) with the Secret Chamber near Nyarma  141 Gerald Kozicz 6 The Dating of the Sumtsek Temple at Alchi  159 Philip Denwood 7 The Iconography and the Historical Context of the Drinking Scene in the Dukhang at Alchi, Ladakh  167 Marjo Alafouzo 8 The Wood Carvings of Lachuse. A Hidden Jewel of Early Mediaeval Ladakhi Art  191 Heinrich Poell 9 The mGon khang of dPe thub (Spituk): A Rare Example of 15th Century Tibetan Painting from Ladakh  226 Chiara Bellini 10 Chigtan Castle and Mosque: A Preliminary Historical and Architectural Analysis  254 Kacho Mumtaz Ali Khan, John Bray, Quentin Devers and Martin Vernier

vi

contents

11 Lamayuru (Ladakh)—Chenrezik Lhakang: The Bar Do Thos Grol Illustrated as a Mural Painting  274 Kristin Blancke 12 The Lost Paintings of Kesar  298 John Bray 13 Tshogs zhing: a Wall Painting in the New ’Du khang of Spituk (dPe thub)  314 Filippo Lunardo 14 From Benaras to Leh—The Trade and Use of Silk-brocade  329 Monisha Ahmed 15 Conservation of Leh Old Town—Concepts and Challenges  348 André Alexander and Andreas Catanese 16 Revealing Traditions in Earthen Architecture: Analysis of Earthen Building Material and Traditional Constructions in the Western Himalayas  364 Hubert Feiglstorfer 17 Conservation of Architectural Heritage in Ladakh  388 John Harrison Bibliography  400 Index  428

List of Contributors Editors Erberto Lo Bue obtained a Ph.D in Tibetan Studies at the School of Oriental and African Studies (University of London) and became Associate Professor at Bologna University, where he taught the history of Indian and Central Asian art, and classical Tibetan. From 1972 he carried out research in Nepal, India and Tibet, his fieldwork in Ladakh starting in 1978 and continuing in 2001, 2002, 2003 and 2005. Most of his over 190 publications are related to Tibetan, Newar and Indian religious art. John Bray is President of the International Association for Ladakh Studies (IALS). His main research interests include the history of Christian missions on the borders of Tibet, and the 19th and early 20th politics of the Himalayan border regions. His publications include: A Bibliography of Ladakh (Aris & Phillips 1988); an edited volume, Ladakhi Histories (Brill, 2005); Mountains Monasteries and Mosques (co-edited with Elena de Rossi Filibeck; Rome, 2009); and articles in The Tibet Journal, Zentralasiatische Studien, and the Journal of the Royal Asiatic Society. Authors Monisha Ahmed is an independent researcher. Her doctoral thesis from Oxford University developed into the book Living Fabric—Weaving among the Nomads of Ladakh Himalaya (Orchid Press, 2002), which received the Textile Society of America’s 2003 R.L. Shep Award. She is the author of several articles on the material culture of Ladakh. Together with Clare Harris, she co-edited Ladakh—Culture at the Crossroads (Marg Publications, 2005) and with Janet Rizvi authored Pashmina—The Kashmir Shawl and Beyond (Marg Publications, 2009). She is the co-founder and Executive Director of LAMO (Ladakh Arts and Media Organisation) and Associate Editor for Marg Publications, Mumbai.

viii

list of contributors

Marjo Alafouzo obtained her Ph.D in 2008 from the School of Oriental and African Studies, University of London, where she has been teaching for a number of years. Her current research is on pre-14th century Ladakhi murals and their iconography. André Alexander (1965–2012) was the co-founder of the Tibet Heritage Fund (THF). His research work included studies on historic buildings in Lhasa (1993–2000), historic monuments in Sichuan (2001), vernacular Hutong architecture in Beijing (2001-2004), and historic monuments in Ladakh (2003–2012). As an architect he led conservation projects in Lhasa as well as Leh Old Town, and he designed the Central Asian Museum in Leh. André received his Master of Science and Ph.D degrees from the Berlin University of Technology. Chiara Bellini is a Research Fellow at the Department of Linguistic and Oriental Studies at the University of Bologna. She has studied the history of Indian and Tibetan art since 1999, focusing on wall painting, and carrying out fieldwork in Ladakh, Nepal (including Mustang) and Tibet. She obtained her Ph.D in Indology and Tibetology at the University of Turin with a thesis on 14th–16th century Ladakhi murals. In 2011 she received a scholarship for a research project on Alphabetum Tibetanum, written in 1762 by Agostino Giorgi. She is currently preparing a monograph on the history of Buddhist art in Ladakh in collaboration with Erberto Lo Bue. Kristin Blancke graduated as a clinical psychologist at the University of Leuven, Belgium. She studied Vajrayana Buddhism at Sherabling Monastery, Himachal Pradesh, India, and in the process started translating “The Hundred Thousand Songs of Milarepa” from Tibetan into Italian. She now lives in Dharamsala and works as a travel consultant specializing in trips to Buddhist areas. Her visits to Ladakh have led her to conduct research as an independent scholar in the field of iconography. Laurianne Bruneau dedicated her Ph.d to the study of the rock art of Ladakh from the Bronze Age to the Buddhist period. She has conducted post-doctoral research on rock and metal images relating to the Eurasian animal style from Ladakh, Gilgit-Baltistan and Rudog. Since September 2012 she has been Assistant

list of contributors

ix

Professor in Indian and Central Asian Studies at the Historical and Philological Department of the École Pratique des Hautes Études in Paris. Andreas Catanese worked for the Tibet Heritage Fund (THF) on restoration projects in Ladakh as well as in Mongolia and Amdo (Tibet) from 2005 to 2008. In Ladakh he made a deep study of the local culture and architectural style. Together with the THF team and with local artisans he tried to improve traditional techniques and materials, adapting them to changing climatic conditions and the new needs of the buildings’ users. Since 2009 he has worked as an architect in Switzerland but retains his interest in Ladakh. Philip Denwood is Emeritus Reader in Tibetan Studies, School of Oriental and African Studies, University of London, and is currently a Senior Teaching Fellow there. His interests include the Tibetan language, epigraphy, art history and architecture. His publications include The Tibetan Carpet (Aris & Phillips 1974) and Tibetan (Benjamins 1999). He spent six months in Ladakh in 1975 studying the Ladakhi dialect and the inscriptions of Alchi monastery. Quentin Devers is a Ph.D candidate in archaeology at the École Pratique des Hautes Études (Paris). His researches focus on the fortified sites of Ladakh (watch towers, forts, defensible settlements), from protohistory to the late historical period. His fieldwork investigations also include a study of the construction techniques of other built structures such as early temples and chortens, as well as the study of protohistorical ceremonial sites. Hubert Feiglstorfer is an architect and research associate at the Austrian Academy of Sciences, Vienna. He is the founder of the Architectural Heritage and Development Fund (AHDF) and teaches architectural history at the Vienna University of Technology. His main research interests are in the fields of sacred architecture as well as building traditions with a particular focus on the use of different kinds of materials and constructions. John Harrison is a historic buildings architect who has been working for some 20 years in the Himalaya and Tibet on research, documentation and building

x

list of contributors

conservation. In Ladakh he has been involved with the Zürich-based Achi Association, and now with Achi Association India, restoring historic Buddhist temples and training young monks in heritage appreciation; and with Indian NGO Ladakh Arts and Media Organisation (LAMO) restoring and converting a historic mansion in Leh into a community arts centre. Kath Howard is the wife of Neil Howard and has shared in all his fieldwork, as his invaluable assistant. In Ladakh she pursued her own research interests in stupa architecture upon which she has published earlier. Neil Howard is an English private scholar who has been studying fortification architecture and associated history of the old Himalayan kingdoms since 1982. His extensive fieldwork has been carried out in Nepal, Himachal Pradesh and Greater Ladakh, India. His results have been published in several academic papers; several more are to come. Kacho Mumtaz Ali Khan is based in Yokma Kharbu, near Chigtan, and is an engineer by profession. His father was the late Kacho Sikandar Khan, the well-known Ladakhi writer and historian, and he is a descendant of the former ruling family of Chigtan. Alongside his professional activities, he is currently working with the Indian National Trust for Art and Cultural Heritage (INTACH) to prevent further deterioration of the Chigtan castle ruins. Gerald Kozicz graduated from the Technical University of Graz, where he received his Ph.D in architecture and technical sciences. He has led two research projects of the Austrian Science Fond FWF. In the course of his research he has been addressing topics related to Buddhist architecture in the Western Himalaya region. His work focuses on stupa architecture and he is currently preparing a database on the subject which can be found at http://stupa.arch-research.at. Filippo Lunardo received his doctorate degree in Indological and Tibetological Studies from the University of Turin. He has worked for the Vatican Museums, and a number of organizations, as well as teaching the History of Tibetan Art at La Sapienza University in Rome. His major research interests currently focus on the image of the tshogs zhing, the merit field, in the dGe lugs pa tradition of

list of contributors

xi

Tibetan Buddhism. He has conducted fieldwork among Tibetan communities in Ladakh and elsewhere in India. Phuntsog Dorjay is an independent scholar specialising in Himalaya Buddhist art. He was born and brought up in Ladakh. After doing his Master’s degree in Ancient Indian History at the University of Jammu (India), he completed a Ph.D thesis entitled “The Development of Buddhist Art in Ladakh from 8–12th Century CE”. He now lives in Hannover, Germany. Heinrich Poell is an Austrian national, was educated as an engineer and works as a consultant for international development agencies. He has a long-time interest in the people and the culture of the Himalayas and the Indian subcontinent and has travelled extensively in the region; he has visited Ladakh more than a dozen times since his first trip in 1978. He has published several papers on Ladakhi wood art and on the iconography of the Buddha life in the early mediaeval period. Tashi Ldawa Thsangspa was born in Leh and is now an Assistant Professor in Zoology at the Eliezer Joldan Memorial College there. For the past 15 years the main focus of his research has been an extended project to document the petroglyphs of Ladakh, and to raise local people’s awareness of their importance. His earlier publications (2007) include a short study of the rock art of Nubra. Martin Vernier holds a Swiss degree in fine arts. He has been conducting archaeological surveys in Ladakh almost every year since 1996, focusing on the early history of the region. His main research interests are petroglyphs and early Buddhist stone carvings. His scholarly activities include the drawing of illustrations pertaining to Ladakhi archaeology, and various collaborations with NGOs engaged in the conservation and protection of Ladakh’s heritage.

List of Illustrations Figure

Caption

1.1

Rock with petroglyph images of people and animals beside the river Indus near Domkhar  15 Masks from Sasoma Nubra  19 Masks found in several regions  20 The Tangtse giant  21 a, b. Giant figures from Tangtse; c. Giant from Bema  22 a–c. Giant figures from Chilling valley; d–n. From Lower Ladakh; o. From Tangtse (Changthang); p. From Karu; q. From Kargil  23 Figures showing social scenes  24 A group of people with raised hands and fingers, together with some animals. Half of the rock is buried under the road debris on the Bema-Domkhar road  25 a, b. Images from Chilling  26 Animal figures from the Nubra valley  26 Animal figures from Domkhar Sanctuary  27 Deer with Chinese inscription at Domkhar  28 The Tangtse chase  29 Tracing of the Tangtse chase  29 A feline chasing deer from Renmudong, Tibet, attributed to the Iron Age  29 The deer in Renmudong  30 A selection of signs and figures from the Tangtse boulders  31 Nestorian cross with Sogdian inscription in Tangtse  32 Signs in Bema (Lower Ladakh); b-d. Kere-Kharu  33 Sign of unknown significance in Kere Kharu. Animal figures to bottom left  33 Map showing the possible early trade routes and important Buddhist rock art sites  37 Avalokiteśvara. Brass, silver and copper, height 100 cm, Tibet, C. 11th century CE  39 Rock carvings at Byama Khumbu (Suru valley) depicting Avalokiteśvara as main figure flanked by gandharvas and lay devotees below  41 Detail of lay devotees at Byama Khumbu  41 Maitreya figure at Kartse (Suru valley)  43 Detail of Kartse figure  44 Maitreya figure at Tumel village (Kargil)  45

1.2 1.3 1.4 1.5 1.6 1.7 1.8

1.9 1.10 1.11 1.12 1.13 1.14 1.15 1.16 1.17 1.18 1.19 1.20 2.1 2.2 2.3 2.4 2.5 2.6 2.7

list of illustrations 2.8

xiii

Famous Maitreya figure at Mulbek village (on Srinagar-Leh highway)  46 2.9 Detail Maitreya sculpture, Mulbek  47 2.10 Lay devotees at Mulbek  48 2.11 Avalokiteśvara and Hayagrīva near bKra shis sgo mang chorten at Changspa (Leh)  50 2.12 Detail of Hayagrīva o, a primary emanation of Avalokiteśvara, Changspa  51 2.13 Padmapāṇi (left) and Maitreya (right) on a stupa at Chute Rantak (Leh Old Town)  52 2.14 Five recently discovered rock figures in Sangkar House (Leh Old Town)  52 2.15 Badly damaged rock sculpture of Maitreya near Sangkar House (Leh Old Town)  54 2.16 Tibetan inscription on the left hand side of Maitreya near Sangkar House (Leh Old Town)  55 2.17 Maitreya relief at Skara village, Leh area  56 2.18 Standing Maitreya at Tyare Rong, recently painted by the villagers of Diskit-tsal (Leh area)  57 2.19 Five tathagatas at Shey  58 2.20 Stele of Maitreya in the courtyard of Lha rdo rje chen mo chapel, Shey  59 2.21 Avalokiteśvara, locally known as sPyan ras gzigs srang ’byung at Sakti village  61 2.22 Padmapāṇi Avalokiteśvara at Digar village (Nubra region)  62 2.23 Mañjuśrī (left) and Maitreya (right) at Tirith village (Nubra Valley)  63 2.24 A slim stele carved with Maitreya at Sumur village (Nubra Valley)  64 2.25 Recently discovered rock carvings in an old stupa at Skyu village  66 3.1 Rumtse castle from the west: the light-coloured rock outcrop summit is against the sky with settlement ruins and rubble on the slopes below and toward the camera  69 3.2 Rumtse castle: ruined masonry and rubble on and below the summit outcrop  70 3.3 Gya fortress from the west: settlement ruins, temples and some of the many stupas  71 3.4 Gya: axonometric sketch of the whole fortress, view from approximately south-west  72 3.5 Gya: axonometric sketch of the core or central area  73 3.6 Rumtse stupa field: diagrammatic sketch plan drawn from photographs  75 3.7 Rumtse stupa field: stupa wall  76

xiv

list of illustrations

3.8

Rumtse stupa field: lha bab mchod rten; this is the stupa photographed by Francke  77 Rumtse stupa field: ‘lotsava’ type stupa; two more and stupa wall behind  77 Rumtse stupa field: ‘lotsava’ type with round stupa; partial collapse shows cobble construction of the base  78 Rumtse stupa field: stupas of double rnam rgyal type  78 Gya stupa field: sketch plan, not to scale  79 Gya stupa field: lha bab mchod rten; cobble base  80 Gya stupa field: brick-built stupa  81 Map of western Tibet illustrating the inheritance of dPal gyi mgon  85 Types A, B and C chortens  102 Ubshi, plan and elevation  103 Ubshi, plan of stone roof  104 Ubshi, right wall  105 Ubshi, plan of stone-roof  106 Nang, first chorten: plan and elevation  107 Nang, first chorten: plan of stone-roof  108 Nang, second chorten: plan and elevation  110 Nang, second chorten: plan of stone-roof  111 Stongde Gandal chorten: plan and elevation  112 Stongde Gandal chorten: left wall  113 Stongde Gandal chorten: detail of a miniature Buddha  114 Thikse: plan and elevation  115 Thikse: back wall of the ground floor  116 Thikse: detail of miniature Buddhas  117 Stok Kadam chorten: plan and elevation  118 Stok Kadam chorten: plan of stone-roof  119 Stok Kadam chorten: western-most wall  120 Stok Kadam chorten: detail of a miniature Buddha  121 Nyoma Khawaling chorten: plan and elevation  122 Nyoma Khawaling chorten: east wall  124 Nyoma Khawaling chorten: south wall  124 Nyoma Khawaling chorten: west wall  125 Nyoma Khawaling chorten: north wall  125 Nyoma Khawaling chorten, details: Māyā’s dream (upper left), Shākyamuni’s first step (upper right), the lokapāla Virūḍhaka (lower left), the siddha Nāropa (lower right)  127 Nyoma Khawaling chorten: ceiling  128 Nyoma smaller gateway chorten: plan and elevation  129

3.9 3.10 3.11 3.12 3.13 3.14 3.15 4.1 4.2 4.3 4.4 4.5 4.6 4.7 4.8 4.9 4.10 4.11 4.12 4.13 4.14 4.15 4.16 4.17 4.18 4.19 4.20 4.21 4.22 4.23 4.24 4.25

4.26 4.27

list of illustrations 4.28 4.29 4.30 4.31 4.32 4.33 4.34 4.35 4.36 5.1

xv

Nyoma smaller gateway chorten: detail of northern wall  130 Nyoma smaller gateway chorten: southern wall  131 Shera: plan and elevation  132 Shera: north-western wall  134 Shera: detail  134 Zangla: plan and elevation  135 Zangla: chorten after action executed by Stupa Onlus  136 Zangla: one of the walls  137 Zangla: detail of the ceilings  138 General view of the Indus Valley from the Nyarma Ensa hermitage  143 5.2 Buddha from the ‘Chorten with the Secret Chamber’ (CSC)  144 5.3 Buddha from the Vajradhātu Temple of Sumda Chung  144 5.4 The CSC from the east  145 5.5 Stupa from the stupa field of Shey  145 5.6 The upper section of the frame of the central Buddha on the main wall of the CSC  146 5.7 Brackets inside the stupa in the Nyarma compound  146 5.8 Cross-sections of the CSC and the decorated stupa in the main Nyarma compound  147 5.9 Floor plans of the CSC  147 5.10 Sealed chambers of the Nyarma Stupa Temple  149 5.11 Main wall of the ground floor with remains of the original Buddha in the centre  149 5.12 Stone masonry of the west (left) wall  149 5.13 View from the ground floor of the CSC towards the corbelled roof  150 5.14 Mandala ceiling of the Alchi stupas  150 5.15 Axonometric comparison of the CSC and the Alchi Entrance Stupa  151 5.16 Axonometric comparison of the different spatial forms  152 5.17 Upper corner area of the main wall and the adjoining right wall with the decorative elements  154 5.18 Decorative banners and the row of chortens depicted above the entrance  154 5.19 Comparison of stupa types from mural paintings  155 5.20 Eyes on a dome inside the CSC  155 5.21 A complete face is depicted on a stupa above Maitreya from Saspol Cave No. 2  155 5.22 Complete interior elevations of the main wall and the opposite entrance wall  156

xvi 5.23 5.24 5.25 6.1

6.2 6.3 6.4 6.5 6.6 7.1 7.2 7.3 7.4 7.5 7.6 7.7 7.8 7.9 7.10 7.11 7.12 7.13 8.1a/b 8.2 8.3 8.4 8.5 8.6 8.7 8.8 8.9

list of illustrations Bodhisattva of the upper storey  156 Inner stupa from the shrine stupa approximately 50 m from the Alchi compound  158 A wooden votive stupa from Saspotse  158 Murals on the top storey of the Sumtsek temple, with members of a Kagyupa lineage in the left-hand panel, and rows of monks on the right-hand panel  159 Left-hand panel showing the Kagyupa lineage with accompanying dbu can inscriptions  160 Panel in the Sumtsek murals with obliterated inscription  161 16th century inscription in the Sumtsek  161 Rows of monks with obliterated inscriptions  162 Inscription including the name of Tshul khrims ’od  164 Mural painting: The Royal Drinking Scene, c. mid-11th century, the Dukhang, Alchi  168 Mural painting, 11th century, Mangyu, Ladakh  169 Gold medallion, 11th–12th century, Iran  172 Stone statue, Turkic balbal with drinking cup from Lake Issyk-kul  173 Detail of The Royal Drinking Scene  173 Detail of the lower scene of The Royal Drinking Scene  175 Mural, late-10th century, Entry Hall, Tabo, Spiti, Himachal Pradesh, India  177 Mural depicting the Pilgrimage of Sudhana, mid-11th century, Assembly Hall, Tabo  177 Detail of The Royal Drinking Scene  179 Seljuq bowl depicting seated and standing figures in a garden dated March–April 1187 CE (Muharram 583)  180 Mural, The Ruler, the Cup and the Hunt, the Dukhang, Alchi  184 Detail of The Ruler, the Cup and the Hunt  184 Detail of The Ruler, the Cup and the Hunt  185 Lachuse temple, view from the south–east and floor plan  192 Lachuse temple, inside view  193 Portal of the Lachuse temple  194 Scene 1, Buddha life frieze, Lachuse doorframe  196 Scene 2, Buddha life frieze, Lachuse doorframe  196 Scene 3, Buddha life frieze, Lachuse doorframe  197 Scene 4, Buddha life frieze, Lachuse doorframe  198 Scene 5, Buddha life frieze, Lachuse doorframe  198 Three scenes from the Buddha life on the Dukhang door in Alchi  202

list of illustrations 8.10 8.11 8.12 8.13 8.14 8.15 8.16 8.17 8.18 8.19 8.20 8.21 8.22 8.23 8.24 9.1 9.2 9.3 9.4 9.5 9.6 9.7 9.8 9.9 9.10 9.11 9.12 9.13 9.14 9.15 9.16 9.17

xvii

Votive shrine; painted wood; Kashmir, 8th/9th century, British Museum  203 Left pillar of the Lachuse portal  205 Right pillar of the Lachuse portal  205 Reconstruction of the pillar pair of the Lachuse portal  207 Reconstruction of the Lachuse portal  209 Portal of the Chigtan temple  210 Remains of the temple façade in Lachuse  212 Central sculptural panel of the Lachuse façade  213 Tripartite façade of the Alchi period  214 Sculptured face of the capital in the Lachuse façade  215 Three-column arrangements in (a) Alchi and (b) Wanla  217 Column capital in the Lachuse temple  218 Central niches of the Lachuse column capitals  218 Pilaster capital over the entrance to the monks’ house  220 Pilaster capital from the vestibule of the Lotsawa temple, Alchi  220 The Monastery of dPe thub  227 The master Shes rab bzang po and his disciples, portrayed in cave No.2 at Sa spo la  228 The master Nam kha’ ba and his disciples, portrayed in cave No.2 at Sa spo la  230 Deity belonging to the cycle of ‘Direction Deities’, painted on the east wall of the mGon khang of dPe thub  232 Pictorial cycle painted on the north wall of the mGon khang of Khrig se  233 The north wall of the mGon khang of dPe thub  236 Yamī painted on the north wall of the mGon khang of dPe thub  238 An “enemy of the Doctrine” painted on the north wall of the mGon khang of dPe thub  240 Acolytes of Yama. North wall of the mGon khang of dPe thub  240 Acolytes of Yama. North wall of the mGon khang of dPe thub  241 Acolytes of Yama. North wall of the mGon khang of dPe thub  241 Vajradhāra. East wall of the mGon khang of dPe thub  242 Deity. East wall of the mGon khang of dPe thub  243 Nāro Ḍākinī. East wall of the mGon khang of dPe thub  244 dPal ldan lha mo. East wall of the mGon khang of dPe thub  245 Brahmā. East wall of the mGon khang of dPe thub  247 The “enemy of the Doctrine”. Detail of painting on the east wall of the mGon khang of dPe thub  248

xviii 9.18 9.19 9.20 9.21

list of illustrations

Śiva and Pārvatī. East wall of the mGon khang of dPe thub  249 Sita Varuṇa. East wall of the mGon khang of dPe thub  250 Vāyu. East wall of the mGon khang of dPe thub  252 The “Eight Masters on Horseback”. South wall of the mGon khang of dPe thub  253 10.1 Chigtan castle in 1909  255 10.2 Chigtan castle in 2009  256 10.3 Sketch plan of Chigtan fortifications  261 10.4 Ruins of the upper tower  262 10.5 Partly enclosed tower. The original stone tower is just visible inside the shuttered mud walls  263 10.6 The stone tower behind the walls of shuttered mud  263 10.7 Rear of the main castle  264 10.8 Timber lacing as seen in 1909  265 10.9 Drawings of the first capital in the mosque  269 10.10 Simplified line drawing showing the patterns in the first capital  270 10.11 Drawings showing the second capital in the mosque  270 10.12 Simplified line drawing showing the patterns in the second capital  271 10.13 Carving from Kammand temple in Himachal Pradesh, dated approximately to the late 14th/early 15th centuries  272 10.14 Carving from Kammand temple in Himachal Pradesh  272 11.1 Statue of Bakula Rangdröl Nyima Rinpoche placed at the centre of the Chenrezik Lhakhang  274 11.2 Overview of the mural painting  280 11.3 Küntuzangpo/Samantabhadra with Küntuzangmo/Samatabhadri  281 11.4 Main Peaceful Deities: the Tathagatas Vairochana, Akshobhya-Vajrasattva, Ratnasambhava, Amitabha, Amoghasiddhi  281 11.5 and 11.6 The inscription below the Vairochana Buddha  282 11.7 On the sixth day of the chos nyid bar do the five peaceful Buddhas appear together, each with his consort  283 11.8 The inscription of the Vidyadhara deities that come to meet the deceased  284 11.9 Main wrathful deities: Mahottara Heruka, Buddha Heruka, Vajra Heruka, Ratna Heruka, Padma Heruka and Karma Heruka, each with his consort  285 11.10 The assembly of the 58 wrathful deities with their entourage  285 11.11 Detail of two Keurima deities: Tseuri and Tramo Marmo  286 11.12 Two Ishvari deities, yoginis of the southern direction: Gawa Marser and Zhiwa Marmo  286 11.13 First image of srid pa bar do  287

list of illustrations 11.14 11.15

xix

Inscriptions on the slate of the monkey-headed assistant  288 Signs for places where the deceased is to take rebirth: lights of the six realms, and feeling of going up, sliding down or walking flat  288 11.16 Appearances arise such as whirlwinds, blizzards, hailstorms, darkness  289 11.17 Image of the four continents  290 11.18 Image of five unfavourable realms where the deceased should not take rebirth: gods’ realm, demi-gods’ realm, animals’ realm, hungry ghosts’ realm, hell realms  290 11.19 The inscription for the last image  291 12.1 Copy of the Kesar wall painting  301 12.2 The first of the Changspa Kesar paintings  304 12.3 The two warriors  306 12.4 Close-up of one of the mounted warriors  307 12.5 The palace  308 12.6 Painting from the Völkerkundemuseum Herrnhut  309 12.7 Detail of painting in Völkerkundemuseum Herrnhut  310 13.1 The tshogs zhing’s oldest iconography, late 18th century/early 19th century  317 13.2 19th century tshogs zhing  318 13.3 19th century tshogs zhing  319 13.4 Central portion of the Spituk/dPe thub New ’Du khang tshogs zhing  321 13.5 Detail of the main figure of the tshogs zhing, Tsong kha pa  323 13.6 The portion of the tshogs zhing to the proper left of the main, central figure, Tsong kha pa  325 13.7 The portion of the tshogs zhing to the proper right of the main, central figure, Tsong kha pa  326 14.1 Weaving Mahakala. A weaver in Benaras shows the image of Mahakala, the God of Protection, that he is weaving on his pit loom  329 14.2 The apron of Mahakala is wrapped around a statue of the same god at Hemis monastery, Ladakh  330 14.3 Appliqué and patchwork thang ka of Gyalse Rinpoche, Hemis monastery  332 14.4 Monks, resplendent in their brocade robes, perform at the annual monastic festival, Hemis monastery  333 14.5 A participant at a horse race is dressed in his brocade robe and hat, Rupshu, Changthang  333 14.6 A woman’s cape made from Chinese silk-brocade from the Shangara family  334

xx 14.7 14.8 14.9 14.10

15.1

15.2 15.3 15.4 15.5 15.6 15.7 16.1 16.2 16.3 16.4 16.5 16.6 16.7 16.8 16.9 16.10 16.11 16.12

list of illustrations Haseen Ahmed Kasim hold up a piece of silk-brocade in his office in Benaras  341 Aminuddin Ansari, a trader from Benaras, shows samples of his silkbrocade to Lobsang Dorje  343 Silk processing activities and preparation of the warp in Pili Kothi, Benaras  345 A piece of brocade is given prominence on a wall in Haseen Ahmed Kasim’s office, flanked by embroidered hangings of ‘Allah’ and pictures of the Dalai Lama  347 View of Leh Old Town showing the royal palace as well as the monasteries, temples and private houses of the king`s ministers just below  350 A thick layer of ‘markalag’ being laid on the roof by mason Sonam Dorjey  355 Composition of ‘markalag’ samples as analysed by the University of Stuttgart  356 A section through a Ladakhi roof shows the main beams, the secondary beams and the ceiling joists  357 The interior of the restored Masjid Sharif  358 Sankar Labrang, now named Lala’s café, in Leh’s Manikhang area  360 Anca Nicolaescu cleaning the mural paintings inside the Red Maitreya Temple  362 Likir. Clay pit with rdza sa, the pottery soil  370 Likir. Clay pit with thab sa, the soil used for making traditional Tibetan stoves  371 Phugtal. Clay pit with rdza sa, locally known as ‘arga’  372 Looking over the red earth (’phred sa) mountain of Basgo  373 Lamayuru. Clay pit with a striped ‘markalak’  374 Wanla, Sumtsek. Interior view of a wall section beside the portal, showing a detachment of the plaster  375 Nyarma, Tsuklakhang, Nangkor. Exterior plaster mixed with a high amount of straw  377 Likir. The white spot in the centre of the picture is dkar rtsi, a local species of lime  381 Wanla, chorten. According to Sandeep Kumar, locally available dmar rtsi was for the red colour  382 Basgo. Clay pit with a specific kind of white shining ‘markalak’  383 Basgo. Traditional Tibetan stove (thab), located in the house of Mr. Angchuk’s brother  384 Tunlung. Clay pit with orange-brown rdza sa  385

list of illustrations 17.1

xxi

17.5 17.6

Prayer flags and telephone mast in Leh: communicating with other worlds  388 Mulbek. Colossal rock-carved statue of Maitreya, between the 9th and 11th century  391 Skurbuchan Khar  392 Leh Old Town, section from Tashi Khangsar House down to Abdul Wahid Choskor House behind the Jami Masjid  395 Lonpo House, Leh  396 LAMO centre and the restored Munshi House, Leh  397

map

Caption

7.1

Alchi, Mangyu, Tabo, Tholing and Dungkar (map not to scale)  167

17.2 17.3 17.4

Introduction Erberto Lo Bue and John Bray Ladakh’s particular interest to scholars lies in its status as a region with close religious and cultural affinities with Tibet that at the same time looks west to Kashmir, south to the Indian plains and north to Turkestan/Xinjiang. At times the Tibetan influence has seemed to predominate to the extent that a collection of wedding songs recorded in the early 20th century in the eastern Ladakhi village of Hanle (Wam le) refers to “us Tibetans” (Filibeck 2009:197– 199). However, the transmission of religious and artistic ideas has not been solely in one direction. Between the 11th and 13th centuries CE Ladakh occupied a particular place in the history of geo-cultural Tibet, representing a bridge between India and the Land of Snows, as exemplified by the painting, sculpture and architecture of Alchi (A lci) in the Indus valley. Today Alchi serves as the best surviving testimony to the highly evolved Buddhist culture of mediaeval Kashmir. In more recent centuries the flow of Buddhist ideas and artistic styles has been more from Tibet to Ladakh rather than in the reverse direction. The tradition of sending Ladakhi novices to Tibet for their higher studies started as early as the 13th century, and Tibet’s spiritual pre-eminence is reflected in later Buddhist monastic art in the region. However, Ladakh at the same time remained subject to multiple cultural influences from the west and the south as reflected, for example, in its musical traditions.1 Nearly half its population is Muslim and its architectural inheritance includes an extensive network of mosques as well as monasteries. An earlier volume in Brill’s Tibetan Studies series, Ladakhi Histories: Local and Regional Perspectives (Bray 2005 and 2011), discusses the impact of these competing influences on the region’s political and cultural history. A second volume, Modern Ladakh, reviews the contemporary processes of social change. This book, which is largely drawn from contributions to the 13th, 14th and 15th International Association for Ladakh Studies (IALS) conferences in 2007, 2009 and 2011, celebrates the region’s artistic and architectural inheritance.

1 On this point, see Trewin (1990:273–276), who discusses both Tibetan and Kashmiri/Islamic influences on Ladakhi folk music.

© koninklijke brill nv, leiden, ���4 | doi ��.��63/9789004271807_�02

2

Lo Bue and Bray

Central Themes

The papers in the collection range widely over time, from prehistory to the present day, and the subjects for analysis extend from rock art to castle architecture, monastic murals and silk brocades. Amidst this diversity, it is possible to identify three underlying themes, perhaps best expressed as questions.



The first is to do with the region’s connectedness. Ladakh lies across the great historical communications routes linking the Indian subcontinent with the regions north of the Karakoram as well as Tibet. So how are these manifold external influences on the region expressed in its material culture? The second question, which follows immediately from the first, concerns local particularity. Anthropologists working on Ladakh have emphasized the distinctive local roots of many aspects of village-level ritual, and contemporary Ladakhis are keen to assert their own regional identity, distinct from Tibet and from other parts of South Asia. So how far should we see Ladakh as a subset of a civilization that has come from somewhere else— whether from Central, South or even West Asia—and how far is it the result of a synthesis? The third question concerns change, evolution and decay. In the Tibetan monastic world it has long been customary to repaint ancient murals in new religious and artistic styles. In Ladakh as elsewhere, ancient buildings are often composite constructions of different layers and styles. As will be seen, these complexities lend themselves to extensive debate on the dating of particular buildings or paintings. Meanwhile, the question whether to replace or discard old structures has taken on a new urgency in contemporary Ladakh: how far is it possible to preserve the region’s artistic and architectural inheritance in an era of rapid social, economic and cultural change?





None of these questions lend themselves to simple or definitive answers. In contributing to a greater understanding and appreciation of the region’s artistic and architectural past, we hope that this book will provide a clear basis for fresh scholarly enquiry, and perhaps for practical decisions on the conservation of its cultural inheritance.

Scholarly Foundations

The contemporary study of Ladakh is itself built on older scholarly foundations and all the papers in this collection build on them. As discussed in Ladakhi

Introduction

3

Histories, the beginnings of a critical historical analysis date back to the 19th century. The single most important source is the La dvags rgyal rabs, the royal chronicle of Ladakh, which dates back to the 17th century. At the same time literary sources need to be supplemented by evidence from inscriptions on rocks and wall paintings as well as the buildings themselves. Many of the early researchers on these topics in the late 19th and early 20th centuries were either missionaries or colonial officials. The two-volume study Antiquities of Indian Tibet by the Moravian missionary August Hermann Francke marks an important early 20th century landmark. In 1909 Francke embarked on an extended research expedition under the auspices of the Archaeological Survey of India (ASI), travelling from Simla to Kinnaur and then via Spiti to Ladakh. The first volume of the Antiquities, which was published in 1914, includes a description and a preliminary analysis of the paintings in Alchi in Ladakh. The second volume (1926) contains a critical edition of the La dvags rgyal rabs. On his research expedition, Francke was accompanied by a skilled photographer, Babu Pindi Lal. His photographs themselves constitute an important part of the historical record, as discussed in the papers by Howard and Khan, Bray, Devers and Vernier in this volume. Tucci’s four-volume Indo-Tibetica, published between 1932 and 1941, marks the next major scholarly landmark. For Ladakh, the first two volumes are of particular importance in that they focus on the architecture of the chorten (mchod rten) and the biography of the Buddhist scholar and translator Rinchen Zangpo (Rin chen bzang po, 958–1055). In the same period, Tucci’s student Luciano Petech followed up Francke’s earlier work on the La dvags rgyal rabs with his 1939 Ph.D. thesis, published under the title of A Study on the Chronicles of Ladakh (Indian Tibet). The outbreak of the Second World War (which Petech spent in an internment camp near Dehra Dun because his Italian nationality made him an ‘enemy alien’) marked the beginning of an extended hiatus in scholarly research on Ladakh. In 1947 and 1948 Ladakh was caught up in the India/Pakistan conflict over Kashmir. From the mid-1950s until 1974 Ladakh was closed to foreigners because of its status as a sensitive border area. Poor road communications and limited flight connections meant that it was scarcely accessible even to scholars from other parts of India. However, in 1974 Ladakh was opened up to tourism, and the period since then has seen rapid social and economic change as well as a proliferation of new research by both local and international scholars. Significant contributions by Ladakhi scholars include three historical studies: Joseph Gergan’s Bla dvags rgyal rabs ’chi med gter (1976), Tashi Rabgias’s Mar yul la dvags kyi sngon rabs kun gsal me long zhes bya ba bzhugs so sgrig pa po (1984), and Kacho Sikandar Khan’s Qadim Laddakh tarikh va tamaddun (1987). The first major new

4

Lo Bue and Bray

contributions by international scholars were Petech’s History of Ladakh (1977), which revised and greatly expanded his earlier work, and the two volumes of The Cultural Heritage of Ladakh (1977, 1980) by David Snellgrove, Tadeusz Skorupski and Philip Denwood. The first volume of the latter work, Central Ladakh, focuses on the murals in the temples of Alchi, which are widely agreed to be among the most outstanding of all Ladakh’s historical monuments. This initial study paved the way for further research by scholars such as Roger Goepper, notably including his lavishly illustrated Alchi. Ladakh’s Hidden Buddhist Sanctuary (1996). Meanwhile, Erberto Lo Bue (1983, 2005, 2011) has documented contemporary Buddhist artists in Ladakh, while Neil Howard (1989) pioneered the study of Ladakh’s ancient fortresses. Major contributions in the art-historical field by Christian Luczanits include Sculpture in Clay (2004), which examines sculptures dating from the late 10th to the early 13th centuries in Kinnaur and Spiti as well as Ladakh. Monisha Ahmed and Clare Harris presented a valuable collection of papers on Ladakh’s material culture in Ladakh: Culture at the Crossroads (2005). The research of the last 40 years has therefore greatly extended the knowledge that was inherited from Francke, Tucci and Petech. However, there are still several lacunae to be filled before it is possible to write a comprehensive new history of the region’s art and architecture. This collection highlights a selection of the most important new research findings.

Historical Perspectives

The papers in the collection are presented in rough chronological order, starting with Ladakh’s prehistory. The themes of international influence and local particularity combined with evolution and decay recur at every stage. Early Petroglyphs and Rock Sculptures Petroglyphs in Ladakh have been the object of research since Francke’s time, and over 20 years ago Giacomello Orofino published an article illustrated by an important set of photographs of rock art taken during Tucci’s visit to the region in the 1930s. However, it is only in recent decades that scholars have come to appreciate the full extent and the sheer diversity of the thousands of petroglyphs scattered all over Ladakh and neighbouring regions. Tashi Ldawa Tshangspa has made an important contribution in mapping out and recording these petroglyphs, and in this volume he presents a small selection of his findings, categorizing them according to their subject

Introduction

5

matter. Petroglyphs reflect the local fauna and economy, for example in the depiction of animals that were hunted by the region’s earliest human inhabitants. However, they also suggest that Ladakh was far from isolated even in the prehistoric era. Images of human ‘masks’ from the third and second millennia BCE have analogies in what is now northern Pakistan and as far afield as southern Siberia. Later images of animals standing on ‘tip-toe’ point to a historical connection with the Scythian cultures of Central Asia in the first millennium BCE. Ladakh was incorporated into the Yarlung dynasty’s Tibetan empire in the late seventh or early eighth century CE, and research by Philip Denwood has shown that the first Tibetan settlement in the region may have taken place during that period. However, at that time as in the following centuries, the most significant religious and artistic influences came from Kashmir rather than Tibet. The gigantic Maitreya rock relief at Mulbek (Mul bhe), on the LehSrinagar road, is a prominent example. Phuntsog Dorjay’s article examines this and other less well-known Buddhist rock reliefs found mostly along ancient trade routes in the Kargil, Leh and Nubra regions. He attributes most of them to the period between the ninth century and the early 11th century, and shows that their style closely resembles that of bronze images produced in Kashmir during the same period. Art and Architecture during the Second Diffusion of Buddhism The history of Ladakh’s art and architecture entered a new phase after the fall of the Tibetan empire in the mid-ninth century and the establishment of a Tibetan dynasty in western Tibet, known as Ngari Korsum (mNga’ ris skor gsum), in the first half of the tenth. After the death of King Skyilde Nyimagon (sKyid lde nyi ma mgon) in c. 950, his kingdom was divided among his three sons. One of them, Palgyigon (dPal gyi mgon), inherited Ladakh and may be regarded as the founder of the Ladakhi kingdom. In practice, however, the region remained politically fragmented, with a number of larger and smaller principalities exercising a high degree of local autonomy. The rulers of the Ngari Korsum kingdoms came to play an important role as the patrons of the second diffusion of Buddhism (bstan pa phyi dar) in geocultural Tibet. In the second half of the tenth century King Yeshé Ö (Ye shes ’od), sent 21 young noblemen to Kashmir to study the Buddhist scriptures. One of the two survivors was Rin chen bzang po, who was to become one the greatest protagonists of the second diffusion of Buddhism in Tibet, famous both as a translator of Buddhist scriptures from Sanskrit into Tibetan and as the founder of temples and monasteries in Western Tibet across Purang, Kinnaur, Spiti and Ladakh.

6

Lo Bue and Bray

This political and religious background sets the context for Neil and Kath Howard’s contribution on the archaeological remains in and around the little studied district of Gya (rGya), in eastern Ladakh. Following up on Neil Howard’s earlier study of Ladakhi fortresses, they first examine the castles at Rumtse (Rum rtse) and Gya, both located along the main trade route from India and on a branch of a route leading from Kashmir to Tibet, before turning to the nearby field of chortens, which they attribute to the time of the second diffusion of Buddhism in Tibet. The third part of their paper discusses Gya’s historical significance as one of the more powerful local principalities within the wider Ladakh region. The authors attempt to set the surviving ruins in a proper historical context suggesting that Gya was an important area about one thousand years ago and that a re-examination of references to it in historical sources might contribute to define major issues for the purpose of further research on the topic. An appendix by Philip Denwood discusses a reference to Gya in the chronicles of Guge Purang. Supporting his case with a careful analysis of the written sources, he argues that this refers to eastern Tibet and not to the settlement in Ladakh as suggested earlier by Roberto Vitali (1996:95). Quentin Devers, Laurianne Bruneau and Martin Vernier develop a related archaeological theme with a carefully illustrated study of ten examples of painted chortens, dividing them into three types: ‘plinth’, ‘Lotsawa’ and ‘gateway’ chortens. They note that all have the external appearance of a chorten while their cores shelter paintings that are similar to those of a shrine. The three authors argue forcefully for greater recognition of these chortens’ significance in the history of Buddhism in Ladakh, while emphasizing the urgent need to preserve them before they fall into decay. In a similar vein, Gerald Kozicz presents a detailed examination of a c­ horten in the area of Nyarma (Nyar ma), the only monastery founded in Ladakh by Rin chen bzang po according to historical sources, which today lies in ruins. In spite of its importance, the Nyarma area has not been the subject of a detailed monograph until now. Jampa Panglung (1983) devoted an article to the monastery and later Kozicz (2007a, b) himself devoted two contributions to its neighbouring area. Here he analyses the architectural structure of what he calls the “Chorten with the Secret Chamber”, together with the iconography of its interior. He compares it with roughly contemporary chortens from elsewhere in the Nyarma area and at the site of Alchi, and concludes by dating it to the late 11th or early 12th century. Alchi and Lachuse The Choskhor (Chos ’khor) monastic compound in Alchi (A lci) is among the most closely studied of all Ladakh’s historical monuments. However, there is

Introduction

7

still no consensus on the dating of the most important buildings in the compound. Denwood (in Snellgrove & Skorupski 1980) examined the inscriptions on the site and concluded that the Dukhang (’Du khang) was founded in the early to mid-11th century while the Sumtsek (gSum brtsegs) belonged to the last third of the century. Goepper (1996:212 and 216–217; 2003:14–24) subsequently published further important inscriptions from the top storey of the Sumtsek, and argued that its founder “must have been active around the end of the 12th to the beginning of the 13th century”. In his contribution to this volume, Philip Denwood restates his opinion and amplifies his hypothesis, based on epigraphic and palaeographic evidence, that the Sumtsek was founded in the last third of the 11th century, arguing that some paintings and inscriptions were added at a later date. He suggests that a later generation of monks resorted to obliteration and rewriting in order to claim a Drigung (’Bri gung) inspiration for the foundation of the temple after it was taken over by the order of that monastery. The most puzzling scene painted in the assembly hall of the monastic site of Alchi portrays a ruler being offered a cup by a woman. Here Marjo Alafouzo relates this painting to its historical context, pointing out that the characters depicted in the mural were instrumental in the founding of the temple. By analysing its iconography and comparing it with art historical material from Tibet as well as from pre-Islamic and Islamic Central Asia, she argues that such scenes are attested to in Turkic literary records and suggests that it depicts a marriage between a Tibetan woman and a foreign ruler, possibly Turkic. Several chapels were built in the wake of Rinchen Zangpo’s followers’ activity in Ladakh. One of the least known is a chapel with a fine carved wood doorframe in the village of Lachuse. Heinrich Poell’s paper relates the carvings to the Indian traditions prevalent in the regions of present-day Himachal Pradesh as well as Kashmir during the 11th and 12th centuries. The 14th to the 18th Centuries With the Islamization of Kashmir from the 14th century, Tibet became Ladakh’s only external source of Buddhist learning. From the following century the Geluk (dGe lugs) religious order joined the Drigung one in renovating existing institutions, and building new temples and monasteries. Just as the Drigung school had taken over foundations such as Alchi and Lamayuru (Bla ma g.yu ru) starting from the 13th century, so the Geluk order took over foundations such as Spituk (dPe thub) two centuries later. In this volume Chiara Bellini deals with the statues and fine murals fashioned presumably in the early 15th century in the mgon khang near the monastery of Spituk. Resemblances between the paintings in Spituk, Thikse (Khrig se) and Phugtal (Phug tal) reveal the

8

Lo Bue and Bray

existence of a school of painting, very likely of Tibetan origin and associated with the Geluk order. The ruined castle in Chigtan (Cig gtan or Cig ldan) stands as testimony both to political turbulence in Ladakh and to the skill of its builders. The castle is thought to have been built on an earlier site at the turn of the 16th and 17th centuries on the orders of a local ruler named Tsering Malik (Tshe ring ma lig). During this period Chigtan was caught up in the conflicts between Ladakh and Baltistan, and the need for fortifications is readily apparent. The paper by Kacho Mumtaz Ali Khan (a descendant of Chigtan’s ruling family), John Bray, Quentin Devers and Martin Vernier examines the castle’s history and its archaeological remains, drawing on the evidence of photographs taken by Babu Pindi Lal in 1909. The castle looks in two directions culturally as well as politically. According to local tradition, Chigtan was founded by a migrant from Gilgit. One of the inner buildings of the castle complex made extensive use of timber lacing in a style reminiscent of the older mosques and castles of Baltistan. By contrast, the front building with slightly tapering walls is firmly rooted in Tibetan architectural tradition. On a similar note, Tsering Malik is understood to have been the first member of his family to adopt Islam and his name is a composite of Buddhist and Muslim elements. The small mosque at the foot of the castle incorporates a ‘lotus’ design in its doorway, as well as hybrid dragons carved on the capitals of the central pillars. Careful examination shows that the capitals must have come from an earlier building, possibly the castle itself. The dragons show features that are reminiscent of the Central Asian steppe cultures’ animal style as seen in some of the most notable petroglyphs discussed in Tashi Ldawa’s paper in the beginning of the volume. The 19th and 20th Centuries In 1834 an army led by Zorawar Singh invaded Ladakh on behalf of his master, Raja Gulab Singh of Jammu. The invasion marks a definitive turning point because it led to Ladakh’s incorporation into the dominions first of Jammu and then of the combined princely state of Jammu & Kashmir within Britain’s Indian Empire. From now on Ladakh’s political alignment was with South Asia even though it retained its cultural and religious connections with Tibet. The Dogra invasion was a profound shock for Ladakh’s Buddhist leaders, not least because it was accompanied by widespread physical destruction and looting. In Lamayuru, according to a chronicle by Bakula Rangdröl Nyima Rinpoche (Ba ku la Rang grol nyi ma), the original buildings of the monastery were all but destroyed. As abbot, Bakula Rangdröl Nyima was responsible for supervising the monastery’s reconstruction and rebuilding the Chenrezik

Introduction

9

Lhakhang (sPyan ras gzigs lha khang) in the northern section of the monastic compound was particularly dear to his heart. As Kristin Blancke explains in her paper, the temple contains a mural representing the visions one encounters in the intermediate state between death and rebirth (as described in the Bar do thos grol chen mo, known in the West as the “Tibetan Book of the Dead”). These murals appear to be without parallel elsewhere in geo-cultural Tibet. Another iconographic theme seldom represented on Tibetan murals is the saga of Kesar of Ling (Gling Ge sar).2 Here, John Bray discusses a set of Kesar paintings that formerly existed in the house of the Kalon (bKa’ blon) family at Changspa, near Leh. The first Western references to the paintings appear in a footnote to an article by the Moravian missionary Karl Marx in 1891 and Francke’s assistant Babu Pindi Lal was able to photograph them in 1909. At that time the owner had already begun to whitewash the paintings and they have since disappeared completely, to the extent that the Kalon family now has no memory of them. Pindi Lal’s photographs are therefore all the more important as a testimony to a lost art work. Despite the destructions of the past, the aesthetic quality of much of the more recent Buddhist painting and sculpture in Ladakh is often of a high quality. Filippo Lunardo’s paper analyses a particular mural in the new assembly hall of Spituk monastery, built and decorated during the second half of the 20th century. That painting is a visualization of the “assembly field (tshogs zhing) of gurus and deities according to the bla ma mchod pa liturgy, a practice aimed at the accumulation of merits necessary to achieve the final goal of the union of emptiness and great bliss, as expected in Vajrayāna praxis. It serves as evidence of the continuing vitality of Buddhism in Ladakh in recent decades. Monisha Ahmed’s paper examines the trade and use of silk brocade in Ladakh, skilfully drawing a link between the art-historical evidence and contemporary practice. She notes that the murals of the Alchi Sumtsek reproduce textiles made with various techniques: some would have been produced in Ladakh itself while others—including brocade—would have come in through trade. When the English traveller William Moorcroft and his companion George Trebeck travelled to Ladakh in 1820, they found that Chinese brocade was among the merchandise that they saw arriving on caravans from Central Asia. However, in more recent times Benaras (now Varanasi) has replaced China as a source for high-quality brocade. The article discusses the role of a 2 During his fieldwork in eastern Tibet (1997) Erberto Lo Bue recorded one instance of a cycle of murals devoted to the Kesar saga, painted on the walls of a cloister in the Sa skya monastery of Wara (Wa ra) in. Whereas the latter survived the destruction caused by the Cultural Revolution, the Ladakhi paintings were not so fortunate.

10

Lo Bue and Bray

Benaras Muslim family, the Kasims, in preserving and developing the tradition to the present day. Contemporary Conservation Challenges The last three papers in the collection confront the problems of conservation. Ladakh is rich in historical monuments from monasteries, mosques and chortens to castles and palaces. Many of them are rapidly falling into disrepair. So what can and should be done to preserve them? The Old Town of Leh is a prime example both of the challenges of conservation and of potential solutions. In their contribution André Alexander and Andreas Catanese report about the challenges of conservation as well as the discoveries they made during the work they carried out implementing the Leh Old Town Project. They pay attention not only to the original structure and urban development of historical Leh, but also to the planning of homes and temples in the town, where they discovered ancient murals under a coat of whitewash in the Maitreya Temple below the royal palace. As explained by the authors, the Old Town has considerable importance as one of the few surviving examples of historical urban architecture in the wider Himalayan and Tibetan world. However, in recent years most of the richer families who used to live there have moved to more comfortable dwellings in the outskirts of Leh. Many of the remaining inhabitants live off low incomes and their lives have been blighted by the lack of such facilities as a modern water supply. Alexander and Catanese’s paper argues that there is no realistic prospect of preserving the old houses of the town unless they can be adapted to the changes in lifestyle of their owners and occupiers. Working with the local community, the Tibet Heritage Fund (THF) has set up a programme to revive traditional crafts and building skills in order to restore the buildings while at the same time making them more habitable. Among other techniques, the craftsmen sponsored by THF have revised the use of a kind of clay known as ‘markalak’ (mar ka lag) to waterproof roofs rather than resorting to concrete or corrugated iron. Tragically, Alexander died suddenly in January 2012 at the early age of 47. As his obituarists pointed out, he was in all respects a ‘master builder’. It is to be hoped that both the buildings themselves and the THF’s wider initiatives will prove to be sustainable long into the future. Hubert Feiglstorfer’s contribution complements Alexander and Catanese’s with a detailed examination of the use of earthen building materials in traditional buildings in Ladakh. He includes a scientific analysis of the different kinds of soil as well as additives such as the pulp of Ladakhi apricots that were used to improve their cohesion. Pointing out that many buildings made with these materials and techniques have survived for hundreds of years, and that

Introduction

11

now few local people know about types of soil traditionally used as building material, Feiglstorfer emphasizes the importance of recording ancient surviving knowledge in order to gain a comprehensive understanding of traditional buildings. Furthermore, the author argues that it is essential to understand the strengths and weaknesses of traditional building materials and techniques when developing new approaches to architectural conservation. The concluding paper by John Harrison looks at heritage conservation questions from a wider perspective, raising basic issues such as what the Ladakhi “heritage” is supposed to be and how it ought to be preserved in a changing society. Like several others in this volume, this paper was originally presented in 2007 at the IALS conference in Rome, where its author observed that even the ‘Eternal City’ had passed through a constant process of adaptation and rebuilding in which ancient monuments were often quarried to provide the raw materials for new constructions. The contemporary significance of ancient buildings has often been contested, both in Europe as well as in other parts of India. Reviewing a series of recent initiatives in Ladakh, Harrison notes that much of the restoration work has been undertaken by outsiders from elsewhere in India and from the West. He suggests that this may be acceptable in the short term, but concludes that Ladakhis must be involved to a much greater extent if the region’s architectural heritage is to “survive and become an essential part of the future”.

Looking Ahead

Taken together, the papers in this book serve as a guide to the insights of recent scholarship while helping to define an agenda for the future. The challenges are formidable. The contributors to the book come from a range of academic backgrounds: history, art history, architecture and archaeology. Looking ahead, it is essential to draw on the insights from all these disciplines while ensuring that they are firmly rooted in historical scholarship, including an understanding of the Tibetan-language written sources. This task demands a willingness on the part of both international and local scholars to cross interdisciplinary boundaries, and to make extensive personal investments in acquiring the requisite linguistic and other skills. The still greater challenge concerns the future of all aspects of Ladakhi culture, including its built heritage. A recent publication by the Namgyal Institute for Research on Ladakhi Art and Culture (NIRLAC 2008), a Ladakh-based NGO, highlights both the scale of the problem and the potential opportunities.

12

Lo Bue and Bray

Starting in 2003, NIRLAC set out to prepare an inventory of Ladakh’s cultural resources, a term which it defines broadly to include both religious and secular constructions as well as features of the landscape that are considered to be sacred. The inventory lists new and restored buildings as well as older ones: it covers 400 villages in Leh and Kargil districts and records 4,250 sites. Even this extensive list is far from comprehensive. For example, it covers no more than a small selection of the hundreds of petroglyphs in the region. All too many of the chortens in NIRLAC’s listing are listed as being in poor condition or even “danger of disappearance” and in recent years many of the most important petroglyphs—for example a historic inscription near Khalatse—have been destroyed in road-building programmes. In rural Ladakh as in the Old Town of Leh, many of the more impressive secular buildings are no longer inhabited because their owners have chosen to move to newer, more comfortable modern constructions. For the architectural historian, restoration may pose its own problems. For example, NIRLAC’s inventory shows that mosques all over Ladakh have been rebuilt in ‘Turco-Iranian’ style in the last twenty years, and there is no indication that the earlier buildings were properly documented before they were pulled down. In other cases insensitive restoration using concrete and other modern materials may have destroyed the aesthetic integrity of older buildings without solving major structural problems. The preface to NIRLAC’s inventory rightly highlights the contemporary religious and social aspects of ‘cultural heritage’. For instance, it cites the example of a distinguished lama who queried the purpose of restoring a ruined temple with exquisite wall paintings if there were no monks to carry out daily rituals there (NIRLAC 2008a, vol. 1:v.). The various components of Ladakh’s cultural heritage have little hope of survival—except perhaps as museum pieces—unless they are valued and form part of the lives of the region’s contemporary inhabitants.

A Note on Personal and Place Names

In historical references most authors in this book have given the Wylie transliterations of personal names and place names, in some cases accompanied by a modern phonetic rendering. The spelling of place names has often proved problematic. In many cases there is no consensus among Ladakhi scholars either on their etymologies or on their correct rendering in Tibetan script. Similarly, there is little uniformity in the Romanized spellings used for phonetic renderings in contemporary Western-language publications. We have therefore had to make a series of pragmatic choices without being able to claim

Introduction

13

complete consistency or uniformity across the papers, some authors showing preference for phonetic transcription, others for transliteration. A few examples illustrate the point. The name of the modern capital of Ladakh is rendered as Sle or Gle in Tibetan script (using the Wylie transliteration), with the latter becoming more common in recent times. Its English pronunciation should more plausibly be rendered as ‘Lé’, but ‘Leh’ has been the most common spelling in European texts since the 19th century and authors in this book follow what they regard as an established convention. In the many cases where there is still no consensus, we have had to make a series of somewhat arbitrary choices, adopting the version that seems to be in most common usage, for example ‘Thikse’ rather than ‘Thiksay’ or ‘Tikse’ for the place name Khrig se / Khri rtse. It would be a welcome step forward if Ladakhi scholars and administrators could agree on a set of standard Romanizations for place names in accordance with a consistent set of linguistic rules. In principle, the task of standardizing existing English phonetic transcriptions of Ladakhi words should not prove too difficult, since there are already two Ladakhi dictionaries (NorbergHodge and Thupstan Paldan 1991:188, and Hamid 1998:277) including entries in Tibetan script with the corresponding English phonetic transcriptions and translations. According to such dictionaries, for instance, the English phonetic rendering of the vowel sound in a place name like Shel is not ‘Shey’ as in ‘grey’, which would include a final diphthong hardly ever found in Ladakhi pronunciation, but simply ‘Shel’, in this case actually identical with the transliteration of the word. Nevertheless, realistically, we may have to accept that there is unlikely to be a consensus on these matters for some time to come: even these two dictionaries report different phonetic transcriptions for the same term, one giving for instance ‘markalak’ and the other ‘markalag’ (cf. Norberg-Hodge & Paldan 1991:132 and Hamid 1998:194). Future scholars working in this area might draw inspiration from the “THL Simplified Phonetic Transcription of Standard Tibetan” of the The Tibetan & Himalayan Digital Library,3 for instance adopting ‘é’ endings for place names such as Tiksé and avoiding inaccurate English phonetic renderings such as ‘Tiksay’ or ‘Tiksey’. However, it should be noted that the transliteration and pronunciation rules that apply to Central Tibet cannot simply be transferred to Ladakh. For example, the descendants of the kings of Ladakh reside in a village whose name is written and transliterated as ‘Stog’ or ‘sTog’ (again using the Wylie system). A Central Tibetan reading the written version of this place 3 See: www.thlib.org/reference/transliteration/#!essay=/thl/phonetics/.

14

Lo Bue and Bray

name would pronounce it as ‘Tok’, but Ladakhis say ‘Stok’, and presumably have done for centuries. In many cases the search for a Tibetan etymology may be misleading anyway. The La dvags rgyal rabs writes ‘dKar skyil’ for the site now known as ‘Kargil’ (Francke 1926:128), and this certainly looks ‘Tibetan’ even if its precise meaning is unclear. However, local oral tradition suggests that a ‘Dard’ named Kargi founded the original settlement and that it was earlier known as ‘Kargilo’ (dKar gyil lo), using the ‘Dard’ suffix ‘-lo’ to mean ‘Kargi-place’.4 As this example shows, the favoured Tibetan literary spelling of a place name at any one time does not necessarily bear any direct relation to its historical or linguistic origin: in geo-cultural Tibet, as well as in many other regions, bogus etymologies have often been created to suit shifting cultural, religious and political fashions. In this as in other respects, contemporary Ladakh reflects diverse linguistic and cultural influences rather than a single source.

4 The etymologies of Ladakhi place names are among the many topics that merit further investigation. We are grateful to Rebecca Norman for her observations on this topic, and to Bettina Zeisler for her own views and the information on ‘Kargilo’ which draws on local sources in Khalatse and Kargil. On this point, Francke (1926:253) cites the spelling ‘dKar gyil lo’ in a Khalatse villager’s account of the Dogra wars. He adds that the ‘munshi’ who wrote down the villager’s oral testimony “contrived to embellish it with as many classical Tibetan words and phrases as he thought necessary, to make the account acceptable to educated men”. In Ladakh, Western writers have applied the term ‘Dard’ to the inhabitants of villages of Dha and Hanu who speak a dialect of Shina. On the issue of the ‘Dard’ presence in Ladakh see Clarke (1977).

Ancient Petroglyphs of Ladakh: New Discoveries and Documentation Tashi Ldawa Thsangspa Although Ladakh appears relatively isolated today, it was for hundreds of years a transport corridor for traders and pilgrims moving between South and Central Asia. These travellers, together with Ladakh’s early inhabitants, left their marks in the form of the thousands of petroglyphs—figures or symbols pecked and chiselled into rock surfaces—that are scattered across the region. Many of the most interesting petroglyphs are to be found on or near the main transport routes, for example near the sites of bridges or river crossings. They constitute Ladakh’s oldest art form.

Figure 1.1 Rock with petroglyph images of people and animals beside the river Indus near Domkhar. Photo by the author.

© koninklijke brill nv, leiden, ���4 | doi ��.��63/9789004271807_�03

16

Thsangspa

As will be seen, the first scholarly references to petroglyphs date back to the late nineteenth and early twentieth centuries. However, the study of Ladakhi rock art remains in its infancy, first because earlier studies have documented no more than a small fraction of the petroglyphs that survive, and secondly because there remain many challenges of interpretation. This article draws on more than fifteen years of research to document the petroglyphs of Ladakh in both well-known and newly discovered sites. It begins with a brief overview of earlier research in Ladakh and neighbouring regions before presenting a small sample of Ladakhi petroglyphs arranged by theme. The petroglyphs are illustrated with tracings and a selection of photographs. As will be seen, their dating remains a challenge. However, many of the images that I have selected date back to the prehistoric era, well before the introduction of Buddhism into the region.

Earlier Studies of Rock Art in Ladakh and Neighbouring Regions

The Moravian missionary August Herman Francke (1870–1930) was the first Western scholar to write about Ladakhi rock art in two journal articles written in 1902 and 1903.1 In accordance with his historical interests, Francke was primarily interested in textual rock inscriptions, but also noted drawings of animals, stupas, hunting scenes, and what appeared to be local deities in and around Khalatse in Lower Ladakh and in Zangskar.2 Giuseppe Tucci rediscovered much of the rock art identified by Francke on the route between Leh, Alchi and Khalatse during his 1935 expedition: a selection of his photographs was republished by Giacomella Orofino (1990). In view of the damage caused by recent road-building operations along this route, these photographs constitute a valuable historical record. Since the 1990s there has been increasing interest in the rock art of the region. Vohra (1993, 1994, 1995, 1999, 2005) made several reports from Lower Ladakh and Nubra. Fonia (1993) reports on rock art as well as Paleolithic and Neolithic sites from Ladakh. Mani (1998–1999) reports on rock art from Lower Ladakh, Nubra and Zangskar. Francfort et al. (1990, 1992) give a valuable analytical report of petroglyphs spreading from the interior regions of Zangskar to Leh and Lower Ladakh, and suggest a number of comparisons with other

1 Karl Marx (1897), an earlier Moravian missionary, had earlier reported on the existence of petroglyphs in a travel account of a medical missionary tour near Achinathang. 2 See, for example, Plates 1–3 in Francke (1923).

Ancient Petroglyphs of Ladakh

17

regions. More recently, Vernier (2007) and Bruneau (2010) have been engaged in a longstanding project to document petroglyphs across the entire region. One of the most important emerging but still tentative themes from recent research concerns comparisons with both neighbouring and more distant regions. It should perhaps come as no surprise to find that petroglyphs comparable with those in Ladakh are to be found in western Tibet (Bellezza 2004). Similarly, in Burzahom (Kashmir) a petroglyph on a slab of rock from the Neolithic period also shows similarities with rock art in Ladakh (Sonwane 1992). However, by far the most important and thorough recent research comes from what are now the Northern Areas of Pakistan. Since the 1970s, a team of Pakistani archaeologists working together with the Heidelberg Academy of Humanities and Sciences (Germany) has been conducting a major research project around Chilas (Jettmar & Thewalt 1985, 1987). In a recent overview drawing on decades of research in northern Pakistan, the German scholar Harald Hauptmann (2007) points to comparisons with rock art in Central Asia. In particular, he identifies affinities between mask-like representations in northern Pakistan and similar motifs from the Okunev culture in the Charkassian-Minusinsk basin of the Altai (Southern Siberia), attributed to the end of the third and the beginning of the second millennium BCE.3 Similarly, he highlights a set of carvings in the ‘Eurasian animal style’ which he attributes to Scythian or Saka Iranian-speaking nomads who crossed into the region in the first millennium BCE. As Hauptmann notes, similar petroglyphs can also be found in Ladakh, and these will be discussed below.

Dating and Common Classifications

So far we have no means of dating petroglyphs precisely. However, comparisons between different sets of petroglyphs often make it possible to decide which is older and which is more recent. One set of clues comes from repatination or ‘desert varnish’, the change in surface coloration that takes place when a rock is exposed to the external environment. When a rock is scratched or pecked, the new surface begins to oxidize and starts darkening. The more the colour resembles the original rock surface, the older the art is likely to be. The superimposition of one set of images on another also indicates the relative dates of different pieces of rock art on particular boulders. Petroglyphs may be classified according to their style and subject matter. For example, the bodies of some figures are fully pecked. In other cases, only 3 On Siberia see also Brentjes (1999).

18

Thsangspa

the outlines are visible, while the main bodies of the images are kept blank. As will be seen, there are many anthropomorphic images in Ladakh, as well as depictions of animals. Some petroglyphs are geometric figures: others may be symbols of religious or magical origin. Some signs or ‘tamgas’ may be linked to particular individuals, families, clans or ethnic groups. We can also classify much of this art according to the presumed artists. As will be seen, some petroglyphs seem to be the work of travellers from outside the region who drew forms or figures that are clearly exotic, while other figures are more clearly linked to the local environment.

Masks and Giants

Some of the earliest rock art in Ladakh appears to date back to the Bronze Age, and a site at Sasoma in Nubra is one of the most important examples. One of the rock surfaces at Sasoma has six rounded figures (Figure 1.2). These are heavily repatinated, which partly accounts for their faintness. The two lowest figures appear to show human mask-like images. They have some resemblance to mask-like figures found in other parts of Ladakh and further afield. For example, Vohra (1993) discovered another masklike figure in Gya, while Francfort et al. (1990, 1992:Figure 5) report one from Zangskar. Similar masks, where the face is cut into quadrants by diagonal lines have also been found at Ziarat I and Chilas in northern Pakistan and in Central Asia (Jettmar & Thewalt 1987:12). Hauptmann (2007:25) argues that the Upper Indus masks have direct analogies in masks belonging to the Siberian Okunev culture in the Minusinsk basin of Altai from the late 3rd millennium BCE. He adds that this clear connection to the northern fringe of the steppe belt shows that the Upper Indus valley formed a part of the interaction zone between Central and South Asia. The Sasoma mask is closer to the images from Siberia than the ones from Chilas, Zangskar or Gya. Similarly, Hauptmann reports that in northern Pakistan there are some fifty ‘giants’—petroglyphs of male anthropomorphic figures with outstretched arms which he likewise attributes to the Bronze Age. As will be seen below, human figures depicted in petroglyphs are usually much smaller, whereas the giants are often as much as 1.5 metres high. Again, similar figures are found scattered across Ladakh. One in Tangtse village has a large head with about eight raised cone-like structures and two long antenna-like structures protruding from the two ears on the sides. The details of the face are not legible. The hands have raised forearms with fingers prominently displayed. The legs

Ancient Petroglyphs of Ladakh

Figure 1.2 Masks from Sasoma Nubra. Photo by the author.

19

20

Thsangspa

b.

a.

c.

e. d.

Figure 1.3 a. Mask found in Sasoma (Nubra), Ladakh; b. Mask from Gya, Ladakh (Vohra 1993); c. Mascoid from Char. Zanskar (Francfort et al. 1990. Figure 5); d. Masks from Chilas on the Indus (Francfort et al. 1990. Figure 6); e. Okunev Culture maskoids from Mugur Sargol in Siberia (Ibid.).

show a squatting position; between the legs there might have been some more details, but it is no longer possible to make these out. On both sides of the main figure there are some rounded circular figures. John Bellezza (1999) has published a similar image from northern Tibet in a posture that he compares to depictions of ‘cosmic birthing goddesses’, whoever they may be. Another giant-like figure (Figs 1.4, 1.5a) in the Tangtse area has similar outstretched arms, while a slightly different one can be seen in Bema, in Lower Ladakh (Figure 1.5c), and there are others—not illustrated here—in Domkhar (Lower Ladakh) and Sasoma (Nubra).

Human Figures

Human figures are definitely a dominant theme in the rock art of Ladakh. They may be shown as individuals or in groups, but are almost always involved in activities such as hunting, dancing, fighting, and conducting rituals. The images below present a small sample.

Ancient Petroglyphs of Ladakh

Figure 1.4 The Tangtse giant. Photo by the author.

21

22

Thsangspa

Figure 1.5 a, b Giant figures from Tangtse; c. Giant from Bema.

a.

b.

c.

d.

g.

k.

e.

h.

i.

f.

j.

l.

m.

23

Ancient Petroglyphs of Ladakh

n.

o.

p.

q. Figure 1.6 a–c. Giant figures from Chilling valley; d–n. From Lower Ladakh; o. From Tangtse (Changthang); p. From Karu; q. From Kargil.

The figures from the Chilling valley (Figure 1.6 a–c) are outline-pecked. However, as with the figures from Lower Ladakh (Figure 1.6 d–n), it is more common to find petroglyphs that are silhouette-pecked. The figures in the Lower Ladakh images are commonly shown with enlarged, raised hands: we have no means of knowing for certain, but it is possible that this gesture is associated with some ritual. Figs 1.1p and 1.1q show human-like figures in association with other figures or symbols, but these are hard to interpret. Many petroglyphs show a range of social activities. The most common subject is hunting with bows and arrows, and often with dogs. Other activities include dancing, fighting and rituals. The hunting scenes in Figs 1.7a,b,c,d show the prey as much bigger than the hunters, and this may have some magical significance, perhaps as part of a ritual to ensure a successful hunt, or as an expression of metaphysical beliefs (Sonawane 1992:280). Figure 1.7e likewise shows a large animal, probably an ibex, surrounded by anthropomorphic figures. Figure 1.7f appears to depict a figure engaging in some sort of ceremony. Another strange composition in Figure 1.7i shows a man on horseback together with a second horse, a feline and a counter-clockwise swastika, a symbol which is associated with—among other traditions—the ancient Bon religion. Figure 1.7j from Tangtse shows a man with a bow and arrow. The man’s dress resembles that of a West Iranian warrior from the Alter rock, near Thalpan Bridge

24

Thsangspa

b.

a.

c.

e.

d.

f.

g.

h.

i.

j.

Figure 1.7 Figures showing social scenes: a–b. from Chilling valley; c–g. from Lower Ladakh; h. from Nubra valley; i. from Kargil; and j from Tangtse.

Ancient Petroglyphs of Ladakh

25

Figure 1.8 A group of people with raised hands and fingers, together with some animals. Half of the rock is buried under the road debris on the Bema-Domkhar road. Photo by the author.

in northern Pakistan, which is dated to the first millennium BCE (Jettmar & Thewalt 1987, photo 6, plate 5).

Animal Representations

The most common animal figures are ungulates such as Siberian ibexes, blue sheep, urials, antelope and yaks: all of these either are or have been common in Ladakh. Horses, camels, dogs and birds are also seen, but less frequently. The animals are represented in a wide range of different styles, and these may offer tentative clues as to their origins. Figure 1.9 shows a group of felines from Chilling: the outlines of the bodies are filled with dots, and the animals have small heads and long tails. The style of these images is unique in Ladakh: it includes a small geometric design at the bottom of the figure (enlarged in Figure 1.9b), and this is without parallel elsewhere in the region.

26

Thsangspa

Figure 1.9 a, b. Images from Chilling.

a.

b.

b.

a.

Figure 1.10

c.

Animal figures from the Nubra valley.

d.

Figures of horses (Figures 1.10a, b) with raised manes and erect penises are a peculiarity of the Sasoma site in the Nubra valley. Figure 1.10c shows ibexes and yaks accompanied by some sort of symbol whose meaning is obscure. Figure 1.10d again shows a group of ibexes, some with extended penises. Ibexes with exaggeratedly long horns are evidently of some cultural significance, although their precise meaning is uncertain. Similar figures have also been shown with exaggeratedly extended penises in northern Pakistan (Jettmar & Thewalt 1987: Photo 53).

27

Ancient Petroglyphs of Ladakh

The ‘Domkhar Sanctuary’ is a constellation of rock art on the right bank of the river Indus in Domkhar village in Lower Ladakh. The closeness of the two banks along with the rocky formation of the Indus at this particular point suggests that there may have been a bridge or a temporary crossing place here. The shiny and highly patinated surface of the Domkhar rocks gives the figures a strong contrast. At the same time, there is a high degree of repatination which points to the figures’ antiquity, and makes it almost certain that they belong to the pre-Buddhist period of Ladakhi history. Together they form one of the most beautiful and aesthetically appealing assemblages of petroglyphs found in Ladakh to date. The most unusual feature of the Sanctuary is the stylistic representation of the animal figures. Many of the animals, which include horses, ibexes and felines, are depicted in motion or on tip-toe. The interiors of the bodies show a peculiar pattern of volutes and curves (Figure 1.11d is accompanied by a Chinese inscription, which is unique in Ladakh—see Fig. 1.12—but it seems to have been added by a different hand, and may be somewhat later than themain image). The long-horned ungulate is undoubtedly a deer since, as Bellezza (2004) points out when discussing a similar image from northern Tibet, no other animal in the region has comparable antlers.

b.

a.

e. Figure 1.11

Animal figures from Domkhar Sanctuary.

c.

f.

d.

g.

28

Figure 1.12

Thsangspa

Deer with Chinese inscription at Domkhar. Photo by the author.

This style points to cultural connections with the Scythian (Saka) nomads who spread out in the Eurasian steppes in the first millennium BCE.4 For example, images of animals that seem to be ‘standing on tip-toe’ are frequent in the Early Iron Age art of eastern Kazakhstan (Hauptmann 2003:5–6). Scythian art based much of its artistic composition on motifs with symbolic meanings. The characteristics of Scythian art include: attack scenes of predators upon prey, lively depiction of animals, exaggerated formation of antlers on deer, and animals with heads facing backwards over their shoulders. The magical use of symbols may have been intended to guarantee the power of the people who had them made. In some cases a particular animal may have been a clan symbol, a mark of tribal identity. One of the most artistically drawn animal figures—which also has Scythian characteristics—is to be found amidst some big boulders on the outskirts of Tangtse in north-east Ladakh. The figure is drawn on a rock surface that is almost vertical and faces approximately south, with a rough surface. The subject is a feline chasing two deer (Figs 1.13, 1.14). A second deer to the right seems to be part of the same chase scene. 4 See Dani (1989:118–123), Bellezza (2004), Francfort et al. (1990:152), Hauptmann (2007:12–13).

Ancient Petroglyphs of Ladakh

Figure 1.13

The Tangtse chase. Photo by the author.

Figure 1.14

Tracing of the Tangtse chase.

Figure 1.15

A feline chasing deer from Renmudong, Tibet, attributed to the Iron Age (Francfort et al. 1990:Figure 12).

29

30

Thsangspa

Figure 1.16

The deer in Renmudong. Courtesy of the Mission archéologique française en Asie centrale (MAFAC) archive.

The Chinese scholar Wu Junkui (1987) has found a very similar composition with felines chasing deer at Renmudong in the Tibetan Changthang (east of Lake Panggong, near Rudok), and attributes it to the Iron Age. The Changthang chase is so similar to the Tangtse one that we could with some degree of plausibility imagine the artist as belonging to the same clan or tribe repeating similar compositions in different places. Images of felines or other carnivores chasing their prey, particularly deer, sheep or goats, are common in a number of different cultures. For example, depictions of felines chasing deer are widespread in Central Asia’s rock-cut art during the Scythian/Saka period (Francfort et al.:154). Francfort et al. (1990:151) also point to close parallels between Zangskar petroglyphs of deer with their heads turned back with similar images on jade plaques or pendants from Western Zhou (China) from the ninth century BCE.

Signs and Symbols

The petroglyphs on the four prominent boulders in the middle of Tangtse village in the Changthang region of Ladakh are among the best known in Ladakh. They include inscriptions in Arabic, Sogdian, Tibetan, Brahmi and Sharada

31

Ancient Petroglyphs of Ladakh

f.

g.

l. Figure 1.17

c.

b.

a.

h.

i.

d.

j.

e.

k.

m.

A selection of signs and figures from the Tangtse boulders.

written by travellers, traders, monks and missionaries. One of the Sogdian inscriptions is dated to around 841–842 CE.5 Besides these inscriptions, the boulders also have many interesting signs as shown in Figure 1.17, including the Nestorian Christian crosses in Figs d–f. The meaning of most of the signs remains tantalizingly obscure. Other signs and symbols are to be found on rock art elsewhere in Ladakh as shown in the following images. Again, their precise meaning is hard to interpret.

Discussion and Conclusion

This essay has presented no more than a small fraction of Ladakh’s petroglyphs, but it is to be hoped that it has been sufficient to demonstrate their historic and artistic importance, and to point to the many unsolved problems of interpretation. Realistically, we must accept that the meaning and precise significance of images created in the remote past is likely to remain obscure. Nevertheless, 5 The Sogdian carvings, and especially the crosses, have been widely discussed since they were first noticed by Moravian missionaries in the early 20th century. For the most recently discussions see: Sims-Williams (1993), Sander (1994) and Vohra (1994).

32

Figure 1.18

Thsangspa

Nestorian cross with Sogdian inscription in Tangtse. Photo by the author.

there are still many clues to pursue. The first concerns a closer analysis of the precise locations of the petroglyphs. As noted above, many are to be found along the historic and pre-historic trade routes of the region, for example at the Indus river crossing near Khalatse. They bear witness to the numbers and cultural diversity of travellers who have passed this way in previous epochs.

33

Ancient Petroglyphs of Ladakh

a.

d.

b.

c.

Figure 1.19

Signs in Bema (Lower Ladakh); b–d Kere-Kharu.

Figure 1.20

Sign of unknown significance in Kere Kharu. Animal figures to bottom left. Photo by the author.

34

Thsangspa

Other sites may have some magical/religious significance. It is possible that future archaeological research will point to links between certain groups of petroglyphs and burial sites. Insights from ethnologists may yet provide further clues to the meaning of the many images that appear to serve some ritual purpose. On a related note a closer analysis of the hunting scenes may provide some clues on the transition from pastoral to more settled societies, or the extent to which both styles of livelihood existed side by side. A second line of enquiry concerns the physical examination of the petroglyphs. By examining ‘palimpsests’—cases where one set of images is imposed on top of another—it may be possible to establish a clearer chronology for different styles. Similarly, a closer study of the process of repatination may make it possible to establish a more accurate system of dating. Thirdly, there is ample scope for regional and inter-regional comparisons. It is striking to observe how even the earliest rock art—for example the masks and ‘giants’—have analogies in what is now northern Pakistan and much further afield in Central Asia. Similarly, the ‘tip-toe’ animal representations point to a historical connection with the Scythian cultures of central Asia in the first millennium BCE. One of the main themes of this book concerns the interplay of internal and external cultural influences in Ladakh. It is clear that this process began in prehistoric times. Finally, it should be emphasized that, despite the hundreds of new petroglyphs discovered in recent years, the task of documentation is far from complete. This task is all the more urgent because of the recent destruction of rock art all over Ladakh as a result of modern development. The fact that petroglyphs are so often to be found along the ancient transport routes means that they are more vulnerable to modern road-building. The rock art sites near the ancient river crossing in Khalatse provide a poignant example: the old wooden bridge has been replaced successively by steel and concrete bridges. The two modern replacements were at different sites along the bank. As a result of their construction, some of the most significant petroglyphs and rock inscriptions in the region have been pulverised. Elsewhere rocks have been broken down for construction purposes, or vandalized with modern graffiti. Ladakh’s rock art bears witness to the experiences and preoccupations of its earliest human inhabitants. The testimony of thousands of years will be eroded and destroyed unless the region’s contemporary inhabitants learn to regard it with greater appreciation.

Embedded in Stone—Early Buddhist Rock Art of Ladakh Phuntsog Dorjay While discussing the history of Ladakh, one invariably slips into talking about how Buddhism came into being in the region. In other words, there is a thin dividing line between the wider history of Ladakh and the initiation and advancement of Buddhism. Moreover, it is difficult to look at Ladakh in isolation from what occurred in other parts of the Western Himalayan region. This is a reflection of the powerful impact of Buddhism, and the extent to which almost all political events revolved around its propagation. Limited information is available on the history of Buddhist art in Ladakh prior to the establishment of the Alchi group of monuments in the 11–13th centuries CE, as there has hardly been any precise documentation of facts and figures. There is also a lack of comparative information on the region and its neighbours. It is even difficult to ascertain when exactly Buddhism was introduced, as there are very few written sources from which the region’s early history can be reconstructed, while the available archaeological evidence cannot yet be securely dated. Since the earliest Buddhist heritage of the region is in the form of rock carvings, a comparative study of rock sculptures and inscriptions, as well as of contemporary trade relations and routes, can throw some light on early art-historical aspects of the region. The major rock carvings of the region have been noted since the colonial period when Western explorers and administrators began to visit Ladakh. Moorcroft’s Travels (1841) and Cunningham’s Ladak (1854) mention the Mulbek Maitreya and the Dras sculptures. However, the earliest scholarly reference to the rock carvings was made by Francke in Antiquities of Indian Tibet (1914), followed much later by Petech in The Kingdom of Ladakh (1977) and Snellgrove & Skorupski in The Cultural Heritage of Ladakh (1977 and 1980). More recent arthistorical analysis acknowledging Kashmiri influence in the early temples of Ladakh was made by Pratapaditya Pal in A Buddhist Paradise: The Mural of Alchi, Western Himalaya (1982) and by Roger Goepper in Alchi: Ladakh’s Hidden Buddhist Sanctuary (1996). In this article I analyse the various rock carvings mentioned by these scholars and travellers as well as presenting new findings of my own. This work is based on a more precise documentation and a comparative study between Ladakhi stone sculpture and Kashmiri bronzes. The idea for such a comparison

© koninklijke brill nv, leiden, ���4 | doi ��.��63/9789004271807_�04

36

Dorjay

had been suggested before (Luczanits 2005:68), but no serious study to that effect had been carried out. I conducted a large-scale documentation of rock sculptures in Kargil and Leh districts, as well as in the Nubra valley, during the years 2001 to 2008. The stone carvings here taken into consideration may be grouped into the following geographical areas of investigation (see map, Figure 2.1): Kargil area Dras, Mulbek, Tumel, Kartse and Byama Khumbu. Leh area Khalatse, Alchi, Saspol, Nimo, Skyu, Taru, Leh , Shey and Sakti. Nubra area Digar, Hundar, Tirith and Sumur.

Early Rock Art

Prehistoric petroglyphs are found throughout Ladakh. The carvings are pecked or chiselled into the dark brown varnished surface of boulders scattered on riverbanks and valley terraces. Dani concludes that two things are evident: that the examples of rock art, spread all over the Western Himalaya/Karakoram region, are interrelated; and that there is a continuity in this rock art from early times up to the historic period in such a way that we may well speak of a language of cultural continuity from prehistory to the early mediaeval period, where more sources of study are available to us (Dani 1982). Prehistoric rock carvings of various stages are found in Ladakh, although I have not been able to date them. The earliest group appears to include simple line drawings of ibex and human figures, representing geometric forms only. Scenes showing hunters with arrows, individually and in groups, are also found, implying that the earlier people who inhabited Ladakh were hunters and food gatherers that preceded Mons and so-called Dards, who practised agriculture. The discovery of a vast quantity of antiquities by the ‘Pak-German Study Group’ in the upper Indus and adjoining parts of the northern areas of Pakistan began in 1979, and to date a total of nearly 50,000 rock carvings and 6,000 inscriptions have been recorded (Hauptmann 2005:23). It is difficult to ascertain precisely when Buddhism was introduced to Ladakh, but it may safely be said that the dwellers of Ladakh came into

Figure 2.1 Map showing the possible early trade routes and important Buddhist rock art sites. Illustration: Phuntsog Dorjay.

Embedded in Stone

37

38

Dorjay

c­ ontact with Buddhism following the Tibetan invasion of Zhang Zhung and after Buddhism had been adopted as the official religion of Tibet, in a period corresponding roughly to 760–840 CE (Beckwith 1987:163). In Ladakh there are several petroglyphs inscribed by Tibetan troops in Ladakh who were stationed at frontier posts (Denwood 1980). The engravings are represented by stupas (mchod rten) with or without Tibetan inscriptions mentioning names of persons, some of them Buddhist, and years from the Tibetan calendar. Thus it is clear that by this time Ladakh, including the adjoining areas now under Pakistan, was occupied by Tibetan people. However, in that early period their culture can have made no real impact on the indigenous population. At that time Tibet itself was receiving Buddhist ideas from mainland India, particularly from Kashmir, and the rock sculptures to be discussed in this paper are certainly not of Tibetan origin. The discovery of inscriptions in Kharoshti and Brāhmī script at Dras and Khalatse has not been decisive in assessing whether Ladakh was part of the Kushana empire (second-third centuries CE) (Francke 1994, Goepper 1996:11), although Kushana influence in the upper Indus valley of northern Pakistan, with which Ladakh shares a long history, is evident (Hauptmann 2005). Kashmiri Influence The north Indian, particularly Kashmiri, influence on the rock cut reliefs discussed in this paper may be deduced through a comparison of the rock sculptures with the available Kashmiri bronzes of the ninth to twelfth centuries or even later, for which more sources are available, at least for a comparative documentation. A study of the early Buddhist rock sculptures of Ladakh shows striking resemblances with Kashmiri bronzes. The workshops of Kashmir were active in bronze casting from at least the seventh to the twelfth centuries CE: they were renowned both for the high quality of their output and for the wide range of image sizes they produced. Unfortunately, none of these larger images have survived. What has survived are the smaller, portable images, many of which found their way into the monasteries of Tibet and Ladakh. Tibetan pilgrims and scholars who travelled to Kashmir brought several of these images back home with them (Heller 1999). Kashmiri craftsmen are known to have been active in Tibet, where their craftsmanship was valued highly, and the subsequent influence of the art of Kashmir upon that of western Tibet is obvious. The special features of Kashmiri bronzes (Figure 2.2) include tough bodies with large shoulders, strong breasts, and abdominal muscles very similar to the sculptures of the Alchi Sumtsek. Among the other typical features are arched eyebrows, a small mouth, a band of pearls under the crown, copper

Embedded in Stone

39

Figure 2.2 Avalokiteśvara. Brass, silver and copper, height 100 cm, Tibet, C. 11th century CE. Pritzker Collections. Photo: Hughes Dubois.

strips inserted in the dhoti and silver in the eyes. Sharp-pointed diadems and lotus leaves are among the distinctive features of the Kashmiri bronzes from Tibet and scholars believe that these were made by Kashmiri artists in Western Tibet (Heller 2007:84, Pal 2003). An early trade route must have existed along the river Indus leading from northern India into Central Asia for centuries, and there were good ­political

40

Dorjay

and cultural contacts between Kashmir and Ladakh. India in general and Kashmir in particular were connected with the southern silk route across Central Asia, and trade took place with the north mainly via three routes. Two of these passed through Gilgit and the Yasin valley up to Tashkurgan where they joined the Kashgar route. The third passed through Ladakh and then split, reaching either Khotan or Yarkand and Kashgar (Warmington 1947). The fact that the village of Khalatse (100 km west of Leh) was situated on a trade route from the time of the Kushanas is proved by the inscriptions mentioned above. That route was important because it was the gateway that led further on to Tibet and other countries in Central Asia. It is therefore obvious that in the early mediaeval period this route must have served as a trade route via the Zoji-la pass to Ladakh and from there to Central Asia, via Yarkand or Khotan. This route must have achieved greater significance when India’s trade relations with Central Asian countries and the West via the Silk Route were disrupted because of the Huna invasion in the fifth century, followed by the rise of the Arabs and the Ghazanavid conquests (Jha 2000). It is therefore highly probable that Buddhist missionaries and pilgrims during and since that time also used the same route. After the fifth century, the Indus valley thus became part of the wider system of the Silk Routes, and various side-valleys could be used to access and exit the main paths (Jettmar 2002). The discovery of several rock carvings bears witness to this fact. Ladakh shares a common language with Baltistan, and its direct communication with that region, mainly via the readily accessible route along the Indus, played a more important role than any other cultural and political links with its neighbours. The three Buddhist rock reliefs from Babur (Baltistan) published by Jettmar, and the famous ninth century Buddha reliefs from Naupura near Gilgit and Manthal near Skardu, share stylistic similarities with the reliefs found in Ladakh, notably in the Kargil region (Jettmar 2002, Hauptmann 2005). Fine examples of direct Indian Buddhist influence in Ladakh include the rock carvings by the wayside at Dras, which depict the Future Buddha Maitreya, a lotus flower, a horseman and the Bodhisattva Avalokiteśvara. Those reliefs represent the most important traces of pre-Tibetan—i.e. direct IndianBuddhist influence in Ladakh. The Buddhist images are clearly of Indian origin and, for iconographic as well as historical reasons they are unlikely to be much earlier than the ninth century CE. Rock carvings of the later phase of the early historical and early mediaeval period (the tenth to eleventh centuries CE) are found at Byama Khumbu, Kartse Khar, Tumel, Sani, Stongde and Padum. The early rock reliefs representing Avalokiteśvara flanked by two goddesses near the village of Byama Khumbu

Embedded in Stone

41

Figure 2.3 Rock carvings at Byama Khumbu (Suru valley), depicting Avalokiteśvara as main figure flanked by gandharvas and lay devotees below. Photo: Phuntsog Dorjay.

Figure 2.4 Detail of lay devotees at Byama Khumbu. Photo: Phuntsog Dorjay.

(about 33 km south of Kargil, see Figure 2.3, 2.4) must have been executed prior to the foundation of the first Tibetan monastic complex of Nyarma in the early 11th century CE (Tucci 1988b:64). A male devotee wears a heavy simple tunic with a belt and a lady a multi-layered skirt. The dresses of the devotees are very similar to those worn by local worshippers in a recently discovered old wall painting from the neighbouring region of Baltistan (Hauptmann 2005), and similar to those worn by the lay devotees in the carvings of the five Tathāgatas at Shey (see below). The same kind of dress seems to have been worn by local

42

Dorjay

nobles and laymen in the Indus belt from Baltistan to Kargil and the upper Indus areas of Ladakh, during the period that the reliefs were carved. The most impressive carving in Ladakh, apart from that at Mulbek, is situated in Kartse (Figure 2.5), near the village of Sanko in the Suru valley, on the way from Kargil to Zanskar. This is the tallest statue carved out of live rock in Ladakh, more than seven metres high. The sculpture is flanked by two flying gandharvas on either side of the head (Figure 2.6). The triple sharp-pointed diadem is characteristic of Kashmiri sculpture, and the pearl bands under the crown are typical of those seen in Kashmiri bronzes. The small twists of fabric extending from the crown and drape to shoulder level, more clearly seen on the left shoulder, derive from the pativ, a Persian emblem of royalty adopted by Kashmiri Buddhists (Heller 1999). The figure wears a lower garment very similar to an Indian dhoti. Another Maitreya sculpture (Figure 2.7) from the early period can be seen near the road to Batalik at the village of Tumel, some 25 kilometres north of Kargil. The figure, unnoticed by the scholars and travellers mentioned above, is carved in the same fashion as at Kartse and Mulbek (see below), although it is smaller in size. It too has features similar to the Kashmiri bronzes, with a tough body, broad shoulders and strong chest, as well as crossed navel and abdominal muscles, very similar to the clay sculptures in the Alchi Sumtsek. The four-armed standing figure of Maitreya (Figures 2.8, 2.9), about seven metres high, at Mulbek has been noticed by almost all scholars and travellers, as it lies along the road from Srinagar to Leh. Maitreya can be recognized principally by the stupa lodged in his hair, a common attribute of his. There are some carved lay devotees, standing in rows and facing upward, on both sides of his legs to be seen (Figure 2.10). In addition, he holds a water pot in his left lower hand, a nāgakeśara flower in his upper left hand and a rosary in his right upper hand, the presence of the first attribute being emblematic in his Indian and Kashmiri Buddhist iconography (Bhattacharyya 1985:62). The sacred thread across his chest and simple scarf draped over his arms are signs of his caste affiliation. Maitreya has to be born into a Brahman family (Pal 1975), and Moorcroft and Francke rightly noticed the characteristic Brahmanical emblems of the figure (Moorcroft 1837, Francke 1994). His ornaments, consisting of a necklace, a pair of bangles, and a belt, are made of pearls. In any event, this image is represented in a form that was popular in tenth-eleventh century Kashmiri bronzes found in west Tibet that portray Maitreya and other Bodhisattvas (Heller 1999:62). These characteristics are also common in clay sculptures at Alchi, Sumda, Mangyu and Tabo, which are generally attributed to Kashmiri artists.

Embedded in Stone

Figure 2.5 Maitreya figure at Kartse (Suru valley). Photo: Phuntsog Dorjay.

43

44

Dorjay

Figure 2.6 Detail of Kartse figure. Photo: Phuntsog Dorjay.



Upper Indus Valley

Travelling eastwards from Mulbek one comes to the village of Khalatse where Francke discovered the inscription from Kushana times cited earlier (Francke 1994). Francke also noticed an early Brāhmī and a Gupta inscription (from the fourth and fifth centuries CE respectively) besides a number of Tibetan royal inscriptions around the castle and stupas. Many of the inscriptions and other carvings at Khalatse were destroyed during the construction of a new bridge in 2005–2006. As noted above, the fact that Khalatse was situated along a major trade route from the time of the Kushanas is proved by those very inscriptions. That route was important because it was the gateway leading on to Tibet and other Central Asian countries. A number of prehistoric carvings as well as early petroglyphs depicting stupas are found along the way on the route eastwards from Khalatse towards Leh in villages such as Nurla, Tingmosgang, Saspol, Alchi, Nimo, Basgo and Taru. Leh Leh seems to have developed into a trade centre from about the ninth century as suggested by the finding of rock reliefs, including early carvings, in and

Embedded in Stone

Figure 2.7 Maitreya figure at Tumel village (Kargil). Photo: Phuntsog Dorjay.

45

46

Dorjay

Figure 2.8 Famous Maitreya figure at Mulbek village (on Srinagar-Leh highway). Photo: Phuntsog Dorjay.

Embedded in Stone

Figure 2.9 Detail Maitreya sculpture, Mulbek. Photo: Phuntsog Dorjay.

47

48

Figure 2.10

Dorjay

Lay devotees at Mulbek. Photo: Phuntsog Dorjay.

around the town. It became a meeting place for traders and Buddhist masters travelling along different trade routes. The rock carvings found in the Leh area are based on Indian models, and are very similar to those found at Dras, Suru Valley and Mulbek. It is not possible to mention all the carvings in the present article, but some of the important ones are discussed below. Sangkar Beside the footpath from the main market of Leh to Sangkar monastery, there is a small enclosure which is open from the front and houses four standing sculptures, each on a separate stone. These are very similar to those found at Changspa, another suburb of Leh (see below), but have been painted recently with yellow, blue and white colours. According to the local people, these carvings were previously lying in the fields and it was only after the construction of the nearby Sangkar monastery, around the last quarter of the 19th century, that they were brought together and arranged in a row under this shelter, which was specially built for the purpose. All four carvings are a little more than one metre in height. Starting from the right, the first sculpture is Avalokiteśvara with a rosary in his right hand. The next figure is Maitreya in the abhaya mudrā posture; the third from the right is also Maitreya; and the fourth figure is similar to the first one, but with a slightly different hand gesture.

Embedded in Stone

49

Changspa There is a large stupa at Changspa in Leh, belonging to the bkra shis sgo mang (‘auspicious because of its many doors’) type. Francke assumed that it might go back to the time of Mons and Dards of c. 700–900 CE (Francke 1977:80), but its architectural style is similar to the Shey sgo mang mchod rten, built around the sixteenth century. There are two rock carvings in its immediate vicinity and another standing some 100 metres to the east of the stupa. Figures of Buddhist deities are carved on both sides of a rock about two metres high. The front side shows the Buddha, standing and displaying a teaching gesture while holding a vase in his left hand. On the reverse side of the same stone is a four-armed figure about one metre high (Figure 2.11). On the basis of its iconography, as it holds a lotus flower in his upper right hand, it is identified as Avalokiteśvara (Francke 1994). The lower right hand seems to hold a rosary, which is common in the iconography. Just below the Bodhisattva’s left hand there is a seated figure in the anjāli gesture (with the palms joined together in front of the chest, signifying reverence or submission), holding an utpala flower in two of his left hands. An antelope skin slung over Avalokiteśvara’s shoulder is the sign of his status as a reclusive ascetic, and also provides the mat upon which he sits in meditation. The same kind of sash, often found in Kashmiri bronzes from Western Tibet, can be seen on the shoulder of the Avalokiteśvara image in the Alchi Sumtsek. Another proof of this figure being Avalokiteśvara is the presence of the smaller adjacent figure of Hayagrīva (‘Horse-necked One’), with his horse’s head emerging at the crown of his head (Figure 2.12), holding a sword in his left hand. Hayagrīva is a primary emanation of Avalokiteśvara and may be regarded as the Bodhisattva’s archetypal fierce, dynamic and undying compassion (Rhie & Thurman 1991). This is the only rock sculpture of Hayagrīva that I have found in Ladakh: it has a striking resemblance with one found by Rob Linrothe in the east Indian region of Ratnagiri (Orissa), dated c. 9th century CE (Linrothe 1999). Leh Old Town Apart from many rock sculptures placed on number of stupas in Leh Old Town (Figure 2.13). There are some interesting sculptures which have been unnoticed by many. A group of five sculptures (Figure 2.14), dating probably from the period c. ninth–eleventh, centuries has recently been rediscovered in Sangkar House in Leh. The sculptures are slightly above one metre in height. The house has been recently been renovated by the Tibet Heritage Fund (Berlin). In all probability, the sculptures were lying outside prior to the building of the house. They appear to have been collected and put in a room on the ground floor,

50

Figure 2.11

Dorjay

Avalokiteśvara and Hayagrīva near bKra shis sgo mang chorten at Changspa (Leh). Photo: Phuntsog Dorjay.

Embedded in Stone

Figure 2.12

51

Detail of Hayagrīva or ‘horse-necked one’ (Tib. rTa mgrin), a primary emanation of Avalokiteśvara, Changspa. Photo: Phuntsog Dorjay.

52

Dorjay

Figure 2.13

Padmapāṇi (left) and Maitreya (right) on a stupa at Chute Rantak (Leh Old Town). Photo: Phuntsog Dorjay.

Figure 2.14

Five recently discovered rock figures in Sangkar House (Leh Old Town). Photo: Phuntsog Dorjay.

Embedded in Stone

53

probably by the owner (from the monastery of Sangkar) of the house, which seems to have been built after the construction of main Leh Palace in the 17th century. The sculptures appear to have been the object of much worship as the room has become dark with the smoke from oil lamps. In the summer of 2008 I photographed and documented these sculptures. Though a detailed study is still awaited, two of them, the one in the front of the entrance and the one in the left corner may be identified as the historical Buddha and as Mañjuśrī, the Bodhisattva of wisdom, respectively. Another figure identified as Maitreya (1.9 metres in height), was lying badly damaged in an open space near Sangkar House. The sculpture has recently been moved and put in front of Sankar House by the local people with the help of Tibet Heritage Fund workers (Figures 2.15, 2.16). Skara In Skara, south of Leh and near the airport, two rock steles have been found. A standing Maitreya (Figure 2.17), c.1.5 metres high and very similar to the one at Changspa, is carved on a smaller stone adjacent to a round stupa of a later period. The stone is carved on both sides with standing figures in abhaya mudrā. On the back Maitreya is depicted with a diadem and his lower garment, similar to a dhoti, is carved in deep relief. The second figure is situated about 100 metres south of this spot and only the upper portion of the figure is visible. An elderly inhabitant of the nearby Karpotok House said that during his childhood local people built a shed to protect the figure from rain and sun. That shelter is no longer there, but the stones of its collapsed wall have now covered the lower portion of the figure. Judging from the head and shoulders, it seems that the image is the same size as the above-mentioned rock carving. The figure can be identified as Padmapāṇi Avalokiteśvara wearing a three-pointed diadem and holding a lotus flower in his left hand. A sash made from antelope skin is slung over his shoulder. On the way to Choglamsar from Leh we have found a boulder carved with Buddhist sculptures on both sides. The sculptures appear to have been recently painted with colours by the villagers, which makes it difficult to assess its age (Figure 2.18). Shey Shey, about 15 km southeast of Leh, was an ancient capital of Ladakh. In the tenth century CE the first king of Ladakh, lHa chen dPal gyi mgon, apparently constructed the hilltop fortress whose ruins can be seen above the present Shey Palace. Shey possesses a number of early Buddhist rock sculptures, many of which are about a metre in height.

54

Figure 2.15

Dorjay

Badly damaged rock sculpture of Maitreya near Sangkar House (Leh Old Town). Photo: Phuntsog Dorjay.

Embedded in Stone

Figure 2.16

55

Tibetan inscription on the left hand side of Maitreya near Sangkar House (Leh Old Town). Photo: Phuntsog Dorjay.

The main carving of five Tathāgatas (Figure 2.19) at the lower edge of the mountain has frequently been mentioned by scholars (Snellgrove & Skorupski 1977). The carving, which is generally dated to about the tenth century, shows the five Cosmic Buddhas standing on thrones identified by their respective vehicles (vahanas) starting from left to right: Ratnasambhava, Akshobhya, Vairocana, Amitabha and Amoghasiddhi. Under a huge lotus supporting the throne, seated in rows and facing upward, are several male lay devotees. Their dress is similar to the Ladakhi ‘goncha’ (a long-sleeved robe worn with a sash) and on their heads is the Ladakhi hat (tibi). A similar dress is worn by the male lay devotees in the Avalokiteśvara carving at Byama Khumbu in Kargil. On the basis of style and the dress of the lay devotees, it can be assumed that the Shey carvings reflect the early stage of Buddhist rock art in the region, presumably, the early 11th century CE. A fine example of early rock carving is found near the sgo mang stupa in the field behind Shey palace. The figure is identified as Maitreya: it is finely carved with deep relief, and the Bodhisattva’s left hand displays the ubiquitous gesture of charity, grasping the long stem of the lotus flower that rises from the base with a vase on top of it. The lower garment resembles the Indian dhoti worn by the Maitreya of Mulbek. The figure wears a three-pointed diadem, framed by a nimbus and aureole. In many ways it resembles the Kashmiri bronzes of

56

Figure 2.17

Dorjay

Maitreya relief at Skara village, Leh area. Photo: Phuntsog Dorjay.

Embedded in Stone

Figure 2.18

57

Standing Maitreya at Tyare Rong, recently painted by the villagers of Diskit-tsal (Leh area). Photo: Phuntsog Dorjay.

58

Figure 2.19

Dorjay

Five Tathāgatas at Shey. Photo: Phuntsog Dorjay.

Bodhisattva Avalokiteśvara, dated c. 1000–1050, published by Pal (Pal 2003–4). The work relates generally to a large group of Kashmiri bronzes believed to be from the tenth to eleventh centuries (Heller 1999). About 500 metres east of the palace, near the lHa rDo rje chen mo chapel, nine figures are arranged in a row and housed under a shelter built recently for that purpose by the villagers of Shey. As in Sangkar, these figures were previously lying in the field. The figures are dark as local people put oil on them as an offering, and the lips and eyes are painted red. In the courtyard of the temple, on the base of central prayer flag (dar chen), stands a thin pillar carved with Maitreya on three sides (Figure 2.20). The detail of figures holding the throne of the central image is of particular interest since similar figures, holding the throne with their raised hands, are commonly seen in Kashmiri bronzes and in the mandalas in Alchi temples, particularly in the Vajraguhya mandala on the first upper storey of the Sumtek temple. Sakti According to the eminent Ladakhi historian Tashi Rabgias, Sakti was one of the centres through which Central Asian trade routes passed (Personal interview on 14 December 2007). It is possible that one of the routes from Kashmir to Leh

Embedded in Stone

Figure 2.20

59

Beautifully carved slim stele of Maitreya in the courtyard of lHa rdo rje chen mo chapel, Shey. Photo: Phuntsog Dorjay.

60

Dorjay

and from there on to Central Asia via the Nubra valley used to pass through the village. Another important route that connected Ladakh with Western Tibet went through Sakti to Tangtse via the Chang-la pass which leads to Rudok in Western Tibet. Sakti was one of the places where early Buddhist missionaries and artists put their energy into making rock carvings similar to the ones described above. A set of rock sculptures about one metre high is found near the ruins of an old castle in the valley, which leads to the Chang-la pass. According to the villagers, altogether nine castles or forts used to exist in the Sakti area. Three carvings are found near one of the ruined castles. The rock figures are badly weathered as they are situated on an open slope exposed directly to the sun and rain. The people from the nearby houses refer to them as Chamba (Byams pa), the Tibetan name for Maitreya. They are almost the same size and style as the series of Maitreya figures near the lHa rDo rje Chen mo Temple in Shey. These carvings lie in the ruins of what may once have been a specially constructed shelter. About a mile away I found another rock carving lying under a tree near a stream. The sculpture is locally known as Chenrezik Rangjung (sPyan ras gzigs Rang ’byung), literally meaning ‘self-born’ or ‘self-originated’ Avalokiteśvara’ (Figure 2.21). This sculpture is highly venerated and villagers said that many people from neighbouring villages visit it hoping that their personal wishes will be granted. The sculpture is housed in a wooden box and is only visible through a closed glass window. The figure is about half a metre high and seems to have four arms, but the box makes it difficult to see the details. Its style and iconography resemble those of the small figure of Avalokiteśvara near the lHa rDo rje Chen mo temple in Shey. Digar The left side of Sakti valley leads to the Wari-la pass (5200 metres) and, descending northwards from the pass, one reaches Digar village. That route may have been used by missionaries, pilgrims and artisans. They may have also used the Chang-la route but this pass is higher (about 5400 metres), and it would take longer to reach Digar. Passing through Wari-la down to Digar the route continues to the Shayok valley of Nubra. That route must have been easier for caravans, compared with the Khardung-la pass (5600 metres), which is very rocky and inhospitable for caravans with animals. The route over the Chang-la and down to the Shayok valley was also a popular trade route till recent times. Digar, a village of about seventy families, is situated at the junction of the Leh-Nubra route via Sakti and an alternative route via Sabu village (about 10 km east of Leh). The Indian Army, which is building a new road from Sakti

Embedded in Stone

Figure 2.21

61

Avalokiteśvara. The sculpture is locally known as sryan ras gzigs sRang ’byung at Sakti village. Photo: Phuntsog Dorjay.

to Nubra via the Wari-La, recently connected Digar with a motorable road. Near the village there are three impressive rock carvings about 1.5 metres high showing Buddhist divinities carved out of a single rock (Figure 2.22). The sculptures are identified as Maitreya, Avalokiteśvara and Vajrapani. They are similar to Indian style ones wearing dhoti-type lower garments. Their torsos, from the shoulders to the navel, show the same upward thrust as portrayed in the Kashmiri style sculptures of Alchi and Kashmiri bronzes.

Nubra Valley

Nubra lies to the north of Leh, between the Ladakh and Karakoram ranges. The valley has been an important halting place on the Ladakh–Central Asia trade route since ancient times. The rivers Shayok and Nubra meet in Diskit, the headquarters of Nubra valley, and then flow to the west, before joining the Indus in what is now Pakistan. One can still see reminders of the trade route from Leh to Khotan, for example the two-humped Bactrian camels that were formerly used as pack animals.

62

Figure 2.22

Dorjay

Padmapāṇi Avalokiteśvara at Digar village (Nubra region). Photo: Tashi Morup.

Embedded in Stone

Figure 2.23

63

Mañjuśrī (left) and Maitreya (right) at Tirith village (Nubra Valley). Photo: Stanzin Rabgyas.

More rock carvings were noticed in Nubra valley during my survey. No systematic study has so far been done on the rock carvings of Nubra valley, which will be part of my next survey as I discover more rock carvings on the route from Tirith to the Karakoram pass, which was used by traders from Yarkand and Khotan to travel to Ladakh until recent times (Bell 1890). In the east of Tirith village near the main road towards the Saser-la, I noticed some carvings on the big rock about five metres high, on whose face four Buddhist divinities were carved (Figure 2.23). From the ruin of a masonry wall it seems that those carvings were once housed inside an enclosure. Although the reliefs of the carving are not as deep as the ones in Kargil and the Indus Valley, the dress and features are almost the same, whereas the faces of the figures are rounder. All the four deities have three-pointed diadems on their heads. Starting from left to right, the first figure is Mañjuśrī, with a sword in his right hand and a book of wisdom in his left hand. The next is Maitreya, holding a rosary similar to that of the Mulbek Maitreya, the stupa on his head being clearly visible. The third and fourth figures stand displaying the varada mudrā gesture of boon-offering. Those deities have still to be identified as many details are missing. Another slim stele of early origin found in Sumur village (Figure 2.24), about 20 kms north of Tirith, shows an about one metrehigh figure identified as Maitreya in varada mudrā, standing on a lotus flower. A small miniature stupa can be seen in his crown.

64

Figure 2.24

Dorjay

A slim stele carved with Maitreya at Sumur village (Nubra Valley). Photo: Phuntsog Dorjay.

Embedded in Stone

65

One further possible alternative route, which might have been used by Buddhist missionaries and artists in the course of the time, is the path from Phyang to Hunder in the Nubra valley. Near the bridge by the Maitreya temple at Hunder, I noticed a rock with a carving of Maitreya very near the fast-running Hunder stream. The detail of the figure is worn out by water as that stream often overflows during the rainy season, but the headdress is still identifiable. Conclusion In conclusion we can say that the rock reliefs found around Ladakh are clearly not of Tibetan origin. If we compare the Buddhist rock carvings from Baltistan, Ladakh, Zangskar and Lahaul, in the Spiti area of modern Himachal Pradesh (Thakur 2001:42), it appears that they share a similar style. It seems that throughout the Western Himalayas there were similar missionary and artistic movements during the ninth to eleventh centuries, though we cannot deny the fact that Tibetan inscriptions and stūpa carvings made during the seventh to eighth centuries are also found in the region. But the rock-cut figures of Bodhisattvas are not just the creation of Tibetan soldiers. They were made by skilled craftsmen, who often had an aesthetic command of their vision and material. There is a balance in the relative proportions of the limbs and appendages, the garments and the ornaments, which is not just the consequence of the application of iconometric and iconographic rules. The harmonious and pleasing details on the front of the figures contrast with the extremely shallow carvings of later periods, which can be seen even on rock slabs placed over the maṇi walls. These Buddhist rock-reliefs represent the most important traces of direct Indian Buddhist influence in the entire western Himalayan belt, and we can attribute them to the period between the ninth century and Rin chen bzang po’s foundation of the first Buddhist monastic complex at Nyarma in the early eleventh century (Tucci 1988b:64). However some rock bas-reliefs, e.g. the Shey sMan la Maitreya, may be of a later period and perhaps were carved by local artists following previous examples. The present work demonstrates the inroads that Buddhist art made into Ladakh and from there on into Tibet and other parts of Central Asia. The people that walked along those roads were traders, missionaries, pilgrims and artisans. Some of those who carried the Buddhist faith along with them not only taught the dharma as they went on, but also commissioned religious artefacts. These varied rock carvings and inscriptions provide an important source for the study of the cultural history of Ladakh and neighbouring countries.

66

Dorjay

However, sadly they are succumbing both to natural weathering and human activities. As noted above, one poignant example is the loss of the Kharoshti inscription from the second century CE at Khalatse during the construction of two bridges (the last one in 2005). It is very important to conserve and protect these carvings before it is too late and all trace of them is obliterated. Hopefully, further studies will make it possible to achieve a more accurate dating of Buddhist sculptures in Ladakh with more precise documentation and detailed comparison with Kashmiri sculptures. Bearing that in mind, my study on the rock carvings of the region continues. Hundreds of rock carvings have been documented so far and the number increase every year as a result of further exploration. Recently we have found number of Buddhist rock carvings during a renovation of an old stupa in Skyu village (Figure 2.25). I also documented numerous stone carvings in Zanskar in the summer of 2008 and plan to make a detailed study of them.

Figure 2.25

Recently discovered rock carvings in an old stupa at Skyu village. Photo: Dorjay Angdu Kaya.

Embedded in Stone

67

Acknowledgements This paper is a shortened version of part of my Ph.D. dissertation, submitted to the University of Jammu (India) in 2005. I should like to express my gratitude to Monisha Ahmed and Christian Luczanits for their comments on my dissertation. I also should like to thank Erberto Lo Bue and John Bray for their comments on the present article. In the initial fieldwork the assistance of Tashi Morup was invaluable. I also thank Tashi Ldawa for sharing with me some important information on prehistoric rock carvings. I am also indebted to my family members in Ladakh and in Germany for their constant encouragement and support.

Historic Ruins in the Gya Valley, Eastern Ladakh, and a Consideration of their Relationship to the History of Ladakh and Maryul

With an Appendix on the War of Tsede (rTse lde) of Guge in 1083 CE by Philip Denwood Neil and Kath Howard

The purpose of this paper is to draw historians’ attention to a neglected part of Ladakh and to review what evidence we have to justify doing so. That evidence is inherently fragmentary, and the fieldwork data second-hand since neither of the authors has been able to visit Gya (rGya). Conclusions will be of the utmost tentativity; but nonetheless it is believed that the exercise will sufficiently demonstrate the importance of the history of the Gya district to inspire others to develop our work. First, we offer some comments on two castle ruins, at Rumtse (Rum rtse) and at Gya followed by an outline analysis of the sacred ruins which lie near them. Then we attempt to set both in their historical context and suggest some directions for future work.

The Castles (N. Howard)

Rumtse Rumtse stands in the upper part of the Gya valley. When A.H. Francke passed through in 1909 he wrote that “the most conspicuous building in the old town [Rumtse] was a round tower of great dimensions”, although he admitted that he did not visit the ruins; and he attributed the castle to previous non-Tibetan inhabitants called Mons (Francke 1914:63 and plate XXVII). As Figs 3.1 and 3.2 show, the round tower is really a high, smooth, palecoloured rock outcrop with masonry ruins on its top and other relics of decayed rooms, buildings and walls down the ridge, and toward the camera. A closer view, Figure 3.2, shows the outcrop and the way in which many of the buildings were built out to its edges or even up from its face. The masonry consists of random stones picked off the hillside and mortared together with mud which contains a great deal of sand and gravel. This combination is very susceptible to attack by frost and rain and is therefore weak; buildings constructed like this

© koninklijke brill nv, leiden, ���4 | doi ��.��63/9789004271807_�05

Historic Ruins in the Gya Valley

69

Figure 3.1 Rumtse castle from the west: the light-coloured rock out-crop summit is against the sky with settlement ruins and rubble on the slopes below and toward the camera. Photo: K. MacDonald.

decay fast without constant maintenance. The rubble strewn down the ­hillside in Figure 3.1 suggests something of the size of the old settlement before it shifted to the valley floor and natural decay destroyed its abandoned buildings. Although there is no evidence to provide a date of origin for these buildings, there is nothing incompatible with their having been begun nine or ten centuries ago when, as we shall see, there is reason to believe the associated religious structures were built. Their builders were ancestors of the present inhabitants, not ‘Mons’, as Francke believed, to whom apparently anything hereabouts is attributed when it is so old that its origins have been forgotten.1 Gya The ruins at Gya, a few miles lower down the valley, are of an altogether grander place although its site and defensive concepts are broadly similar to Rumtse. Figure 3.3 shows the ruins from the valley floor and Figure 3.4 is a sketch made 1 See N. Howard (1999:224–27).

70

Howard

Figure 3.2 Rumtse castle: ruined masonry and rubble on and below the summit outcrop. Photo: K. MacDonald).

from photographs showing the whole settlement, from the stupas along the approach path near the river to the highest defensive tower. The tower on the sky-line in Figure 3.3 is tower 3 of Figure 3.4. The ruins occupy a spur of the mountains on the north side of the Gya valley where a tributary torrent joins the main river. The round towers on the higher ridges provide defence against an attack from above. The strong point (labelled ‘castle’ on Figs 3.4 and 3.5, originally a large rectangular building or buildings) and the smaller residential buildings below it form the core of the settlement. It is clear from those buildings which survive that they were built close against each other on the edges of the steep slopes so that their outer walls provided a defensive barrier without building separate defensive walls. This arrangement

Historic Ruins in the Gya Valley

71

Figure 3.3 Gya fortress from the west; settlement ruins, temples and some of the many stupas. Photo: K. Macdonald.

is shown diagrammatically in Figure 3.5. Large quantities of rubble indicate the area originally occupied by buildings, for example down the right-hand slope as far as the double stupa in Figure 3.3. The ruins at Rumtse reveal the same building plan, but it does not show clearly in the photographs; equally the building materials and techniques already illustrated for Rumtse are the same as those at Gya. Further defences here must somehow have protected the lower part of the settlement, facing the river, but we have insufficient detail for a reconstruction except to note that they included at least two more round towers: tower 1 (which has a lha tho on it today) and another which may have guarded, or been part of, the gateway or entrance. According to the defensive systems recognized in N. Howard’s earlier paper on the fortresses of Ladakh, the defensive principles employed at Gya, and less extensively as far as we can tell at Rumtse, are similar to those which have survived from early times in the history of the kingdom of Ladakh. If we consider the settlements and defences of Temisgam (gTing mo sgang) before the mud walls and towers were built, or of Wanla before the building of the mud towers (Howard, N. 1989:251–54, 257–61), we have parallels with Gya: a defensible settlement of closely packed dwellings (Figure 3.5) with outer defences of

72

Howard

Figure 3.4 Gya: axonometric sketch of the whole fortress, view from approximately south-west (G. Kozicz, not surveyed).

Historic Ruins in the Gya Valley

73

Figure 3.5 Gya: axonometric sketch of the core or central area (G. Kozicz, not surveyed).

round towers (Figure 3.4); no mural defences, but protection in depth against a flank attack by means of detached towers. Such fortified settlements built on steep ridges, but without towers, used to be common all over greater Ladakh. Surviving examples can be seen at Hankar in the Markha valley (Howard, N. 1989:261–64), Tshazar and Zangla in Zangskar (Howard, N. 1995a:88,97), at Henasku and Chigtan in Purig, and probably in the oldest parts of Lamayuru village (Herdick 1997:203–34). They are not confined to Ladakh and can be seen, for example, at Jarkot and Kagbeni in the northern, the Tibetanized, part of Nepal (Howard, N. 1984 cover photograph; 1995b:26). It is the simplest form of defence for a settlement and, not surprisingly, was built wherever and whenever perilous conditions made small communities erect the best protection they could. If settlements grew larger a walled enceinte may have been added to the original defensible settlement, as can clearly be seen at Leh, Shey , Basgo and Temisgam, but obviously not at Rumtse and Gya. In the core area of Gya are religious buildings. The tall white one with a wall painted dark red (Figure 3.3) is the older of the two temples. In his pioneering survey Kozicz (2006:57–60) has identified two of the rooms inside as a gonkang (mgon khang) and a dukhang (’du khang). The dark red paint is on the left side is of the gonkang. These two rooms contain elaborate mural paintings; the others are bare and empty. Amongst the paintings on the walls of the Dukang, Kozicz has reported portraits of leading teachers of the Kargyu (bKa’ brgyud) school. He concluded that the schemata of the mural paintings would be appropriate to a date in the early 17th century; and this is compatible with what little we know of the history of Gya at that time, as we shall see. The

74

Howard

undistinguished architectural style of the temple building would be, at least, not inappropriate to such a building date. Kozicz’s fieldwork has revealed that the temple stands upon the ruins of an earlier building and that its own walls are in two parts: random-textured mud-mortared stone in the lower parts and sun-dried brick in the upper.2 All three can be seen in the facing elevations in Figure 3.5. The other complete buildings, left and above the old temple in Figs 3.4 and 3.5, are a temple, a dwelling and a latrine, all newly-built structures.3 The temples, old and new, are the only maintained buildings on the site; all the rest have been abandoned, perhaps for two centuries or more. According to an inscription discovered by Francke (1914:63) on a nearby maṇi wall “of the time of bDe-ldan rnam rgyal” (ruled 1642–c. 1691) Gya fortress was named rGya mkhar rmug po, The Dark Red Castle of Gya; whether the colour refers to some natural colouring in the rock or to paint on lost buildings we cannot tell. There has been a lavish provision of stupas, both on the approaches and in other parts of the settlement. This is a regular sign of a royal residence and the whole assemblage seems quite appropriate for the old capital of the Gya chieftainship or kingdom. However, with the exception of the dukhang and the 17th century inscription containing the fortress’s name, we do not yet have any evidence by which to date the settlement. Given the right circumstances, as I have suggested elsewhere, building techniques and the shape of certain defensive buildings can provide evidence of building date (Howard, N. 1989:passim). Here the masonry style is so commonplace that all the buildings could date from any period. There are two partial exceptions to this rule: had they been built under Ladakhi royal influence between c. 1550 and the late 17th century they would have included square towers, shuttered mud (tamped earth) and/or banded textured masonry, all of which are absent here. However from the little history we do know, it is more likely that the ruling house of Gya was in opposition to the kings of Ladakh rather than in co-operation during that period (see below). Both these fortified settlements, and the ruined fort in a side valley opposite to, and mid-way between, Rumtse and Gya (noticed by Francke) need thorough exploration and documentation, and the original function of every surviving building should be recorded. Meru is completely unknown to us and there may be others, too. They should all be visited as a matter of urgency; and Kozicz’s work shows what rewards may be in store.

2 Kozicz, personal communication 21 April 2009. 3 Kozicz, ibid.

Historic Ruins in the Gya Valley



75

The Sacred Ruins (K. Howard)

Rumtse The sketch plan, Figure 3.6, has been drawn from a photograph taken through a long lens from the summit of Rumtse castle; it is of the stupa (mchod rten, chorten) field shown in Francke’s Plate XXVIb (Francke 1914). We have here a very interesting complex of individual stupas, whose variety of designs include some which were apparently used only at the time of the Second Diffusion, and stupa walls (or rows; but the building method seems to be a wall with stupas along its top). mChod rten walls are particularly common in the eastern parts of Ladakh although fine examples have survived in other places. Here there are three stupa walls, almost certainly of 108 each, but too decayed to be counted. mChod rten walls like these, e.g. in Figure 3.7, were considered by Tucci to be associated with the most ancient centres of lamaistic Buddhism in western Tibet (Tucci 1988:50). In Ladakh circumstantial and locational evidence suggests they were built no later than the early 15th century. There are many between Shey and Ranbirpur and more at Sabu (Sa phud), also on the Alchi (A lchi) side of the Indus opposite Saspol (Sa spo la), and at Khalatse (Kha la tse). At Sabu, we are told, lHa chen Khri gtsug lde built two rows of 108

Figure 3.6 Rumtse stupa field: diagrammatic sketch plan drawn from photographs (K. Howard).

76

Howard

Figure 3.7 Rumtse stupa field: stupa wall. Photo: J. Bray.

mchod rten (Francke 1926:99). Lo Bue (2007:184) has recently implied that King Khri gtsug lde still ruled in the 1440s which may place his mchod rten walls late in their history and shortly before the introduction of Gelugpa beliefs into Ladakh/Mar yul. It remains to be seen whether this last is a significant observation or merely a coincidence of chronology. Some of the mchod rten here will now be briefly described. Figure 3.8 shows a particularly fine lha bab (or “Divine Descent”) mchod rten. Its base is circular (rather than square as is more usual) and there is a lotus decoration on both the base and the dome. Between base and dome there are four square tiers with facetted corners, and traces of ladders or stairs up the four main faces. Francke (1914:Plate XXX1a) published an excellent photograph of this mchod rten in which the detail is less weathered but his caption incorrectly says it is at Gya. He also attributes its construction to the Mon people, but this is without foundation. Another stupa form characteristic of the Second Diffusion is seen in Figure 3.9. In an earlier paper K. Howard (1995:62) gave this type the provisional classification of ‘lotsava type’. Characteristically, it has a tall plinth, either square or rectangular, and a somewhat disproportionately small mchod rten, or more than one, on its upper surface. The mchod rten on this example has four tiers and a dome, with niches in the middle of the sides of the second and uppermost tier. At the far side of the complex there is another lotsava type mchod rten, the stupa of which has a circular base and several facetted tiers below the dome,

Historic Ruins in the Gya Valley

Figure 3.8 Rumtse stupa field: lha bab mchod rten (Photo: J. Bray); this is the stupa photographed by Francke (1914: Plate XXX1a).

Figure 3.9 Rumtse stupa field: ‘lotsava’ type stupa; two more and stupa wall behind. Photo: J. Bray.

77

78

Howard

Figure 3.10

Rumtse stupa field: ‘lotsava’ type with round stupa; partial collapse shows cobble construction of the base. Photo: J. Bray.

Figure 3.11

Rumtse stupa field: stupas of double rnam rgyal type. Photo: J. Bray.

Historic Ruins in the Gya Valley

79

Figure 3.10. The extensive weathering shows the mud-mortared cobble-stone construction of the base. Further away there are examples of double stupas on a single base. These are of the rnam rgyal type, which is not usually doubled today, but they appear to be as old as their neighbours, Figure 3.11. A detailed survey here would yield materials for the classification of stupas typical of the Second Diffusion since, to judge by the photographs, all are of much the same age, irrespective of their shape. Gya The mchod rten complex at Gya is larger than that at Rumtse. From Kozicz’s survey and sketch plan, Figure 3.12, we find there is one wall of 108 mchod rten and the ruined remains of a shorter row, perhaps of 8 stupas. There are more examples of mchod rten of the types described at Rumtse. There are fine examples of lha bab mchod rten with detailed design differences from those at Rumtse, e.g. Figure 3.13, which has no projecting cornice above the base and a smaller dome. Some are clearly built of mud brick, a construction feature apparently not used at Rumtse (see Figure 3.14). There are also remains of a mud-brick building, perhaps a temple; its ruins are marked on Figure 3.12. This may be the mud-brick ruin which Francke (1914:64) suggested was a monastery; Kozicz considers it to have been a temple. The bricks used in this building appear to be very similar in size and bond pattern to those at Nyarma. Finally, there is a new stupa, the Sankasyastupa shown

Figure 3.12

Gya stupa field: sketch plan, not to scale (G. Kozicz).

80

Figure 3.13

Howard

Gya stupa field: lha bab mchod rten; cobble base. Photo: K. MacDonald.

on the map. It is of the rnam rgyal type, as large as any of the ancient ones, painted white, and stands under a masonry canopy painted with red and white stripes. It is quite unlike any of the ancient structures. Our overall conclusion is that all the old ruined mchod rten, and other structures in the Gya stupa field, date from the early period of the Second Diffusion of Buddhism, the time when the temple and monastery complex at Nyarma was being built, or only a little later. The brickwork in the stupas at Gya suggests a very close date. The relics at Rumtse may have been built a little later still, if we speculate that the Rumtse field was a sort of replica of that at Gya. One name missing from our sources for the study of the history of Gya is that of Rin chen bzang po, the Great Translator and founder of Nyarma. His

Historic Ruins in the Gya Valley

Figure 3.14

Gya stupa field: brick-built stupa. Photo: K. MacDonald.

81

82

Howard

biographies make no mention of Gya; although, if he was personally present at the building or consecration of Nyarma monastery as his biography states that he was, he must have passed through Gya. For this reason we suggest that the stupa fields of Gya and then Rumtse followed a few decades after the building at Nyarma. We would also add that religious complexes like this are not built by a few lonely monks in a thebaid; they are the product of rich—and therefore royal—patronage and powerful religious inspiration. This in turn suggests that the castle ruins were at the centre of a significant political entity between nine hundred and one thousand years ago.

The Historical Context (N. Howard)

It might be thought that the first reference we have to Gya in Ladakh is in La dvags rgyal rabs (henceforward LDGR) in the early section on the emperors of Tibet. There we learn that King Sad na legs (early ninth century) built a temple named sKar chung rdo dbyings in rGya sde, i.e. in the Province of Gya, which Francke (1926:89) suggests was in Western Tibet. Philip Denwood points out that “this temple is the well-known temple and stone pillar (“rdo rings”) at “Skar cung”, southwest of Lhasa published by Hugh Richardson (1985:72–81). The epithet rGya sde found in some sources does not occur in the early period and one modern source spells it rGyal sde which is perhaps more likely, i.e. royal community (of monks)”.4 However, from other evidence we do know that Ladakh was then part of the Tibetan Empire.5 Gya and the Establishment of the Kingdom of Ladakh-Maryul in the Mid-Tenth Century After the collapse of the Tibetan empire, a descendant of the last imperial king migrated to the west, married into the aristocracy of Purang (Pu hrang) and established himself as the ruler of mNga’ ris skor gsum. His name was sKyid lde Nyi ma mgon; and some of his sovereignty may have been nominal rather than actual because LDGR tells us that he left Maryul (Mar yul) undisturbed. At that time “Upper Ladakh of Maryul” was held by rulers who claimed to be descendants of Gesar, whilst Lower Ladakh was split up into small independent principalities (Francke 1926:93). We might tentatively deduce an insight into the political geography of those times: that Upper and Lower Ladakh were both considered to be parts of Maryul but that Maryul was greater than the

4 Denwood, personal communication by e-mail 14 October 2008. 5 E.g. Denwood (1980:155–163), Denwood and Howard (1990:81–85).

Historic Ruins in the Gya Valley

83

two parts of Ladakh. We shall return to this matter. It is especially difficult to understand what is meant by “Upper Ladakh of Maryul”. The same passage in the LDGR informs us that Nyi ma mgon built a castle named rTse tho rGya ri: it has been suggested that this might have been on the hill above Gya and built in the year 935 CE (Petech 1997a:232). If the suggestion is correct Gya must have been outside Maryul which sKyid lde Nyi ma mgon “left undisturbed”. However, Gya is such a vague toponym that the reference may have nothing to do with Gya in eastern Ladakh. The photographic evidence used in this paper is inconclusive as far as building date goes, as we have already noted. However far Nyi ma mgon may have personally extended his power into Maryul or Ladakh, the LDGR makes it plain that the kingdom known to history by these names was put together by the eldest of his three sons, IHa chen dPal gyi mgon, who inherited: Mar yul of mNga’ ris, the inhabitants using black bows; Ru thogs [Rudok] of the east and the gold mine of ‘Gog [in the same area]; nearer this way6 lDe mchog dkar po [Demchok]; at the frontier Ra ba dmar po [?]; Wam le to the top of the pass of the Yi mig rock [Hanle and the nearby Imis Pass]; to the west to the foot of the Kashmir Pass [Zoji La], from the cavernous stone upwards hither; to the north to the gold mine of ‘Gog; all the places belonging to Gya [translation and identifications from Francke (1926:94)]. Although other Tibetan sources differ over the territories inherited by the younger sons there is more or less unanimity about what IHa chen dPal gyi mgon gained. Only in a rather late source7 quoted by Petech (1997a:232) and Vitali (2003:55) is it suggested that Rudok was inherited by the second son, along with Guge (Gu ge). We have no means of deciding between these statements, except to observe that since Rudok is on the far side of Demchok from Guge the geographical integrity of the inheritance is weak. It is, of course, perfectly possible that the original inheritances changed hands relatively soon after they were made. Accepting for present purposes the territorial division given in LDGR, there are still problems of interpretation. At first sight it might be read as a list of 6 The expression ‘nearer this way’ in Francke’s translation might imply that the source manuscript from which this passage of LDGR was transcribed or compiled was originally written in Guge? 7 sTag tshang pa Śribhutibhadra, 1434, Gya bod yig tshang mkhas pa’i dga’ ’byed, Thimpu 1979.

84

Howard

separate districts given to King dPal gyi mgon: Maryul plus those which follow. On the other hand, with reference to the map, those places which are named after Maryul may be included in order to define the frontiers of Maryul territory in terms of the main routes into its neighbours, i.e. they may be regarded as places on the borders of his inheritance. Before we try to decide the truth, let us briefly consider “a new translation” of this passage in LDGR by Zahiruddin Ahmad, recommended by Petech but apparently not noticed by any of his readers (Petech 1977:17 n.4). Ahmad’s translation reads: To the eldest, dPal-gyi-mgon, (he gave) Mar-yul (Ladakh) of mNga’ ris, the inhabitants using black bows. (The frontiers of Mar yul (Ladakh) were:) In the east, Ru-thogs and lDe-mchog dKar-po which resembles (sic) the gold-mine (called) ‘Gog; Ra-ba dMar-po at the frontier (itself); (and) up to the top of the rocky pass of Yi mig (I mis) (in the territory of) Wam-le (Han-le). In the west, the foot of the Kashmir Pass (Zo Ji La), from the cavernous rock upward ( yan chad) (towards Ladakh). In the north, the lands up to (tshun chad) the gold mine (called) ’Gog (from the north side?) belong to China (Ahmad 1968:340–41, including his bracketed interpolations). So Ahmad supports the proposition that the places named after Maryul define the borders of Maryul and justifies it by changing the punctuation used by Francke. Furthermore, on the basis of some earlier research published elsewhere, he identified Gya and its “places” with China and placed ’Gog in the Yarkand river valley. Finally, he apparently did not consider the significance of the well-known gold fields east of Rudok or of the other sources of gold along the Indus river towards Baltistan. On balance, this translation is not helpful. Before drawing our discussion to a close with the aid of Figure 3.15, let us note that so far no mention has been made yet of the famous, earlier, goldproducing country, Suvarnagotra. Philip Denwood (2008:9) has recently concluded that Suvarnagotra lay north of Zhang zhung and east of modern Ladakh: “in the area of Rudok in modern Western Tibet, with likely extensions in the direction of Nubra and the Shyok valley to the west, the Lingzi Thang (Aksai Chin) to the north, Gartok and the headwaters of the Indus to the south, and the gold fields of Thok Jalung to the east”. Suvarnagotra no longer existed as an independent kingdom, but its resources were still there for the taking. Ra ba dmar po is not known. Hanle is west of Demchok and Francke says that the Imis Pass is at the headwaters of the Hanle River, which must mean it is a pass into Chumurti; so the border with Guge is defined here. All sources assign Zangskar and Spiti-Spichog to other brothers, which define by exclusion the limits of

Historic Ruins in the Gya Valley

Figure 3.15

85

Map of western Tibet illustrating the inheritance of dPal gyi mgon; Maryul is not named because its boundaries cannot be precisely defined.

dPal gyi mgon’s inheritance to the south-west. The Zoji La is the pass at the far western end of Purig; and the mention of the second gold mine of ’Gog, which is said to lie to the north, may be both a textual corruption of its name and a reference to some gold-yielding district of the Indus river or in western Nubra. The Yarkand river, Ahmad’s preference, is too far away, as is Wakhan, which in some Tibetan sources is also called ’Gog. Finally “all the places belonging to Gya” fit into the scheme of definition as the lands between the Hanle river basin, Spiti, Zangskar, and the junction of the Gya river with the Indus. We cannot be certain, but perhaps these included the grazing grounds of Rupshu, east of the Taglung La as far as Tsho Moriri. Within and including this circuit of boundary markers and districts lay the kingdom inherited by dPal gyi mgon. Bearing in mind the evidence for a more congenial climate a thousand years ago, with both agriculture and p ­ astoralism on the high plains where they are not possible today (Bellezza 2000:1–61, Denwood 2005), this was a rich inheritance indeed for the eldest son. The

86

Howard

words of LDGR can be interpreted as both a list of separate territories inherited by dPal gyi mgon and as a definition of his kingdom’s boundaries. Notably missing, however, is mention of any place west of the junction of the Dras River with the Indus, vis. present-day Baltistan. The reason is that the country lying between that point and Gilgit (and the northern areas of Pakistan today) was then the separate kingdom of Bolorian Tibet (Minorsky 1982:93). Finally, in this context it is worth mentioning ‘Gya mur orr’. An intriguing passage in the Kulu vaṃśāvalī states that during the reigns of its 47th to 50th kings, Kulu was involved in wars with Baltistan, Tibet and a place named Gya mur orr (Hutchison and Vogel 1933:438). From the context this last must be in or near Ladakh. Although several details of the story are fantastic, the curious precision in naming this unidentified place suggests some reality lies in the background. The events are undateable from the source but the existence of Gya and Meru at the gates of Ladakh and on the great trade route between India and Central Asia via Kulu and Rupshu may be more than a coincidence. We have no field data from Meru, but Francke (1914:65) states that it was “famous in olden days” and used to be home to one of the important monasteries of Ladakh. On the other hand, Gya mur orr might be a corruption of rGya Mar yul. The Problem of the Location and Extent of Maryul and Its Relationship to Ladakh and Gya We can already see that the issue of the location of Maryul may be central to an understanding of the history of Gya. Whether “all the places belonging to Gya” should be included in Maryul is probably too fine a question to debate, given the present evidence; however these words might seem to imply that Gya was the centre of a small, previously-independent kingdom, or a substantial chieftainship at least. A review of the documentary evidence for the name Maryul throws up some interesting historical propositions. The earliest attestation to the existence of a prototypical name which became Maryul may have come down to us is in the memoirs of the Chinese Buddhist pilgrim Xuan Zang, who wrote in the early seventh century. He mentions a country named Mo-lo-so which, in its geographical position, must have been approximately where we find Ladakh today. But Mo-lo-so cannot be related etymologically to the name Ladakh. However, it has been demonstrated that an original toponym *Mars/*Mras could have become, through a Sanskritized intermediary, Xuan Zang’s Mo-lo-so; and through a Tibetan borrowing, in the form of Mard, become Mar yul (Uray 1990:222). Chronologically, the next attestation may be in the Persian geographical treatise Hudud al-Alam (The Regions of the World) written in 982 CE but incor

Historic Ruins in the Gya Valley

87

porating both contemporary and earlier material. Here we read of “N.ZVAN, a wealthy country of Tibet”, in which there is a tribe named “Mayul from which the kings of Tibet come” and a “village named Muyul”. N.ZVAN is not readily identifiable and there are several other difficulties; but from the context Huhud’s Tibet includes the countries of Zhang zhung, Ladakh and Baltistan (Minorsky 1982:93).8 Therefore Mayul and Muyul might be taken to be references to Maryul. Philip Denwood has pointed out that “Mayul from which the kings of Tibet come” is reminiscent of the Chinese Mo yu, “far to the west”, from which a descendant of the kings of Tibet was invited to the throne of the kingdom of bTsong kha, east of the Kokonor, in 1008. Mo yu was identified by R.A. Stein with Mar yul, which tends to support the present author’s argument (Stein 1959:231),9 though others have doubted the identification. Moving forward in time, we should consider the wording of two inscriptions in the temples at Alchi written less than 150 years after the lifetime of sKyid lde Nyi ma mgon. Inscription 3 in the Dukang mentions “The religious house of Nyarma in Maryul” (Denwood 1980:146) not in Upper Ladakh as we would say today. Nyarma is near Thikse and Shey. Unlike LDGR, which achieved the form we have today long after the founding of the kingdom which it describes, our sources so far are have been recording the place-name usage of the time at which they were written—and there has been no mention of Ladakh. The earliest occurrence of that name known to this author is in Inscription 7 at Alchi which describes the gSum brtsegs Temple as being “here at Alchi of Ladakh, in Lower Maryul of Upper mNga ri” (Denwood 1980:145, 148).10 The inscriptions are written in verse and the constraints of verse form may have condensed the geographical reference. But in the biography of Rin chen bzang po, edited and translated by Snellgrove and Skorupski (1990:91) in the same volume as the inscriptions, we find his monastery of Nyarma stated simply to be in Maryul. These three sources do seem to exclude any part of Ladakh from eastern Maryul. Several centuries later, Deb ther dmar po gsar ma includes the following passage: “From them the kings of Shel [Shey] in Mang yul [Maryul] and Nub ra are descended and also those of Glo pa, La dags [Ladakh] and Zangs dkar. These (last) five rulers honoured, in their heart, as their chaplains, only the dGe lugs pa; in the monasteries subject to them no other system was followed” 8 9 10

I am grateful to Marjo Alafouzo for directing me to Hudud al-Alam and Minorsky. Philip Denwood, personal communication, 14 October 2008. Despite recent attempts to re-date the gSum brtsegs temple two centuries later, I follow the translator’s dating of these inscriptions: middle to second half of the 11th century. See Denwood, this volume.

88

Howard

(Tucci 1971:169). The author, bSod nams grags pa, lived from 1478 to 1554 and by the time he died a revival of the ’Bri gung pa was beginning in Ladakh; he was ­obviously referring to conditions in the second half of the 15th century. As before, Shey is in Maryul, and Ladakh is a different country. Mirza Haidar Dughlat, an acute observer, described these countries as he found them in the years 1533 to 1535. In his notes on their settlements he includes the following: “Balti, which is a province of Tibet; Balti in turn comprises several [smaller] districts such as Purik, Khapula, Ashigar, Askardu, Runk, and Ladaks, and each of these contains fortresses and villages” (Elias and Ross 1895:410). At that date Tibet was understood by Central Asians to be what we know as Ladakh at its maximum historical extent; Ashigar is Shigar; Askardu is Skardu; Runk is unexplained, but might it be Rong mdo / Rongdu / Rondo and other spellings on the maps? It is situated down the Indus river to the west of Khapalu, Shigar and Skardu, all mentioned by Mirza Haidar. The topographical context seems persuasive.11 For Mirza Haidar, “Ladaks” is a small part of the total. A few lines further on he lists the districts he conquered: “Balti, Zangskar, Maryul, Guga [Gu ge], etc”. Ladakh is not listed—although he certainly did conquer that country—presumably because he saw it as part of Balti. So Ladakh is consistently different from Maryul and must have been west of Maryul. In the rest of his memoir he frequently refers to Maryul and often to Balti but never again to Ladakh. When his Khan decided to return home because of ill health, leaving Mirza Haidar to destroy “the idol temple of Ursang (i.e. Lhasa)”, he “set out from Maryul in Tibet, for Yarkand” (Elias and Ross 1895:443–445).12 He crossed the “pass of Sakri”, which must be that above Sakti (not the Kardung pass as Elias and Ross suggest), descended to Nubra and died at a camping place named Daulat Beg Uldi which is two-and-a-half hours below the Karakoram Pass (De Filippi 1932:413–14). Maryul here must refer to Upper Ladakh as we call it today, most probably the fertile oasis of Shey. This identification seems to be confirmed when Mirza Haidar reached Ladakh again after the failure of his raid: “The Chui [ Jo plural, i.e. rulers] of Maryul, named Tashikun and Lata Jughdan, . . . . . . gave us the castle of Sheya [Shey] which is the capital of Maryul [to live in during the winter]” (Elias and Ross 1895:460). Less than a century later, towards the end of 1631, the Jesuit missionary Francisco de Azevedo passed through Gya on his journey from the Mission at Tsaparang to King Seng ge rnam rgyal of Ladakh. He records that he was warmly welcomed by the Gya chief “whom the King of Ladakh had deprived 11 12

Denwood and Howard in conversation. “Ursang” means, of course, the Tibetan provinces of dBus and gTsang.

Historic Ruins in the Gya Valley

89

of the kingdom of Mariul [Maryul]” (Didier 1996:89–90; Wessels 1992:107, 304).13 Both the translations used here, and the original Portuguese, continue with the information that two days later Azevedo and his companions reached a night stopping place within Ladakh. The implication seems clear that the first night’s stop (at an unknown place) was still in the former territory of Maryul. It was approximately 45 miles from Gya to Leh by the old paths as reported by William Moorcroft, two centuries later still. He made relatively short daily stages and took five days over the journey, recording each day’s mileage (Moorcroft and Trebeck 1841:140–46). Azevedo, in a hurry before winter set in, took three days. Two stages, 30 miles, would have brought him to a point between Martselang and Shey, e.g. Stakna or Chushot, but his description suits the country around Martselang better. Wessels suggests Upshi, but that marks only one third of the journey. We may conclude that, before the chief of Gya was deprived of Maryul, his border with Ladakh lay somewhere between Upshi and Martselang and that Maryul was a more easterly territory than formerly, excluding Shey. Hesitantly, we may propose from the foregoing that for over a thousand years the part of the map which we know today as Ladakh was formerly known as Maryul or some prototype of that name. In the 10th and 11th centuries, and possibly as late as the 16th century, the western parts of Maryul were also known as Ladakh. Recent research has drawn attention to a dynasty or dynasties of kings of Maryul who ruled at Shey apparently in parallel with the Ladakh dynasties we know from LDGR.14 We know almost nothing about the takeover of Shey/Maryul by the Ladakhi line of kings except for the words of LDGR: “This king [lHa chen Bha gan] being very fond of fighting, he and the people of Śel [Shey], having formed an alliance, deposed and subjected the sons of the King of Sle [Leh] . . . etc. . .’ ” (Francke 1926:102). The statement is fraught with difficulties and almost totally opaque; but Bha gan’s son was King bKra shis rnam rgyal who is said by the LDGR to have “conquered [all the 13 14

I am grateful to John Bray for pointing me towards Didier’s work. For more than 100 years the kings listed in LDGR have been presumed to have been kings of Maryul as well as of Ladakh. In fact many people have assumed that both names referred to the same country. As long ago as 1977 Petech (1977:14–27) pointed out in some of the inadequacies of LDGR, including the fact that many kings’ names must be missing and some of the names listed are obviously not Tibetan (pp. 14–37). Coincidentally in 1976 Joseph Gergan published Bla dvags rgyal rabs chi med gter which contains a genealogy of the kings of Shay-Maryul, preserved in gDung rabs zam ‘phreng, almost none of whom occur in LDGR. It has attracted little attention and no serious study. More recently a second genealogy of the kings of Shay-Maryul was discovered and published by Vitali; see Vitali (1996:495 n. 833 and 1996:57–49) for both genealogies. There are several discrepancies between the two.

90

Howard

country] from Purig upwards and from Gro-śod downwards hither” (Francke 1926:103). Gro-śod is the valley of the upper gTsang po east of the Mar yum La on the border between Guge-Purang and southwest Tibet. bKra shis rnam rgyal established his capital at Leh where he built a palace-fort and his conquests must have included all Maryul and the Gya valley. His successor Tshe dbang rnam rgyal is reported in the same source to have reconquered (for reasons not given) the same territory (Francke 1926:105). Gya and Maryul would have been gathered in again; but perhaps the Gya chiefs retained some measure of independence still. Then, at the beginning of the next reign “all the vassal princes in one place after another lifted up their heads” against their king ’Jam dbyangs rnam rgyal (Francke 1926:106). We may guess that the independence of Gya and Maryul was restored. The next king, Seng ge rnam rgyal, slowly rebuilt the kingdom to its former extent, concluding with the conquest of Guge and the overthrow of its royal line in 1630. His predecessors had not removed the conquered rulers of Guge and may have been equally prepared to have the Gya chiefs as their vassals. But, as in the case of Guge, Seng ge rnam rgyal appears to have stripped the Gya chiefs completely of any remaining vestiges of their independence; hence, the situation Azevedo records. In this context it is worth recalling that when Francke visited Meru (lower down the valley than Gya) he reported that Seng ge rnam rgyal had made Meru monastery the “mother” of his new monastery at Himis and had transferred the “spirit” of Meru monastery and “most of the images” to Himis (Francke 1914:65). Was this in order to deprive the Gya chiefs for ever of the spiritual power necessary to support their independence? Thus it may be attractive to speculate, within this historical scheme, that the Gya chiefs were the heirs of the Maryul kings, whose territory had been gradually reduced by the expansion of the rNam rgyal kings of Ladakh. But in the light of presently available sources that must remain a speculation. Gya, Gesar and the Dards Nobody has yet explained who were “the descendants of Gesar”, the great mythical hero of the Tibetan world, whom Nyi-ma-mgon found ruling in “Upper Ladakh of Maryul”. We have been told that an oral history of the chiefs of Gya is still known, in which it is recounted that they were descended from Gesar.15 We await its publication with great interest. In the mNga’ ris rgyal rabs we read of a war in 1083 in which King rTse lde of Guge fought a campaign against a chief named “Gye sar”. The editor of the 15

Bettina Zeisler, personal communication, Rome, September 2007; Nawang Jinpa of Hemis, correspondence late 2012.

Historic Ruins in the Gya Valley

91

rGyal rabs, Roberto Vitali, locates this war in eastern Ladakh and concludes that Gesar and his army “were of Iranic stock from the Indo-Iranic borderlands, most likely Dards, and thus belonged to the pre-Tibetan local ethnic substratum” (Vitali 1996:123–24, 323–327). A review of the Tibetan text by Philip Denwood, which forms our appendix, suggests that this war or raid had nothing at all to do with western Tibet or Gya of eastern Ladakh. Furthermore, it cannot be repeated too often that Dards never inhabited or ruled Ladakh. The name was invented by a genius and charlatan named G.W. Leitner in the late 19th century in order to support certain theories he had about the tribes living between Kabul and Chitral, including Chilas and Gilgit, and based it on ethonyms found in the classical Sanskrit texts and in Mediterranean writers who had never been within a thousand of miles of the upper Indus River.16 Francke took up the name and applied it in Ladakh on the basis of several misunderstandings, not least that the place named Rong chu rgyud in the Dardic hymnal was Rong chu rgyud, east of Martselang, whereas Gergan and Hassnain in their introduction to a reprint of Francke’s A History of Western Tibet (1977:9–10) point out that what is really intended is the Rong chu rgyud in Skardu tehsil. Somehow the name stuck to the people of Dras and of Da-Hanu. None of Francke’s archaeological evidence for the Dards stands up to reanalysis (N. Howard:1999:224–227). It is over thirty years since Graham Clarke’s masterly survey of the ethnographic literature demolished Francke’s Dard theories on other grounds but few workers on Ladakh seem to have noticed (Clarke 1977). Later Importance of the Descendants of the Gya Chiefs The chief, or maybe still king, of Gya at the beginning of the 17th century was an early supporter of the ’Brug pa and a sufficiently substantial personage to give accommodation to sTag tshang ras pa on his first visit to Ladakh in 1616, and was his host again in 1622 (Petech 1977:38, 40). The building or repainting of the older temple as we see it today may have been done under his inspiration. Shortly afterwards sTag tshang ras pa and his religious order became the favourites of King Seng ge rnam rgyal while, as we have seen, the Gya chief perhaps simultaneously lost the last of his independence. Despite this serious set-back, and a prophecy by King Tshe dbang rnam rgyal of Ladakh (c. 1575–95) that no member of the Gya chief’s family should be given high office, or “they will put a bridle on the king of La dvags [Ladakh],

16

For a brief account of Leitner’s life and works see chapter 7 of Keay (1982).

92

Howard

and on the country, and behave [like] a rider” (Francke 1926:226),17 the ruling family of Gya did become very important in the kingdom of Ladakh at the end of the 17th century and into the 18th. In the 1690s the chief of Gya was ‘Brug grags; his younger brother was named bSod nams lhun grub. The latter began his career in a monastery, then abandoned religious life and became first a no no, then a bka’ blon, and finally chief minister to King Nyi ma rnam rgyal, twice his ambassador to the Dalai Lama and a great land-owner (Petech 1977:93–94). His son was the general Tshul khrims rdo rje who, after an initially brilliant career, fell from grace as a result of his support for the “Uncle King”, bKra shis rnam rgyal of Purig, was tried in 1750 and probably executed (Schwieger 1997:224–5); at least, he disappears from history. But the Gya family was still sufficiently important after the Dogras’ conquest of Ladakh for its chief to be obliged to live in the court at Jammu as a hostage in the 1830s. When William Moorcroft passed through Gya in 1820, he visited the “Raja” as he refers to him, presumably following the local custom, named “Tsimma Panckhik”18 and found him living in a “house” in the village by the fields. Later he notes: “Opposite to the town [village], on a lofty ridge of the rocks, was a large pile of houses, formerly inhabited by the raja; and lower down one belonging to the lama” (Moorcroft and Trebek 1986:138–40). We have seen what little remains today. Finally, when Francke came this way in 1909, he reported that the old ruling family was extinct and all its records lost (Francke 1914:63). Conclusion Clearly there is a most urgent need for a full survey and documentation of all the ruins and ancient buildings in the Gya valley. The illustrations used here reveal an assemblage of sacred relics related to the Second Diffusion which are inferior in extent and number only to those in the desert between Shey and Ranbirpur, which includes Nyarma. The ruling family which sponsored and presided over this building work in the Gya valley cannot have been insignificant local chiefs. As might be expected, living as they did close to the southeastern borders of Ladakh-Maryul, these chiefs had a need for security and built several fortress settlements about which we know little at present.

17

18

This prophecy, made in the late 16th century when, according to the testimony of Azevedo, the chiefs of Gya must have still been the rulers of a reduced Maryul, hints at fascinating possibilities in the unknown history of the Gya chiefs . Perhaps ’Jigs med phun tshogs. See Francke 1926:227 note.

Historic Ruins in the Gya Valley

93

In parallel, more historic documentary sources are needed. As we have seen there is little in English but there may be something which can be quarried from historic Tibetan sources and much waiting to be discovered in monastic archives in Ladakh if Ladakhi historians explored them systematically and published their results in English. Even some remote collateral branch of the Gya chiefs’ family may still exist. And along the way we may come upon a better solution to the problem of the origin and extent of Maryul. Acknowledgements John Bray and Ken MacDonald provided the photographs which illustrate our paper and Abdul Ghani Sheikh acted as their guide to Gya in 2005 in unusually wet summer weather. Gerald Kozicz has generously shared the results of his architectural fieldwork in the Gya valley with us and his skilled draftsmanship has provided three of our illustrations. Bettina Zeisler has helped our understanding of the Tibetan language of the LDGR with respect to the inheritance of dPal gyi mgon and the founding of the kingdom of Maryul/Ladakh (although our conclusions are our own). We are deeply grateful to all of them.

Appendix: The war of Tsede (Tse lde) of Guge, 1083 CE

Philip Denwood In his publication The Kingdoms of Gu.ge Pu.hrang (1996), Roberto Vitali has drawn attention to a military campaign of 1083 CE which according to the mNga’ ris rgyal rabs was waged by king Tsede (rTse lde) of Guge in Western Tibet against “Gye sar” and “rGya”. Vitali identifies rGya with the place, Gya, and kingdom of that name in Ladakh. The text (Vitali 1996:72) runs as follows: gnam lha babs kyi Gyal po/ sa lhun grub gyi mnga’ bdag bod kyi lha btsan po rtse lde’i zhal mnga’ nas/ chu pho phag gi lo la/ gye sar ‘dul ba’i dbus byang nges pa thang nas gdan stegs nas/ Gya’i du ram thang du Gya dang snol thabs mdzad pa’i tshe sang nang na blon chen dbang grub dang/ sang nang ba chibs dpon snang grags dang/ sang nang ba’i sku tsha bo jo sras Gyal mtshan rdo rje/ gshen blon g.yung rdor dang/ sang nang ba gser rje mu ru dang lngas dpa’ ba’i dar mdzad pa’i ‘khrul ston nas/ Gya’i Gyan po bcu gsum skum nas/ chab srid phyag tu phul/ de nas gdan gtegs te/ byang go la gongs su/ gdan gting nas bzhugs pa’i dpung/ gye sar gyis

94

Howard

phyag phul nas gling gi mi bzhi phyag la btang te phyag ‘bul gyi skor len du ‘thengs pa zhu zhus pas/ spyan jus mdzad nas sang nang ba blon chen dbang grub dang snang grags gnyis rdzans nas . . . phul/ khral bca ‘bul ba mang po bsnams nas/ shar phyogs mdo smad la sogs pa sna gsum bGya’i gong kha dmag ru tshun chad mnga’ ‘og tu bsdus/ Vitali’s translation (Vitali 1996:123ff) is: This king, who descended from the sky, the mnga’.bdag (‘lord’) of the fertile land, Bod.kyi lha.btsan.po rTse.lde related: “In the water male pig year, as we proceeded (gdan.stegs sic for gdan.btegs) from the plain of the Byang.nges.pa (sic for Byang.ngos.pa-s), which was the centre [of the lands] captured by Gye.sar (sic for Ge.sar) Gye.sar ‘dul.ba’i dbus Byang.ngos.pa thang.nas gdan.btegs.nas), when [the kingdom of] rGya was defeated at Ram.thang of rGya as Sang.nang.na (sic for ba) blon. chen dBang.grub, Sang.nang.ba chibs.dpon sNang.grags, Sang.nang.ba’i sku.tsha.bo Jo.sras rGyal.mtshan rdo.rje, gShen.blon g.Yung.rdor, Sang. nang.ba gser.rje Mu.ru these five, exhibited miraculous performances of bravery, the thirteen headmen of rGya (rGya’i rgan.po bcu.gsum) withdrew [from the conquered area], [and] political power (chab.srid) was surrendered [by rGya Gesar]. Then, we proceeded further (gdan.gtegs sic for gdan.btegs). As the troops posted on the top of Byang.go.la were surrendered over by rGya Gye.sar, four men of the territory (gling.mi.bzhi) were sent as envoys. As they earnestly requested [me, rTse.lde] to hold a session of talks, a plan was made. Since Sang.nang.ba blon.chen dBang. grub and sNang.grags were sent, many treasures were obtained as tribute, . . .” As much was given in tribute, mDo.smad in the east as far as Gong.kha dmag.ru [with its] three hundred sna (“divisions”?) was brought under [rTse.lde’s] dominion. My translation: At the time when the heaven-descended king, master of the self-created earth, the Tibetan bTsan po Tsede, having moved on from Nges pa thang of Üjang (dBus byang) where Gye sar had been defeated, and given battle with the Gya at Du ram Plain of Gya, . . . [five ministers] . . . performed great martial feats, and thirteen kings [rgyal po; or “elders” rgan po] of Gya having been bent (into subjection), their dominion was given into his hands.

Historic Ruins in the Gya Valley

95

Then moving on, while he was staying on the Changgo La (Byang go la),19 Gye sar, making obeisance, sent four men of Ling (Gling) for the obeisance who said, “please come20 and accept the treasures we offer,” and he approved and sent the great Sang nang ba ministers dBang grub and sNang grags to whom they gave . . . [long passage listing the treasures] . . . They brought the many tributes and offerings, and the three eastern provinces including Dome (mDo smad) as far as Gong kha dmag ru of Gya21 were added to the dominions. This passage can only be fully interpreted if the place names are identified. Assuming with Vitali that Gye sar corresponds to the well-known title ‘Gesar’ famous from the Tibetan epic, what leaps to the eye at first sight is the association of Gesar with Ling, the name of the principality in northwest Kham which is Gesar’s kingdom in the epic, corresponding to the modern Lingtsang (gLing tshang). Vitali translates gling simply as “territory”. While this is not impossible, in my experience gling as a free-standing word normally means “island” or “continent”. The long list of “treasures” including weapons presented by Gesar is also highly reminiscent of similar passages in the epic. Lingtsang lies some 1700 km from Guge and in the opposite direction from Gya in Ladakh, perhaps an unlikely quarter for an adversary of Guge. However there is another apparently unambiguous reference in the passage to a place far to the east: (Gya’i) Gong kha dmag ru, identifiable as Gongka Maru (Gong ka ma ru) southeast of the Kokonor Lake: “the well known border of China and Yar lung Bod, where a rdo.rings of the Sino-Tibetan peace treaty was placed in 821–822” (Vitali 1996:331). The phrase mdo smad la sogs pa sna gsum recalls the mdo smad khams gsum recorded from the early categorisations of eastern Tibet, and Dome is thought by Uebach to comprise “parts of present day Khams and A-mdo south of the Yellow River” (Dotson 2006:273, 90, 93, 109. 207, 303). The way the passage on the extension of Tsede’s rule over these eastern regions follows immediately on the account of the campaign against Gesar and Gya without even a sentence break strongly suggests that the extension was a direct result of the campaign. To Vitali, the extension of the boundaries of Tsede’s kingdom as far east as this is “highly improbable” (Vitali 1996:331). However, if looked at in the context of contemporary trade routes and the historical background to the Gesar epic, 19 Reading dpung as dbung or dbus ‘while’ or dus ‘when’: if dpung ‘army’ is retained, a particle is missing either from dpung or phyag. 20 Reading ’thegs for ’thengs. 21 Reading Gya’i for bGya’i.

96

Howard

the claims made in the passage—though no doubt fantastic if implying much degree of actual control over these vast territories—can be seen to be based on plausible realities. I have argued elsewhere (Denwood 2005) for the existence of a Changthang Corridor or Musk Route running along the southern fringe of the Changthang Plateau of northern Tibet from Ladakh via the Panggong Lake across the gold mining area of Thok Jalung and continuing eastwards via the lakes: Dangra Yumtsho, Ziling Tsho and Tengri Nor (Nam tsho) to the Nakchu area of the upper Salween, then north-eastwards across the headwaters of the Yellow River to the Kokonor area. Archaeological evidence shows that in former centuries this route was significantly wetter than today, permitting colonisation in a chain of agricultural oases. This ties up with the picture derived from a study of the Gesar epic. R.A. Stein rightly cautions against identifying Gesar and others in the Ling version of the epic as historical characters and the events as historical events. Nevertheless, as he acknowledges, the general historical scenario envisaged in the first two chapters is clearly that of the eastern parts of this Changthang Corridor in or around the 11th century. Regular caravans of traders from Ladakh and Western Tibet (mNnga’ ris, no doubt including Guge) bearing gold and silver, and from China bearing silk and tea, provide a lucrative source of revenue for those able to extend their ‘protection’ over parts of the route. In the epic, the latter are the principality of Ling—more precisely, the young Gesar who is temporarily exiled from Ling to the upper Yellow River—and his rivals the “Hor ser” or Yellow Uighurs extending southwards from the Ganzhou area along the silk route. They are the precursors of more recent groups such as the Goloks of the upper Yellow River area, also notorious for preying on trading caravans. Stein has drawn attention to two Karmapa sources from which we learn that in about 1216 the monastery of Tshurbu northwest of Lhasa had been destroyed or badly damaged by the forces of Ling and Beri (Be ri) (Stein 1959:228, 239 n.). Since Beri is a well known neighbour of Lingtsang, it would seem that this refers to the same Ling to which the epic became attached, and that its armies had been engaged in a foray along the Changthang Corridor and down into the basin of the Kyichu. It is in the areas north of Lhasa that it may be possible to locate two of the places mentioned in the mNga’ ris Gyal rabs passage. In the phrase dbus byang nges pa thang Vitali takes dbus to mean “the centre of (the lands) conquered by Ge sar” in spite of the lack of a subject-marking particle on Ge sar. Byang nges pa he takes to mean “plain of the Byang ngos pa-s”, that is, the inhabitants of the Byang ngos area of Guge. Byang ngos pa is an odd form for a place name, and if the pa means ‘inhabitants’ one would expect a genitive particle.

Historic Ruins in the Gya Valley

97

However, there are other possibilities. One is the existence (in modern times at least) of a place called Üchang (dBus byang, Chinese Wujiang) at the eastern end of the Panggong Tsho, missed by Vitali. Other instances of Chang (Byang, ‘north’) used as a toponym are listed by Vitali himself in another article: including “The term Byang also denotes ‘Phan-po, the district bordering on lHa-sa, appearing in names such as Byang Rwa-sgreng and Byang sTag-lung” (Vitali 1997:1023 fn). This area north of Lhasa (in fact extending beyond Phenpo (‘Phan po) right up to the Tengri Nor and into Nakchukha) is also sometimes called Üchang (e.g. Drigung (‘Bri gung) Monastery is called “Üchang Drigung”). The c.1216 incursion to Tshurbu from Ling and Beri must have passed through this general Chang area of Ü (dBus) province. The pass between the Lhasa Valley and ‘Phenpo area of Üchang is known as the Phenpo Gola or just Gola (Tibetan spelling uncertain), which would seem to make it a plausible candidate for the Changgo La of the mNga’ ris Gyal rabs passage; as far as I know the Chang La (Byang la) in Ladakh is never called Changgo La. (There is also an Üchang monastery somewhere in the Derge/Katok area—probably quite close to Lingtsang.) Nges pa thang defies identification for the time being. There is a Nges kha rag listed in Das’s dictionary, but the only Kha rag otherwise known to me is in the valley of the Tsangpo northwest of the Yamdrok Tsho. On the route between this Üchang area north of Lhasa and Gongka Maru is the ancient royal burial ground whose Chinese name is Dulan. Could this be the ‘Du ram Plain’ of the mNga ris Gyal rabs passage? This is phonetically possible given conventional correspondences between Chinese and Tibetan. Vitali’s translation of this passage does not follow the grammar of the Tibetan. These places are certainly a long way from Guge, but modern-day traders have traversed comparable distances, and if Ladakhi trading caravans could travel the length of this route in the 11th century there is no reason why a small army could not do so. The important question now remains: who or what is Gya—evidently a kingdom, clan or people—in this passage? Gya of course commonly refers to China, but sometimes also to India. Vitali (1996:323) writes, “rGya cannot be either rGya.gar or rGya.nag, otherwise Guge would have been improbably engaged in a war with India or China, nations too imposing not to be defined with greater accuracy. A simple but conclusive consideration to rule out a reading of rGya as rGya.gar or rGya.nag is that no ruler of either is defined as Gesar”. Gya is certainly used to mean ‘China’ elsewhere in the mNga ris rgyal rabs,22 and it is used both for ‘China’ and for Gya of Ladakh in the La dvags rgyal rabs. 22

E.g. Vitali (1996:29, line 14).

98

Howard

The Gya of rGya’i Gong kha dmag ru would seem quite likely to be China, since this spot is well known as being on the Chinese frontier. Vitali assumes that Gesar is the ruler of rGya (east Ladakhi Gya): he goes so far as to call him “Gya Gesar” (an expression not occurring in the text) and erects an elaborate argument connecting him with the ‘Dardic’ population of the ‘Indo-Iranic borderlands’. This seems to be based on no more than the facts that (a) an 8th Century King of Kabul struck coins bearing the title Khrom ge sar; and (b) the La dvags rgyal rabs claims that in the 10th century Upper Ladakh was held by “members of the lineage of Gesar”. The link with Kabul seems to me highly tenuous rather than Vitali’s “definite”, particularly when one considers that the Dardic hypothesis has been exploded, while even the languages of the peoples formerly classified as ‘Dardic’ are no longer regarded as being Indo-Iranian by many authorities. R.A. Stein has clearly shown that the Tibetan Gesar epic crystallised in the 11th century at the earliest, and that its Ladakhi manifestations were imported subsequently from Eastern Tibet. From a consideration of, for example, the ideology of marriage rituals it is clear that the Gesar of the epic was then adopted as a culture hero in Ladakh in a totally unhistorical way.23 A close reading of our passage reveals that in fact Gesar is not closely associated there with Gya. It is after defeating Gesar that Tsede moves on to confront Gya, before returning to receive Gesar’s submission. Even more so than with the toponym element Chang, there is an embarrassment of places and human groups bearing the name Gya in Tibetan geography and history. The examples of such ethonyms as Hor, Sog po and Sum pa show that a given name can apply to several different ethnic groups as well as to a Tibetan group, clan or clans. Vitali himself has drawn attention to the Gya clan, members of which were influential in the area of southern Tibet/Bhutan from the Yarlung Dynasty period. One of the six clans of Ling in the epic is also called Gya. The many places called Gya or including the syllable in their names include rGya ma east of Lhasa, and rGya mda’ and rGya rdzong in Kongpo. The Gyade (rGya sde, ‘Gya communities’) of the basin of the upper Salween (sometimes called the rGya ma chu) on the Changthang Corridor, northwest of Lingtsang and northeast of Üchang, or their forerunners, come to mind as possibilities for the “Gya” of the mNga’ ris Gyal rabs. These nomadic communities number thirty-nine or in some lists thirteen (3×13=39): the number of “elders” or “kings” of Gya defeated by Tsede. In modern times, although Tibetan in culture and under Tibetan rather than Chinese administration, the Gyade have supposedly been “originally of Chinese stock” (Teichman 1922:49). In view of 23

E.g. Phylactou (1989, Chapter 9).

Historic Ruins in the Gya Valley

99

the pastoral economy of the area their origins can hardly be sought among Han Chinese, though very likely they stem at least partly from Uighur (the Hor ser of the Gesar epic) and/or Mongol groups. A possible scenario for the events of 1083 is as follows: an armed body sent from Guge, perhaps to escort a trading caravan, clashed with a similar force designated as Gesar from Lingtsang, also engaged in trying to control this trade route, somewhere in the area north of Lhasa. Proceeding on to the upper Salween and beyond to Dulan, a further clash or clashes with some group called Gya occurred. Returning to the Changgo La, north of Lhasa, formal submission was received from the defeated forces of Gesar. These victories, no doubt considerably exaggerated in official propaganda, provided the basis for some degree of control of this trade route and for the claim to ‘rule’ the whole stretch of territory from Guge along the Changthang Corridor to the Chinese frontier. If this interpretation of the episode as related in the mNga’ ris rgyal rabs is correct, it not only revises Vitali’s version of relations between Guge and Ladakh, but provides the earliest evidence for the use of the title Gesar in a historical context; and moreover associates this title with the principality of Ling.

An Archaeological Account of Ten Ancient Painted Chortens in Ladakh and Zanskar By Quentin Devers, Laurianne Bruneau and Martin Vernier1 In memory of André Alexander, his inspiring work and wonderful achievements. Introduction The most comprehensive study of the stupa in Ladakh remains that of Tucci published in Indo-Tibetica I (1932). In this book dealing with Tibet and surrounding Himalayan regions, the famous Italian Tibetologist translates various Tibetan texts dealing with the chorten (mchod rten), the Tibetan equivalent to the Sanskrit stūpa, and presents the material gathered during his expeditions, mainly consisting of clay tablets (tsha tsha). Since then only three publications dedicated to the stupa architecture of Ladakh have been published. Pieper (1980) makes an inventory of various buildings types of chortens in the Indus valley, between Lamayuru (bLa ma g.yu ru) and Hemis (He mis), and shows that there is a considerable variety in their architecture. Mention is made of the ‘signal value’ of some chortens but there is no comment about their origin or possible dating. In Stupa and its Technology: a Tibeto-Buddhist Perspective Dorjee (1996) describes important chortens between Hemis and Spituk (dPe thub). According to him they fall into two categories: late and early stupas. The architecture of the late stupas corresponds to the eight types of chortens acknowledged by the Tibetan tradition since the first quarter of the 14th century. Finally, the study of Ladakhi chortens by Kath Howard (1995) is devoted to monuments belonging to the 11–15th centuries. The author classifies them into five types: stupas associated with Rinchen Zangpo (Rin chen bzang po), Lotsawa (lo tsa ba), Lhabap (lha bab), Gomang (sgo mang) stupas and a fifth miscellaneous type. The chortens are dated according to their traditional asso1 All sites presented in this paper were documented by Devers in 2009, 2010 and 2011, with the exception of the chortens in Nang which were first documented by Vernier in 2003. This fieldwork was funded by the Centre de Recherche sur les Civilisations de l’Asie Orientale (UMR8155: CNRS / EPHE / Paris Diderot-Paris 7 / Collège de France) in 2009, and by the École Française d’Extrême-Orient in 2010 and 2011. All plans and elevations are by Vernier.

© koninklijke brill nv, leiden, ���4 | doi ��.��63/9789004271807_�06

An Archaeological Account

101

ciation with Rinchen Zangpo or their link with other remains such as temples. Since 2005 a ‘Stupa Project’ aiming at documenting and categorizing architectural concepts related to the development of Vajrayāna Buddhism in Ladakh has been undertaken by Kozicz (2010b). At the present state of research, no absolute dating can be provided for any of the ancient chortens of Ladakh. The task is difficult because the monuments are very often in an advanced state of decay due to the nature of the materials used for their construction and also because they are frequently rebuilt or whitewashed. The idea for this article originated in the summer of 2010 when one of the present authors witnessed the reconstruction of a painted chorten in Zangla (bZang la) in Zangskar (Zangs dkar), as will be discussed below. In terms of conservation this action is a major source of concern since a similar reshaping is intended for the neighbouring chorten of Malakartse Khar (Ma lag mkhar rtse mkhar), whose artistic and important historical value has been acknowledged by experts along with similar monuments in Nyarma (Nyṇar ma), Basgo (Bab sgo), Alchi (A lci), Sumda Chung (gSum mda’ chung), Mangyu (Mang rgyu), Lamayuru, Karsha (dKar sha) and Ichar (gYi char, I mTsar).2 If not properly restored, these treasures of Ladakhi religious architecture are threatened, as are other unknown or little-known chortens. The authors wish to draw the attention of locals and scholars to ten unknown or little-known monuments so that they can be appropriately preserved and studied in the future. In their capacity as archaeologists, the authors choose to stay out of art-historical debates, including questions about dating, and present the chortens according to their architecture. As duly noted by Tucci (1988a:15), Tibetan architects divide “[. . .] the building into two different parts: the basis or throne and the stupa as such, in the same way that is usually done for a statue [. . .].” Accordingly, the chortens presented below are classified under three types differentiated by the nature of the 2 On painted chortens, see: Malakartse Khar: Linrothe (2007a); Nyarma: Kozicz (2007b, 2010b—Nyarma. Stupa 1, Nyarma Northern Section Stupa Triad), and his article in this volume, Panglung (1983); Basgo: Francke (1914:87), K. Howard (1995, pl.4), Kozicz (2010b— Basgo Entrance stupa I and II); NIRLAC (2008b:18); Alchi: Francke (1914:42), Goepper (1993), Luczanits (1998:156–57), NIRLAC (2008a:26–27), Snellgrove and Skorupski (1977:24, 29, 78–79, ill.27, 72); Sumda Chung: NIRLAC (2008b:488); Mangyu: Ham (2010:36–41, 138–157), Kozicz (2010b—Mangyu Caitya stupa and Entrance stupa), Linrothe (1994), Luczanits (2004:170– 174), NIRLAC (2008a:264, 269); Lamayuru: Francke (1914:98); Kozicz (2010b—Lamayuru Entrance Stupa I and II, Luczanits (1998:157–59, 2005:87–89), NIRLAC (2008a:219), Snellgrove and Skorupski (1977:21); Karsha: K. Howard (1995: 67–68, pl.9); Linrothe (2002:83–87, 89–92); Linrothe and Kerin (2001); Snellgrove and Skorupski (1980: 48); Ichar: Linrothe (2002:87–89).

102

Devers, Bruneau and Vernier

Figure 4.1 Types A, B and C chortens. Drawings: Vernier, 2011.

throne, hereafter referred to as ‘plinth’ in order to follow a neutral ­architectural terminology. Type A chortens are characterized by a solid plinth, that of Type B is enterable, and that of Type C is opened on both sides forming a passage through which one can walk (Figure 4.1). The presentation of each type is followed by an individual description of the chortens concerned.

Type A

As stated above, this type of chorten is characterized by a solid plinth. However, the lower part of the chorten erected on this plinth presents a small opening revealing an inner painted space. Such monuments are known at Nyarma (see footnote 2 for bibliographical references) and two new specimens of this kind were documented at Ubshi (Ub shi) and Nang (Nang). Ubshi Francke (1914:66) mentioned several ancient chortens in the village of Ubshi, one of which bears paintings.3 It consists of a tall rectangular plinth on top of which rests a square-based chorten with two receding steps, equal in height, supporting a small hemispherical dome. Each element is separated from the other by a row of projecting horizontal slates (Figure 4.2). There is no upper structure but a thin wooden pole sticks out of the dome. The plinth is built of stones whereas the chorten is a mix of stones and mud bricks. The interior of the chorten is fully stone-roofed (Figure 4.3). The whole monument is plastered and white-washed. 3 Francke writes that there are several ancient chorten in the village of Ubshi, and that he examined the interior of one of them. His description of the paintings (colours, figures) matches the one documented by the present authors.

An Archaeological Account

Figure 4.2 Ubshi, plan and elevation. Drawings: Vernier, 2011.

103

104

Devers, Bruneau and Vernier

Figure 4.3 Ubshi, plan of stone roof. Corbels and lintels are view from underneath looking up. Stones with a darker colour rest on stones with a lighter colour. Area with hatching could not be directly seen as we were asked to enter inside the chorten. Drawings: Devers, 2011.

A low square opening (0.7m × 0.7m) at the base of the chorten on the southeastern wall conceals a small rectangular inner space (1.2m × 1.3m and 0.9m high) painted with a limited palette: red and white on a blue background. Although the murals are in a bad state of preservation, their original composition is visible. The back and sidewalls are centred on a damaged haloed deity painted within a square frame and surrounded by rows of miniature seated Buddhas (46 on the left and back walls, 52 on the right one). Two columns of five such Buddhas frame the opening on the fourth wall. The miniatures are outlined in black, thus clearly delineating the figures from the plain blue background (Figure 4.5). Friezes of curtains and geese (haṃsa) are painted on top of the murals, extending in some places to the stone lintels of the ceiling.

An Archaeological Account

105

Figure 4.4 Ubshi, right wall. This image is the result of the merging of several pictures—there might as such be some geometrical distortion. Photo: Devers, 2011.

There are traces of colouring on the latter but it is difficult to know whether it was originally fully painted or not. The chorten is still an object of devotion as shown by the fact that clay tablets (Figure 4.4) and used or incomplete sacred texts and threadbare prayer flags have been deposited within the chamber, while stones carved with sacred syllables (ma ni) rest on the outer walls. Nang In the village of Nang a row of decayed chortens, two of which were painted, mark the approach of a ruined castle locally known as Kenpa (Khen pa).4 Both chortens are entirely built of stone and exhibit a unique roofing technique. One is made of a squarish plinth on which rests a platform of similar shape, although slightly smaller in width and height. The two elements are separated by a row of horizontal stones (Figure 4.6). A heap of stones, dressed with mortar, covers the upper platform, recalling the hemispherical shape of a dome. Two openings in this platform (in the north-western and south-eastern walls, both 0.6m × 0.8m) reveal a painted space (1.7m × 1.9m and 1.1m high) arranged around a stone pillar supporting radiant stone lintels (Figure 4.7). The walls, the ceiling and the central pillar were once painted, as evidenced by remains of coating with orange-ochre lines as well as blue, green, red and white colouring. 4 These painted chortens were first documented by Vernier in 2003 thanks to information provided by Tsering Tundup, a Gelugpa (dGe lugs pa) lama from Zangskar trained at Karsha monastery.

106

Devers, Bruneau and Vernier

Figure 4.5 Ubshi, detail. Photo: Devers, 2011.

Although one can still distinguish halo lines and the faded remains of small seated figures, the paintings are too decayed to propose any identification. In the same row of chortens, a second monument provides a link between type A and type B chortens. It uses the same original roofing technique and conceals a similar small inner space. However, the opening is located in the plinth and is accessible at ground level.

Type B

Type B buildings are characterized by an enterable plinth supporting one or several chortens. In some cases this plinth, very long and large, is the most

An Archaeological Account

Figure 4.6 Nang, first chorten: plan and elevation. Photo: Vernier, 2011.

107

108

Devers, Bruneau and Vernier

Figure 4.7 Nang, first chorten: plan of stone-roof. Corbels and lintels are viewed from underneath, looking up. Stones with a darker colour rest on stones with a lighter colour. Photo: Devers, 2012.

prominent element of the monument.5 Famous examples of such chortens are known in Nyarma and Mangyu (see footnote 2 for bibliographical references).

5 Such chortens from Nyarma and Stok are described by K. Howard (1995:62). According to her, monuments with a large plinth on which is set one or several disproportionately small chorten(s) are known as ‘Lotsawa’ by Ladakhi scholars. She forwards a piece of information provided by Gelong Thupstan Paldan (Jammu & Kashmir Academy). This is confirmed by the recent publication of the gateway chortens of Basgo, Stok (NIRLAC 2008b: 18, 468) and Lamayuru (NIRLAC 2008: 219) locally known as Lotsawa.

An Archaeological Account

109

Four monuments of this type were documented at Nang, Stongde (sTong sde), Thikse (Khrig se / Khri rtse) and Stok (sTog). Nang, Second Chorten The other monument in Nang, also entirely built of stone, consists of a rectangular plinth on which rests the crumbling structure of a chorten, made of one or two platforms supporting a circular dome. An opening (0.4m × 0.8m) into the plinth in the south-eastern side leads to a small inner painted chamber (1.3m × 1.5m and 1.4m high). There is a smaller opening in the back wall meant for light only. The inner space and stone roof are identical to that of the chorten already described (Figure 4.9). The only difference is that the central pillar is square and composed of two cut stones on top of each other whereas that of the other chorten is made of one large, untouched, rounded stone. The inner walls of this second chorten retain more plaster, and therefore more traces of their former murals. The range of colours is similar with the addition of black lines. One can still distinguish parts of miniature chortens and a frieze on the top backside of the central pillar with red, blue and black lines, and a stripe of lotus petals. In this second chorten, and in a third unpainted one built close by with the same stone roofing technique, there are numerous clay tablets, some impressed with inscriptions in Nāgarī script. Their presence is noticeable since such tsha tshas are not frequent in Ladakh.6 Stongde Marpa Ling Gompa Gandal Chorten Located among other chortens in the vicinity of the monastery of Stongde, this painted chorten is locally known as ‘Gandal’.7 The monument is built of stones and composed of a square-ish plinth on which only remains the square base of a chorten. There are no remains of the upper structure which, according to 6 Francke and Tucci collected tsha tsha with ‘Indian characters’, either termed ‘late Gupta’, ‘ancient Nagari’ or ‘śāradā’ from various places in Ladakh. For a list of these clay tablets, see Francke (1914:111–117) and Tucci (1988:73–109). One of the models of tsha tsha found in Nang representing a sgo mang chorten closely resembles one found in Thikse by Tucci bearing the ye dharmā formula and “[. . .] a very corroded inscription in an Indian script of the X–XI century.” See: Tucci (1998:76, No.16). 7 This name might be a derivation of gSal ldan, the location where Indravami built the Descent from Heaven chorten, according to Jamyang (Jamyang Zhaypa Ngawang Dorjee, mChod rten gyi thig rtsa mdor bdus, Collected works of Jamyang Zhaypa, Vol. I, New Delhi 1974) as quoted in Dorjee (1996:19). The name of the builders, localities and even of the eight types of stupa of the Tathāgata acknowledged by the Tibetan tradition varies in the sources: Dorjee (1996:11– 21); Tucci (1988: 21–24).

110

Devers, Bruneau and Vernier

Figure 4.8 Nang, second chorten: plan and elevation. Photo: Vernier, 2011.

one of the senior monks of Stongde, collapsed about 40–50 years ago. Since then, a roundish undressed low wall has been built to replace the collapsed northern corner probably in order to keep local herds out of the chorten. The opening leading to the small inner painted chamber (1.5m × 1.8m) was in the now crumbled north-eastern wall.

An Archaeological Account

111

Figure 4.9 Nang, second chorten: plan of stone-roof. Corbels and lintels are viewed from underneath, looking up. Stones with a darker colour rest on stones with a lighter colour. Area with cross-hatching is roofed with a succession of small corbels, the progressive overhang of which covers the area. Photo: Devers, 2012.

The inner paintings are slowly getting covered by mortar melting from the upper level of the walls and are, as a result, in a bad state of conservation. The wall facing the entrance bore clay sculptures as indicated by plastered aureoles and banners as well as holes left by the struts used to attach them. The two sidewalls display a mandala, of which only the outer circle is still visible on the partly collapsed right wall. The mandala on the left wall is surrounded by seated Buddhas in various mūdras (Figures 4.11, 4.12 & 4.12 ). Some have their complexion rendered by means of colour gradient. The collapsed entrance wall retains traces of a protective deity in pratyāliḍha. This posture and jewellery details suggest the representation of Māhākāla (Tib. mGon po). The small

112

Figure 4.10

Devers, Bruneau and Vernier

Stongde Gandal chorten: plan and elevation. Photo: Vernier, 2011.

An Archaeological Account

Figure 4.11

113

Stongde Gandal chorten: left wall. Photo: Devers, 2010.

figure of Śrīdevī (Tib. dPal ldan lha mo) above it, in the upper right corner, corroborates this identification, as does another female protector, seated on a peacock right behind her. The red fish she holds in her right hand and her pointed ears correspond to a form of Vārāhī (Tib. Phag mo). A series of bands (curtains, crossed flowers, oval motifs) runs around the chamber on top of all four walls. The faded murals show a limited palette of red, blue and white. Some areas were repainted, as can be seen at several spots on the upper part of the murals where the friezes were patched with rougher mortar painted in brighter colours, thus pointing to an ancient repair of the chorten. It should be noted that most characters depicted in the murals had their faces deliberately scratched out. Thikse Double Chamber Chorten8 This chorten is located in the vicinity of a ruined temple at Thikse locally known as Kiki Lhakhang. The wider environment includes three other ruins of temples, two rows of 108 chortens and another set of three painted chortens 8 This chorten was first reported by Kozicz during the 13th conference of the IALS in 2007 (see Kozicz in this volume). It is referenced in Kozicz (2010) as ‘Nyarma Northern Section. Stupa with the Hidden Chamber’.

114

Figure 4.12

Devers, Bruneau and Vernier

Stongde Gandal chorten: detail of a miniature Buddha. Photo: Devers, 2010.

(Kozicz 2007b). The monument is composed of a high plinth built of stones, on which lies a crumbled brick chorten. It seems to have been composed of a rectangular base supporting two rectangular steps in turn supporting a large circular flatten dome. There is no upper element. The inner space of the monument, damaged by successive whitewashing, is painted at two levels accessible by two openings: one in the plinth and one at the bottom of the chorten, respectively on the south-eastern and northeastern sides. In the ground floor chamber, at least two brick walls were built in addition to the outside stone walls, creating a double wall structure. The wall facing the entrance is damaged, and a hole in it reveals a small space between the brick and the stone wall (Figure 4.14) The murals of the ground floor are painted using variations of green, blue, red and white, while black is used for outlining. The walls are covered with miniature seated Buddhas, the number of which cannot be asserted as only some parts of the paintings have survived

An Archaeological Account

Figure 4.13

Thikse: plan and elevation. Photo: Vernier, 2011.

115

116

Devers, Bruneau and Vernier

the process of decay (Figure 4.15 & 4.15) These small figures once framed central ones of bigger size that appear to have been enclosed within an elaborate architectural frame. Symmetrically disposed between the rows of miniature Buddhas are lotus buds motifs and white rectangles. A closer look reveals that the latter were once filled with Tibetan U-chen (dbu can) inscriptions. Most rectangles and Buddhas’ faces were cautiously abraded. While there are bands of painted curtains and geometric designs running on the upper part of the ground floor murals there is a frieze of geese on the second level. There the northern wall, seemingly displaying a Buddha flanked by two standing Bodhisattvas, is best preserved. The space between the two levels is open, although stone lintels coming out of the lower walls, at the level of the painted curtains, may indicate an original intermediary ceiling. Between the two floors there is a painted row of chortens of various types resting on a frieze of lotus with down-turned petals. There is no painted or wooden ceiling and the inner dome reveals the mud brick structure of the chorten.

Figure 4.14

Thikse: back wall of the ground floor. Photo: Devers, 2009.

An Archaeological Account

Figure 4.15

117

Thikse: detail of miniature Buddhas. Photo: Devers 2009.

Stok Kadam Chorten9 Mentioned to us by André Alexander, the late co-director of the Tibet Heritage Fund, in the spring of 2011, the Stok Kadam (sTog bka’ gdams) chorten is located in the vicinity of a ruined temple locally known as Lotsawa Lhakhang and other groups of chortens. The whole structure is white-washed and built with bricks except for the lower part of the walls, which is made of stones. The high square plinth is entered through a low door (about 0.8m wide and 1.1m high) on the northern side. A row of horizontal stones marks the top of the plinth which supports a crumbled chorten. There might have been a base or steps but only the tall, slightly conic, dome with an opening, located above the door of the plinth, is preserved. A row of stones crowns the top of the dome. The inner space of the plinth is unique: a small corridor leads to a central space under the dome and in the back wall, opposite the corridor, there is a niche. The chamber opens on two parallel side corridors, which, altogether, 9 A presentation of this chorten was made by Kozciz under the form of a poster at the 21st conference of the European Association for South Asian Archaeology and Art held in Paris in July 2012. A short description and a photograph of this chorten under the name ‘Lotsava’ can be found in NIRLAC (2008b:468).

118

Figure 4.16

Devers, Bruneau and Vernier

Stok Kadam chorten: plan and elevation. Photo: Vernier, 2011.

delineate an area of about 4.3m×4.3m. All three corridors are covered by a ceiling (2.5m high) entirely made of stone lintels (Figure 4.17). This technique is remarkably well suited to the ground plan, reminding that of a temple. Whether the roofing technique was chosen because it fits well the plan or the

An Archaeological Account

Figure 4.17

119

Stok Kadam chorten: plan of stone-roof. Corbels and lintels are viewed from underneath, looking up. Stones with a darker colour rest on stones with a lighter colour. Photo: Devers, 2012.

other way around remains an open question. Above the chamber two stone lintels indicate the existence of an intermediary ceiling, most likely opened so that the light passing through the opening of the dome reached the core of the building. All 17 walls were once painted but unfortunately only the upper half of the murals is preserved. The colouring includes yellow ochre, deep red, greenish

120

Devers, Bruneau and Vernier

earth, black and white. The walls display three rows of mutilated seated Buddhas in various mūdras (western, eastern and southern walls), numbering 108— if the murals were symmetrical—and representations of varied chortens (northern walls) (Figures 4.18, 4.19 & 4.19) The top of the walls retains a frieze of curtains under a band of geese. Two different styles of curtain friezes suggest two stages of painting. A third, more recent, stage corresponds to the repainting of the miniature chortens on the walls of the entrance corridor and that of a small portion of the frieze of curtains on a wall of the southern section. The side corridors and the back niche were used over the centuries to deposit tsha tsha, the layer of which reached half a metre in height in some places. In 2012, restoration work was carried out by the villagers in the chorten. It included the putting up of a wooden doorframe, wooden columns to support the roof and wooden shelves to sort the tsha tsha.

Figure 4.18

Stok Kadam chorten: western-most wall. Photo: Devers, 2011.

An Archaeological Account

Figure 4.19



121

Stok Kadam chorten: detail of a miniature Buddha. Photo: Devers, 2011.

Type C

Like the previously described types, Type C is composed of a high rectangular plinth on which is erected the chorten proper. Its particularity is that it has an opening into the plinth through which one may pass and raise one’s head to look at the hollow, sometimes painted, inner core of the chorten. In Ladakh, this type of chorten is locally known as Kagan, Kakani or Kankani.10 The conventional English expression for this kind of building is ‘gateway chorten’. Such monuments are known all over the Tibetan cultural area. In Ladakh early examples are located in Karsha, Lamayuru, Basgo, Alchi and Mangyu (see footnote 2 for bibliographical references). The authors have 10

On the etymology of the term, see Francke (1914:87, 98); Goepper (1993:140); Linrothe (2006:171–74).

122

Figure 4.20

Devers, Bruneau and Vernier

Nyoma Khawaling chorten: plan and elevation. Photo: Vernier, 2011.

An Archaeological Account

123

d­ ocumented four: two at Nyoma (Nyo ma), one at Shera (Wylie unknown) and one at Zangla. Nyoma Khawaling Main Gateway Chorten11 This large gateway chorten belongs to a group of chortens erected below the monastery, ancient fortification and settlement of Nyoma. A mani wall and two small three-tiered square-ish chortens were built right up against the plinth of the gateway chorten, respectively on the eastern and western walls. The tall square-ish plinth and the chorten resting on top of it are built of stones that were formerly plastered. The remains of red paint and plastered motifs are still visible on the plinth and the square-ish base of the chorten. The upper part of both elements is enhanced by cornices. The base of the chorten supports a flight of four receding steps, equal in height. A circular dome, now partly sinking into the upper step, is crowned by a rectangular harmikā surmounted by a central pole. Two very low and narrow doors on the northern and southern sides of the plinth enable one to pass underneath the chorten and discover a surprisingly rich inner iconographic programme. Francke (1914:57) reports the discovery of this chorten by Dr F.E. Shawe who wrote that it was “. . . by far the finest piece of mchod-rten decoration” he had seen.12 The inner space of the chorten is cubic (2.2m × 2.2m and 2m high) with four painted sidewalls and a decorated wooden ceiling. Photographs of the ceiling and two details of the walls were recently published by Ham (2010:27–29) and NIRLAC (2008c:362).13 Below the murals, at the base of the walls, stone consoles may indicate an original lower ceiling, underneath the paintings. The collapsed dome is likely at the origin of the cracks observed on the walls, but luckily reinforcement by wooden pillars ensured a fair state of preservation of the murals. The painted walls and ceiling are predominantly dark blue and red, with the additional use of pink, green, white and very thin black lines. The composition of the four sidewalls is identical: there is a main deity surrounded by six secondary ones horizontally framed by two rows of minor divinities (Figures 4.21, 4.21, 4.22, 4.23 & 4.24). The complexion of the figures is rendered by means of colour gradient. The murals are arranged around four central enthroned 11 12 13

This chorten is referenced as ‘Nyoma Stupa 1’ in Kozicz (2010). In NIRLAC (2008c:362) it is referred to as ‘Kaling’. Francke reproduces a letter of Dr. Shawe dated 19th July 1906. A paper entitled The Painted Stupa at Nyoma / Ladakh is under preparation by Ham (2010:173).

124

Devers, Bruneau and Vernier

Figure 4.21

Nyoma Khawaling chorten: east wall. This image is the result of the merging of several pictures and there may therefore be some geometrical distortion. Photo: Devers, 2011.

Figure 4.22

Nyoma Khawaling chorten: south wall. This image is the result of the merging of several pictures and there may be some geometrical distortion. Photo: Devers, 2011.

An Archaeological Account

125

Figure 4.23

Nyoma Khawaling chorten: west wall. This image is the result of the merging of several pictures and there may be be some geometrical distortion. Photo: Devers, 2011.

Figure 4.24

Nyoma Khawaling chorten: north wall. This image is the result of the merging of several pictures and there may be some geometrical distortion. Photo: Devers, 2011.

126

Devers, Bruneau and Vernier

cosmic Buddhas (Skr. Jina, Tib. rGyal ba), as indicated by their individual colours, mudrās and vehicles (Skr. vāhana), flanked by two standing Bodhisattvas. This central group is surrounded by two seated divinities at the shoulder level and four, also seated, painted below. This group of seven deities is framed by an upper and lower row of varied standing and seated figures numbering eleven for the former and nine to eleven for the latter. The lower band depicts invariably one of the lokapālas in the middle surrounded by siddhas (Tib. grub thob) as well as wrathful and protective deities (Figure 4.25).14 Below are narrative scenes of the Buddha’s life (Figure 4.25). They begin on the eastern wall with a first set of eight scenes in a chronological sequence, but they are then in a mixed order and unevenly distributed on the three remaining walls (seven scenes on the southern wall, nine on the western and northern ones). The bottom of the painted area is dotted at some places with Tibetan U-med (dbu med) lettering partly erased. The depiction of scenes of the Buddha’s life in the lower register of paintings is typical of Drigung (’Bri gung) shrines as evidenced for example in the Lakhang Soma (lha khang so ma) at Alchi, Wanla Sumtsek (Wam le gSum brtsegs), the Guru Lhakhang (Gu ru lha khang) near Phyang (Phyi dbang) or the Secret Room of Lingshed (Ling shed) monastery (Linrothe 2007a:51). The association of Nyoma Khawaling gateway chorten with the so-called Central Tibet-derived style of painting is confirmed by the colour palette and stylistic details such as the Buddhas’ thrones (for example the scrollwork emerging from the makaras (Tib. chu srin) beside the head nimbus and Garuḍa at the top of the latter), their stripped dhotis and peculiar crowns (triangular crests set into three tiers “[. . .] surmounted by a broad chignon with a fan-shaped splay of ribbon at each side of the crown and a round flower above each ear” (Linrothe and Kerin 2001:59). Another stylistic marker is the triangular plaque above the shoulders, at the intersection of the head and body nimbus, of some figures. The ceiling, conceived as a mandala, is made of wooden planks forming five concentric squares which diminish in size as they rise in height (Figure 4.26). Both vertical and horizontal surfaces are painted. The four outer corners of 14

Noticeable is the representation of a dark-skinned Indian siddha naked with a yogapata band around his knees. Similar images are known at Alchi, in the Sumtsek and the Great Stupa, as well as in the Sumtsek of Wanla. This siddha is either identified as Nāropa (Goepper 1996a:102, 109; Tsering 2009:49) or Padampa Sangye (Linrothe 2007b:65). The representation, one beside the other, of the goddesses Śrīdevī (Tib. dPal ldan lhamo) and Remati (Re mati) is another highlight of the murals. Similar representations of Remati are known in Alchi (one in the Sumtsek, one in the Dukhang, one in the Manjushri Lhakhang and one in the Lakhang Soma) and Thikse.

An Archaeological Account

Figure 4.25

127

Nyoma Khawaling chorten: details: Māyā’s dream (upper left), Shākyamuni’s first step (upper right), the lokapāla Virūḍhaka (lower left), the siddha Nāropa (lower right). Photo: Devers, 2011.

the ceiling are each marked by a painted vase of plenty (Skr. pūrṇaghaṭa). The two outermost squares, supported by wooden consoles, are painted with scrollwork on their horizontal surfaces and geese on the vertical ones. The next two squares exhibit horizontally golden mantras in Tibetan U-chen script in relief and miniature seated Buddhas and deities vertically. The topmost, inner square bears the four gates of the mandala on its horizontal surface and miniature deities on the vertical one. The centre of the mandala is occupied by a representation of Akṣhobya in typical bhūmiśparsa mudrā and seated upon his emblem, the vajra. He is surrounded by eight secondary deities placed in lotus petals. All nine central figures are in clay and gilded. Among the inscriptions on the outer squares of the ceiling one can read the mantra of Akṣhobya Om kankani kankani. This tends to confirm the hypothesis that Kankani chortens are so-called in reference to this particular dhāraṇī written on their inner walls (Linrothe 2006:172–73).

128

Devers, Bruneau and Vernier

Figure 4.26

Nyoma Khawaling chorten: ceiling. Photo: Devers, 2011.

Beyond doubt, a detailed study of the iconographic programme of Nyoma Khawaling chorten will prove to be very valuable for the history of early Buddhist painting not only in Ladakh but also in the whole of the Western Himalayas. Nyoma Smaller Gateway Chorten15 In the vicinity of the Nyoma Khawaling Chorten is another, smaller, painted gateway chorten. This is not reported by Shawe or Francke although it stands right next to the one already described, the passage of the former being in the axis of the door of the latter. Smaller in size, it is also entirely built of stone, once plastered. The squarish base of the chorten supports two receding steps, apparently equal in height. The upper part of the monument is a mix of melted 15

This chorten is referenced as ‘Nyoma stupa 2’ in Kozicz (2010b).

An Archaeological Account

Figure 4.27

129

Nyoma smaller gateway chorten: plan and elevation. Photo: Vernier, 2011.

earth and stones. The inner space is cubic (1.4m × 1.5m and 1.1m high) and topped by a lantern ceiling. As in the chorten of Thikse previously mentioned, there are inner mud bricks walls. Those are supported by wooden beams resting on a layer of stones. Although two of the four walls are not well preserved, the inner composition of the murals is clearly visible. The colour range is limited to shades of red and black on a white background thus creating a high contrast effect. Some figures are unfinished: one can clearly see ochre preliminary drawings, while the finished figures are drawn with black lines. The northern wall shows a central seated and enthroned cosmic Buddha flanked by two standing Bodhisattvas

130

Devers, Bruneau and Vernier

and surrounded by four secondary seated deities. (Figure 4.28) They face a mandala on the southern wall framed by two vertical rows of miniature deities. (Figure 4.29) The opposite eastern and western walls are covered with five horizontal rows of six seated Buddhas. A band of geese runs on top of the murals. The lantern ceiling is built of stone lintels coated and painted, although partly unfinished. There are beautiful geometric motifs and scrollwork and one can faintly distinguish depictions of deities. Shera Gomang Gateway Chorten The gateway chorten at Shera (She ra) marks the approach of a beautifully preserved ancient fortified settlement, locally said to have been besieged by the armies of the Fifth Dalai Lama. The plinth, built of stones, has its upper part decorated by a row of plastered downward pointed lotus petals, an outer

Figure 4.28

Nyoma smaller gateway chorten: detail of northern wall. Photo: Devers, 2011.

An Archaeological Account

Figure 4.29

131

Nyoma smaller gateway chorten: southern wall. This image is the result of the merging of several pictures and there may be some geometrical distortion. Photo: Devers, 2011.

decoration rarely preserved on Ladakhi chortens.16 The brick-built chorten is composed of a square base on which lie three receding cross-shaped steps decreasing in height. There are three openings in each step; accordingly the model of chorten is that known as “Many Doors” Stupa (sgo mang mchod rten). A high circular dome tops the structure. Each element of the chorten is separated from the other by projecting slates (Figure 4.30). The passage in the plinth is roofed with wooden beams. Below the chorten, within each wall of the passage there is a niche marking the location of a now gone statue. 16

On lotus decoration see: K.Howard (1995:69). Other examples of chortens with downwardpointing petals plastered on their plinths are known at Tragkhung Kowache (Ham 2010— the righthand image on page 25 is identical to the one shown in Kozicz (2010b) as ‘Tragkhung Kowache Stupa with Niches’); and Nyarma (Kozicz 2010b: Nyarma Southern Section. Stupa 7). Other specimens were documented by the present authors at Shera, Shernos and Ensa (’En sa, dBen sa). For the chortens of Ensa see also Kozicz (2010b). Chortens with upward-pointing plastered petals are found at Sabu (K. Howard 1995:Pl.11), Rumtse (Rum rtse) and Trakhung Kowache (Francke 1914:plate XXXI, which is erroneously labelled as being in Gya (rGya); Kozicz (2010b – ‘Rumtse Stupa with Niches, Trakhung Kowache Stupa with Lotus Plinth). A crumbled chorten with plastered trifoliate leaves and upwardpointing lotus petals was documented by Devers at Staglung (Stag lung) near Nyoma.

132

Figure 4.30

Devers, Bruneau and Vernier

Shera: plan and elevation. Drawing: Vernier, 2011.

An Archaeological Account

133

The walls of these niches were also once ornamented by two painted standing figures of which contour lines only remain. The inner space of the chorten (1.8m × 2m) is composed of four painted sidewalls covered by a painted wooden lantern ceiling. Unfortunately, successive whitewashing over the years has damaged the paintings, and the plaster is coming off. A simple range of colours is used: brown, blue, orange ochre and red, all applied on a white background. The murals give the impression of an unfinished work. Various stages of drawing completion are obvious, some figures being left as preliminary sketches while others are more finished. Colour gradient is used to render the complexion of some characters. The four walls are composed in the same manner: in the middle is a rectangular frame containing a large seated enthroned deity flanked by two standing Bodhisattvas and two seated monastic figures (Figure 4.31). The four central deities and accompanying attendants have their crowns and jewels in gold paint (Figure 4.32). Around the rectangular frames are rows of seated Buddhas (68 on the north western wall, 61 on the north-eastern and south-eastern walls, 55 on the south-western wall). There are fewer seated Buddhas on the latter because below the main figure is the depiction of a drinking scene. There are monastic and lay figures seated in two rows. The upper row is composed of three monks facing a couple of lay (?) persons holding a cup. The lower row shows three lay persons, a man with arms surrounded by two ladies: all are holding a cup offered by two attendants. One of them is represented kneeling while the other is standing and carrying one of the six jars represented. The lantern ceiling retains painted geometrical motifs in the corners and its structure, made of wooden boards, is perfectly visible. Zangla mkhar Gateway Chorten, a Preoccupying Case The gateway chorten of the ancient palace of Zangla, the ancient capital of lower Zanskar, is located among other chortens and mani walls marking the approach of the ruined site. This particular chorten is a concerning case for the conservation of Ladakhi heritage. In 2009 it was restored by means of cement, not only destroying its original shape forever—and consequently any archaeological information—but also endangering the inner murals by sealing with concrete what was once a breathing earthen structure (Figure 4.34).17 17

The ‘restoration’ was carried out by the Italian association Stupa Onlus. For the period 2011–2013 this association planned similar work for the painted chortens of Malakartse Khar in Zangla and the Kadampa (bka’ gdams pa) chorten in Karsha as well as Shey’s (Shel) Gomang chorten (locally known as ‘Nepali’ chorten) and other chortens in Tingmosgang (sTing mo sgang), Lamayuru, Mulbekh (Mul bhe), Rangdum (Rang ’dum) and Phey (Phye).

134

Devers, Bruneau and Vernier

Figure 4.31

Shera: north-western wall. This image is the result of the merging of several pictures and there may be some geometrical distortion. Photo: Devers, 2010.

Figure 4.32

Shera: detail. Photo: Devers, 2010.

An Archaeological Account

Figure 4.33

Zangla: plan and elevation. Drawing: Vernier, 2011.

135

136

Figure 4.34

Devers, Bruneau and Vernier

Zangla: chorten after action executed by Stupa Onlus. Photo: Devers, 2010.

The original rectangular plinth was built of stones. Round wooden beams lay on top to support the earthen structure of the chorten, which was composed of a squarish base topped by three seemingly circular sections increasing in height. The intermediary section retained projecting elements on one side, probably symbolising steps (Figure 4.33). The type of chorten that served as model thus might have been ‘the Descent from Heaven’ (lha bab). A decayed small hemispherical dome from which a pole came out crowned the chorten. No upper elements (harmikā, umbrellas or finial) were to be seen before the restoration of the chorten. The section under the steps, the steps and the dome were separated from one another by protruding slates resting on rounded sticks. The whole structure was originally plastered. Over the passage, the inner base of the chorten is partially closed by a frame made of wooden boards. The four corners are dotted with round flower designs. The boards on either side of the opening repeat the same designs: the traditional endless knot pattern (dpal be’u) and a design composed of alternative spirals framed by bands of pearls. The inner space of the chorten is cubic. The four painted sidewalls and the lantern wooden ceiling are painted using predominantly orange, red, green, black and white. Although the murals are

An Archaeological Account

137

not well preserved, their composition, identical on the four sidewalls, is still identifiable (Figure 4.35). In the middle, there is a large enthroned seated deity painted within a square frame. The figure is seated on a lotus throne supported by pillars and alternate facing elephants and lions. The main deity is surrounded by rows of seated miniature Buddhas. The composition of one of the walls is distinguished from the others by the representation of both male and female laymen immediately below the throne of the central figure, a common way to include sponsors and donors. The boards of the ceiling are painted with alternate flower, geometric and check board patterns, the central top space being occupied by a representation of a multi-headed Vairocana, as confirmed by the presence of lions on either side (Figure 4.36). The two well-preserved central figures of the sidewalls are also accompanied by their vehicles (a Garuḍa and a horse). The complexion, mudrā and orientation confirm that the iconographic programme of the chorten corresponds to that of the five Jinas.

Figure 4.35

Zangla: one of the walls (proportions may be affected due to perspective correction). Photo: Devers, 2010.

138

Devers, Bruneau and Vernier

Figure 4.36

Zangla: detail of the ceiling. Photo: Devers, 2010.

Conclusion About half of the ten painted chortens described in this paper mark the approach of ruined settlements and fortifications (at Nang, Nyoma, Shera and Zangla), while the others are linked to religious sites.18 All three types described above are hybrid: they have the external appearance of a chorten but their core, sheltering paintings and sometimes sculptures, is similar to that of a shrine. The iconographic programme occupies the inner space either of the plinth or of the chorten, sometimes both, as in the case of Thikse. When occupying the plinth, the main deity is represented on the back wall and the entrance wall bears protective deities, as in Stongde for example. When occupying the chorten, as it is the case for half of them (Ubshi, Nang first chorten, both Nyoma chortens, Shera and Zangla), the ceiling is invariably painted. It is even the focus of composition in gateway (Type C) chortens: it is where the main deity is depicted while four secondary ones are represented at the centre 18

In Ubshi the environment of the chorten has likely changed a lot over the years and the chorten is now standing in the middle of the village. In its present state, it is therefore not possible to define its original archaeological context.

An Archaeological Account

139

of the walls. When the inner decoration is preserved well enough to enable the identification of the divinities, as in Nyoma Khawaling chorten, Shera Gomang chorten and Zangla, the iconographic programme corresponds to the five Jinas. This organisation is known in other gateway chortens in Karsha, Ichar and Alchi (Linrothe 2002; Linrothe and Kerin 2001:55–57; Luczanits 1998:156). All three types of chortens are characterized by the fact that the paintings are seen with difficulty. In types A and B chortens the narrowness of the openings undoubtedly shows that the murals were not meant to be seen, at least not by everyone. In type C chortens the paintings are very high up, beyond the reach of the viewer.19 From a construction point of view all but one of the chortens (Nyoma’s Khawaling chorten) have a lantern ceiling. The ceilings are made of wooden boards, apart from Nyoma’s second chorten in which the lantern is made of stone lintels. In fact, half the chortens described above feature stone roofs: Ubshi, both Nang chortens, Stok and Nyoma’s second chorten. Although usually not noticed, such a roofing technique is used for other well-known painted chortens, as in Malakartse (Linrothe 2007a) and Thikse-Nyarma (Kozicz 2007b), and also unpainted ones, as observed by the authors during fieldwork. In two cases, in Nang, the roofing technique may even have affected the iconographical programme. Indeed, a construction with a central stone pillar might have influenced the conception of the murals, as a central element had to be painted in addition to the four walls of the chamber.20 It would have been interesting to see how the artists had responded and adapted to it, but unfortunately the paintings are too decayed. We propose to link the use of stone roofing in these chortens to a building tradition that shares common roots with the numerous ancient stone-roofed buildings (watchtowers, castles, fortified and unfortified settlements) documented by the present authors throughout Ladakh and in Upper Tibet by Bellezza (2008:32–37, 56–57). In the latter region the use of stone roofs for early, small, Buddhist structures is also documented.21 19

20

21

For a possible interpretation of the ‘invisibility’ of the paintings in gateway chortens (type C) see Linrothe (2002:94–95). His hypothesis that “These images are not made for viewing; indeed they are rendered more powerful by being unseen. For the patron, merit accrues from the making and the presence of the external structure in the landscape, but not by subsequent acts of viewing the interior paintings”. This principle could also be applied to Type A and B chortens. This configuration recalls that of a mandala. The construction with a central pillar might have been intentional thus tending to a three-dimensional mandala with murals organized accordingly. One wonders whether the roofing technique was adapted to such a design or the other way round. Private communication with John Vincent Bellezza, June 2011.

140

Devers, Bruneau and Vernier

All ten painted chortens described above have a stone plinth supporting a brick-built or mixed (stone and bricks) chorten.22 An interesting feature observed in Thikse and Nyoma is the building of an internal brick wall in addition to the outer stone one. This element could be observed only in two damaged chortens where the internal structure is visible and it is possible that others use such internal brick walls. The reason for this construction feature might that brick walls have a better resistance to water infiltration, due to the inherent absorptive capacity of this material’s homogeneous mass. Indeed, we repeatedly observed that coatings, and consequently paintings, are better preserved when applied on a mud brick support as opposed as one in stone masonry. Turning to the murals, features such as bands of geese running on top of the walls, the rendering of the characters’ complexion by means of colour gradient or else specific topics such as Buddha’s life scenes and drinking scenes are without doubt artistically and historically significant. All deserve a detailed iconographic and stylistic analysis by competent art historians. The aim of our short description of each of these chortens was to show their richness and great potential for the study of the history of Buddhism in Ladakh and the Western Himalayas. It is of the utmost importance to protect and preserve them. As we have seen in Zangla and can be observed throughout Ladakh, the action undertaken by some NGOs and various other organisations is, in some cases, a threat to Ladakh’s cultural heritage. Addressing this issue in a previous note (Vernier, Bruneau and Devers 2011) and workshop,23 we reassert here the urgent need for an authority and legislation to protect Ladakhi heritage, which, considering its richness and the destructions that have occurred in the neighbouring regions, is the unique key witness of a history that goes far beyond its frontiers. 22 23

The plinth of Stok chorten is built of bricks and has foundations in stone. The authors together with André Alexander organised workshops in May and August 2011 at the Central Asian Museum in Leh to bring together the different actors in conservation in Ladakh with the aim to improve cooperation and protection of Ladakhi built heritage. This workshop resulted in the writing of a proposal presented to the Ladakh Autonomous Hill Council for the creation of a ‘Heritage Authority’ and legislation for ‘Ladakh Protected Monuments’.

The Chorten (mChod rten) with the Secret Chamber near Nyarma Gerald Kozicz In Ladakh there are no surviving architectural structures that can be securely dated to a period earlier than the final decades of the first millennium CE, when the rapid spread of Buddhism resulted in the construction of a large number of religious monuments all over the Western Himalayan region. The regional centre, from which this second Diffusion (Phyi dar) of the Buddhist faith pervaded the social and political entities of the region was the monastic complex of Nyarma (Nyar ma) in the Indus valley south of Leh. This part of today’s Ladakh, by that time known as Mar yul, had come under control of the Western Tibetan kingdom of Guge and Purang by the middle of the 10th century. The foundation of Nyarma is ascribed to Rinchen Zangpo (958–1055 CE), the ‘Great Translator’ (Lotsawa—lo tsa ba) who brought numerous Buddhist scriptures back to the region from Kashmir.1 The fact that the kings of Guge and Purang provided the funds for the erection of their second largest religious centre in the very northern border region of their empire is noteworthy and may be explained by the fertility and prosperity of that part of the Indus Valley. Nyarma apparently became the key-site for the sustainable spread of Buddhism in Ladakh and influenced neighbouring Buddhist communities.2 And of course it influenced the religious activities of noble monks from the Alchi (A lci) area, whose affiliations with Nyarma has been attested by Philip Denwood through textual evidence documented in the famous, and today still mainly intact, temples of the Alchi compound.3 The early ‘Golden Decades’ were probably followed by a loss of royal patronage and financial support causing a continuous decline. The monastery of Nyarma was eventually abandoned, and the remaining temple ruins and stupas (chortens—mchod rten) are in poor condition today. 1 Rinchen Zangpo has been credited with the translation of a large number of tantric texts as well the planning of no less than 108 temples in the western Himalayan kingdom. So far Nyarma is the only Rinchen Zangpo foundation in Ladakh identified among the list of 108 on secure textual grounds. See Vitali (1999:24). 2 For a discussion of the political history of Ladakh and the difficulties relating to the definition of regional and sub-regional borders see the contribution in this volume by Neil Howard. 3 See Denwood’s translation in Snellgrove and Skorupski (1977:30, 45, 48).

© koninklijke brill nv, leiden, ���4 | doi ��.��63/9789004271807_�07

142

Kozicz

This essay focuses on one such structure which I describe—for reasons that will be explained—as the ‘Chorten with the Secret Chamber’ (CSC). After years of obvious neglect and vandalism, this chorten is not far from the point of collapse, and it is essential to study it closely while this is still possible. The essay analyses its apparently unique architectural structure, together with the iconography of its interior, and compares it with roughly contemporary chortens from elsewhere on the Nyarma site and in the monastic complex of Alchi.

The Setting

Nyarma lies in the very heartland of Ladakh where the Indus valley is broadest. The site of the monastery seems to have been carefully chosen to the east of the Indus on a flat but slightly elevated, barren part of the valley, perhaps in order to save agricultural land. Its topographical situation may be described as in a ‘bay’, as the flanking eastern hills turn towards the Indus to the south and north of the monastery. On the rocky spur to the north there is an ancient castle,4 and a small hermitage was built on the eastern hill. The latter is still named Ensa (dBen sa—‘hermitage’) and according to local tradition it was used as a place for retreat by the Lotsawa himself. Soon after the monastery’s foundation, several temples and a great number of stupas were built beyond the borders of the main compound,5 including to the north of the castle where another desert field extends for about a kilometre. The structures of this northern section included two temples, three rows of 108 stupas and several smaller clusters of stupas.6 In the 1970s David Snellgrove and Tadeusz Skorupski (1977:19) pointed out that ‘the temples are just empty shells today’. All the roofs of the temples were removed long ago and their decorations have faded away during the centuries 4 For a description of the castle see Neil Howard (1989:269–270). 5 It is noteworthy that additional structures were erected outside the compound despite the enormous size of the enclosed area, thereby leaving a large portion free from architectural structures for purposes that are so far unidentified. The original setting inside the enclosure consisted of four temples only, two of which formed a joint spatial unit. See Kozicz (2010:34–39). 6 Among the ruins in the very centre of the field two neighbouring ruins of square floor plan on a shared platform or basis deserve a separate note. Both of the chambers face west and their remaining walls are made of two layers, one internal and one external. As the entrances were originally framed by colour painted frames created of plaster instead of the wooden frames of temple architecture, the hypothesis can be put forward that these are the ruins of comparatively large stupas.

The Chorten with the Secret Chamber near Nyarma

143

Figure 5.1 General view of the Indus Valley from the Nyarma Ensa hermitage. The Nyarma compound is to the left, and the castle is on the hill near the centre. To the north (the right of this picture) lies the desert stupa field stretching as far as Thikse.

since then. However, some murals have survived within small chambers inside a few of the stupas. Soon after Snellgrove and Skorupski visited Nyarma, a survey by Jampa Losang Panglung (1983:286) brought to light some intact murals inside a stupa within the compound, and he dated these to the 15th century CE. He also noted fragments of an older layer underneath the south wall of these murals, but his findings were insufficient to provide clues for a dating of this early phase on stylistic grounds. Only in the northern section of the site beyond the castle have some minor pieces of architecture survived to reflect the original artistic and religious splendour of Nyarma. Here there are four chambers, all housed inside stupa-like structures and decorated with murals comparable to the early painting style of the Alchi group of monuments. Three of these form a single structural unit sharing a common massive base and in an earlier essay I therefore described them as the ‘Stūpa Triad’.7 The fourth decorated monument is the stupa or 7 In the course of field research by the author in 2002, 2004, 2006 and 2009 intact murals were documented inside. The Stupa Triad is located exactly 300m to the east-south-east of the Chorten with the Secret chamber. The term “triad” was introduced because three stupas share a massive base with the second and third stupas in line behind the front one. Their

144

Kozicz

Figures 5.2 and 5.3 Buddha from the ‘Chorten with the Secret Chamber’ (CSC) compared to the depiction of a Buddha from the Vajradhātu Temple of Sumda Chung, one of the sites listed among the so-called Alchi group of monuments and dateable to the 12th century at the latest. The depiction of the robes and the design of the background, lotus blossoms on blue ground underline its stylistic similarity.

chorten on which this article focuses: the Chorten with the Secret Chamber (henceforth CSC). It stands in the very northern part of the large stupa field to the south of the monastery of Thikse (Khri rtse), founded by the Gelugpa (dGe lugs pa) on a prominent hillside in the 15th century. Thus, this stupa stands in a rather isolated position at the very edge of the cluster. Today it is almost completely surrounded by trees, which adds to its sense of separateness.

Space and Construction

As with the three decorated stupas nearby, the CSC’s exterior may once have reflected the type prevailing in the neighbouring area of Shey, i.e. the stupa may have been erected upon a broad base with several projecting levels and groups of steps, including flights of stairs, on each level and each side. Kath Howard (1995:64) classified this type as the ‘Lotsawa’ type of chorten. Unfortunately, the exterior is now rather eroded thus allowing for not even a tentative reconstruction of its original form. conception must not be confused with the concept of the Three Protectors, i.e. the Bodhisattvas Mañjuśrī, Avalokiteśvara and Vajrapāṇi, visualized through three stupas. Each stupa has a tiny chamber above the shared base and is decorated with murals dateable to the 12th century at the latest. For plans and a description of the iconographic lay-out see Kozicz (2007b:61–64).

The Chorten with the Secret Chamber near Nyarma

145

Figures 5.4 and 5.5 The CSC from the east (left) compared with a stupa from the stupa field of Shey.

Like the painted stupa inside the Nyarma compound recorded by Panglung, the CSC may be accessed through a narrow gate where it is impossible to walk upright: one has to crawl on one’s knees. While passing through the entrance tunnel, one may already note that the wall construction of the basement is made of two different materials: brick masonry was basically used for the core of the construction, above which the ‘stupa compartment’ was erected, and this provided an ideal physical foundation for the murals. Interestingly, the builders chose stone masonry for the construction of the west wall and for the massive outer section of the base.8 After passing through this tunnel, which is some two metres long, one reaches a chamber which is 1.80 metres wide and 2.05 metres from front to back. Inside, the walls were once completely decorated with murals starting at a level of approximately 40 cm above the ground. Opposite the entrance is the main wall with the remains of a mural painting of Buddha Śākyamuni that once presided over the ground floor’s iconographic programme. The architectural layout of the ground floor is based on a clear north to south axis, and the spatial order therefore focuses on the facing wall. In this respect it is similar to the plan of a temple but, as will be seen below, definitely differs from the fourfold symmetrical layout of the stupas of the Alchi group of monuments. About 1.80 metres above ground level, two stone slabs project from each of the lateral walls. These slabs have rather awkward and irregular shapes, and show no traces of decoration or even plastering.

8 The mud bricks are remarkably large (approx. 45 × 25 × 10 cm). They conform to the type (40–45 × 20–25 × 10 cm) used for the very early temples of Ladakh, as already noted by Neil Howard (1989:219). The bricks used for the CSC are even larger than most of those in the Nyarma temples. I thank Neil Howard for pointing out the significance of this detail, which points to a very early date for the structure.

146

Kozicz

In several respects, the CSC shows significant similarities to the Nyarma compound stupa documented by Panglung. First, the entrance is of similar narrowness, although the tunnel to the latter’s chamber is only 1m 20 cm deep. Second, the traces of the paintings on the facing wall again centre on a depiction of the Buddha, although it should be kept in mind that these probably date from a later renovation, and therefore do not necessarily display the original iconography. Third, inside the Nyarma stupa stone slabs projecting from the lateral wall serve as brackets for the stone beams that bear the corbelled ceiling above the ground floor’s chamber. Fourth, the chambers of the two stupas are of approximately the same size. However, what differentiates the CSC from the Nyarma stupa is the existence of a second storey above its ground floor chamber. This ‘upper storey’ can be accessed through a small opening from the east thereby causing a 90° turn of the main orientation, since the main wall usually faces the entrance. The floor basically consists of an L-shaped cat-walk along the northern and eastern walls only allowing the central space to expand as far as the former stupa dome. Thus, the overall order of the interior space lacks a clear orientation and causes an asymmetry that is highly unusual in such a monument. The most significant aspect of the CSC’s spatial programme appears in the area of the main wall. There, a large hole reveals a once-sealed chamber, which was hidden behind the central image of the ground level. It is from this formerly sealed chamber that I have derived the name of the chorten. As shown in the floor plan of level 1 (the ground floor), the hidden chamber was situated within the core area of the stupa, i.e. the back wall or facing wall was moved

Figures 5.6 and 5.7 The upper section of the frame of the central Buddha on the main wall of the CSC includes a winged lion inside a stupa structure. Its face design is practically identical to the lion motif of the stone brackets of the stupa of the Nyarma Compound (right), suggesting that the decoration of those brackets was not done during the restoration phase in the 15th century. Instead, it is part of the earlier layer noted by Panglung, and this confirms the early date of the monument.

147

The Chorten with the Secret Chamber near Nyarma

Stone Slabs

Hidden Chamber 0

4M

Figure 5.8 Cross-sections of the CSC (left) and the decorated stupa in the main Nyarma compound (right). The position of the secret chamber of the CSC hidden behind a brick wall opposite the entrance is clearly depicted. By contrast, the stupa in the compound has only one chamber on the ground floor. A further difference is the use of stone for the construction of its base only.

Hidden Chamber

Catwalk Brick Stone

Stone

Ground Floor

0

4M

Upper Floor

Figure 5.9 Floor plans of the CSC.

towards the entrance. Accordingly, the mid-point of the accessible chamber did not correspond to the central, vertical axis of the stupa. So far, no stupa with a comparable spatial concept has been found in Ladakh, although similar sealed chamber-like spaces can also be found in the Chorten or Stupa Temple of the Nyarma compound (see Kozicz 2007a:49–52). In this temple the function of the spaces is unclear but at least their existence can be explained by the geometric principles according to which the building was designed. The double-wall construction that caused the cavities at the Chorten

148

Kozicz

Temple was the result of a geometrical ‘necessity’ arising from the extension of the main niche which had to enshrine the whole stupa. If the assembly area in front of the niches had been erected in accordance with geometric principles, half of the stupa would have been within the niche while the front part would have been part of the assembly area. However, the external form of the shape was designed according to geometric principles. Despite their flanking positions, the resulting cavities might have served as chambers containing important sacred objects such as damaged manuscripts or ritual tools, but they are completely empty today. The lateral cavities or sealed chambers of the Nyarma Stupa Temple recall the plan of the Lotsawa Temple of Nako in Kinnaur, which is datable to roughly the same period. There, the lateral spaces extend as far as the back wall of the niche and the shape of the temple’s external ground plan becomes a square, while internally the conception is based on the duality of the main niche and the assembly area in front. As already outlined by Giuseppe Tucci (1988:26), stupas and mani walls in Tibet have no funerary character, and there is no evidence in the case of the CSC to contradict this statement. For example, there are no traces of any tsatsa (tsha tsha), the small votive stupas made of clay to which the ashes of a deceased person were sometimes added. Whatever was hidden behind the now-broken brick wall will probably remain a mystery. Another question that arises is: how could anyone find out about the existence of the chamber? It would seem likely that it was broken into while the history of the chorten, the circumstances of its erection and its functional purpose were still part of a living tradition. Proceeding towards level 2 (the upper floor) it must be noted that only the western wall rises vertically, whereas the other walls lean inwards. This leaning is mainly caused by the fact that the builders of the stupa had to compensate for the non-centric position of the ground chamber (see Figure 8). Moreover, the L-shaped catwalk situated on the northern and the eastern sections had to be structurally supported. This structural necessity is particularly apparent on the right-hand, the east side. On the north side, the backward shift of the upper storey already creates the basement for the catwalk. This part of the catwalk simply rests on the wall below that separating the ground floor chamber from the hidden chamber. The catwalk is three metres above today’s ground level. The upper chamber has a rectangular floor plan as well, measuring approximately 1.95 × 1.75 metres.9 The average height of the chamber is 1.45 metres. 9 All measurements given here have to be treated with caution because of the severe building irregularities and the difficult conditions under which the measurements had to be taken on the upper floor. The accuracy of the measurements is ± 5 cm.

The Chorten with the Secret Chamber near Nyarma

Figure 5.10

Two sealed chambers flank the apse of the Nyarma Stupa Temple which contains the remains of a stupa today.

Figures 5.11 and 5.12

149

Main wall of the ground floor with remains of the original Buddha in the centre. This has been broken and badly damaged by intruders. The stone wall in the back is clearly distinguishable from the brick masonry of the painted main wall. Right: The dilapidated condition of the plaster brings to light the stone masonry of the west (left) wall.

The result of these geometric relations of the structural components is an irregular agglomeration of spatial elements. Above the upper chamber there is a corbelled roof, i.e. a stepped brick construction that covers the interior space like an inverse pyramid. This is another major difference from the Alchi stupas, where each chamber is covered by a

150

Kozicz

Figures 5.13 and 5.14 View from the ground floor of the CSC towards the corbelled roof. By contrast, the chambers of the Alchi stupas, the mandala ceiling of the Great Entrance Stupa being the most complex specimen, were covered with elaborately decorated lantern ceilings made of wood.

so-called ‘lantern ceiling’ made of wood and the iconographic systems centre on the vertical, central axis. By contrast, the iconographic centres of the Nyarma stupas are clearly found on their facing walls. These completely different spatial concepts are clearly revealed through a comparison of their axonometric sections. While the Alchi Entrance Stupa has a symmetric, absolutely centric concept based on the principle that all three axes—two in the horizontal plane and one in the vertical—share a common centre, the Nyarma stupa reveals an asymmetric concept. The difference in the spatial systems becomes even clearer when the interior spaces are shown as solid forms.10 It is noteworthy that all the other decorated stupas in the Nyarma area, including the one within the compound, mirror the spatial concept and construction principles of the CSC’s ground floor. Each of the decorated Nyarma stupas has only one access i.e. there was no multi-entrance structure comparable to the Alchi Group stupas, and all the Nyarma stupas share the corbelled roof construction as described above.11 Incidentally, the fact that no wood 10 11

For a documentation of the Great Stupa of Alchi see Goepper (1993). In this regional context no contemporaneous corbelled roof constructions outside the Nyarma area have come to the knowledge of the author so far. I am grateful to Neil

The Chorten with the Secret Chamber near Nyarma

Figure 5.15

151

Axonometric comparison of the CSC (left) and the Alchi Entrance Stupa (right). By contrast with the Nyarma stupa, the Alchi Entrance Stupa has four entrances and can be accessed from all directions, meaning that there is no ‘facing wall’. Note that the inner stupa is not depicted in this drawing.

was used in their construction may have helped ensure their survival, since all the wooden materials of the temples were removed, and probably re-used to construct of new temples, thus contributing to the collapse of the original buildings.

Howard for recalling the existence of a corbelled roof inside the ruins of the Nyarma castle (N. Howard 1989:[53] 269). Accordingly it may be of importance to mention that this kind of construction was a prevailing technology of Buddhist temple architecture in Bengal, e.g. at Mainamati in today’s Bangladesh where inside the cruciform temple of Rupban Mura one corbelled roof has survived until today.

152

Kozicz

Figure 5.16



Axonometric comparison of the different spatial forms (CSC to the left, Alchi Entrance Stupa to the right).

Geometric Order of the Iconographic Programme

The fundamental architectural differences between the Alchi stupas and the CSC continue on the decorative and iconographic level. While the Alchi stupas have almost no decorations at the ground level—only the stupa in front of the Alchi Sumtsek (gSum brtsegs) and the stupa in the backyard of the Alchi Lonpo’s residence have a horizontal row of stupas in the upper-most sections of their lower floors—the ground floor of the Nyarma stupa was fully decorated with murals on all four walls. Of the main figures of the facing wall only the mandorla is visible today, since the wall was destroyed, when some intruder broke into the secret chamber. The mandorla is framed by some kind of throne which is surmounted by a stupa-like construction with lateral turrets. In the centre of this architectural frame stands the winged lion. It is depicted frontally with forelegs firmly placed on the horizontal plane above the Buddha’s head. This composite figure is very similar to a Garuda. Both motifs were frequently placed between the deer above the outermost doors of the painted mandalas of the Alchi Group temples. On both sides of the central figure two seated Buddhas in monks’ robes are situated in vertical order. On both of the lateral walls three rows of four Buddhas are depicted, each of them showing different mudras. Their names were written in small rectangular fields or ‘captions’ next to their lotus thrones, but these have almost completely faded away. All four walls of this ground floor level once shared a blue background colour. On the entrance wall, which is the most badly damaged, only one Buddha figure has survived. Above the field of the Buddha figures the curtain motif, which is so well-known from the Alchi temples, serves as a frame of this iconographic section of the interior space. The curtain motive consists of two layers

The Chorten with the Secret Chamber near Nyarma

153

of which only the larger, upper one has kept its blue original colour. The colour of the lower section, which is slightly smaller, has faded away and appears white today. Above the curtain, the usual geese motif creates a symbolic line that sets a clear iconographic demarcation. It separates the programme of the lower chamber from the upper section of the interior. Stylistically, the whole decorative arrangement of the curtain motif and geese exactly compare to the decorative system of the early Alchi temples.12 Interestingly, these geese are not situated just underneath the catwalk of the upper story. Instead, the row of geese is placed right under the stone slabs. Accordingly, there is a distance of approximately a metre between this demarcation line (i.e. the ending of the iconographic programme of the ground floor) and the beginning of the upper chamber. The decorations above the rows of geese have remained intact only on the southern wall and partly on the eastern wall, thus permitting no more than a limited description of the iconography of this intermediate spatial component that—in terms of iconography—obviously belongs neither to the ground floor nor to the upper floor. The topics depicted there are the following (starting right above the geese in vertical order): first is a stripe, which shows only a few flower elements. No background colour can be identified. Next follows a row of banners of a sofar unique type painted in blue and red lines. Aesthetically, it is very close to a vertical lotus motif. This decorative row lacks a background colour, while the lines have been drawn in blue and red. The most interesting aspect of this intermediate section of decoration is the third part, where several stupas are depicted. By contrast with the Alchi stupas, where ornamental motifs made of stupas (repeating one selected type) were widely used, this horizontal stripe in the CSC does not show a repetitive, ornamental stupa motif, but a clearly differentiated visual representation of various stupa types. It is also remarkable that the stupa drawings inside the Nyarma stupa reflect a structural order closely related to the classical Tibetan types of stupa as noted by later scholars and historians of the 14th century.13 All the four surviving stupa images are the same size. They recall the typical Tibetan stupa structure consisting of a plinth, a base, three or four steps, a dome, a harmika and the rings with the finial. They differ only in the n ­ umber

12 13

Interestingly, such combinations of decorative motifs do not occur in any of the chambers of the Triple Stupa. For a comparison of the two versions of the proportional system of Tibetan stupa architecture—by Buston and Desid—see Dorjee (1996:69).

154

Kozicz

Figures 5.17 and 5.18 Upper corner area of the main wall and the adjoining right wall with the decorative elements, namely the curtain motif, the row of geese and a band of triangles. The protruding stone slabs emerge from this section. Right: Decorative banners and the row of chortens depicted above the entrance.

and form of the steps in their middle sections.14 By contrast, all the stupas painted inside the Alchi monuments display a completely different archetype, based on a pyramid-like structure upon a lotus-like base and with an impressively large pile of disks topped by an umbrella to which scarf-like textiles are attached. It is noteworthy that the stupas depicted inside the Nyarma stupa have eyes in their vases or domes (i.e. the egg-form component).15 This is the earliest appearance of this motif, which can be found regularly in later monuments, such as the temples of Tsatsapuri at Alchi or the nearby Saspol cave temples, from the 13th century onwards.16 The ‘eye-motif’ was never drawn on a stupa of the Kashmir type.17 The field containing the depictions of the stupas ends right below the level of the catwalk. The clear boundary between the iconographic programme of the ground floor (level 1) and this intermediate section which is dominated by

14

15

16 17

The Tibetan types of stupa number eight and like the stupas inside the CSC these differ from each other only in the section of the steps. For a detailed description and drawings of the Tibetan types of stupa architecture see Gutschow (2006:198–199, Abb.2). Despite those compositional similarities it must be noted that the number of stupas depicted within the horizontal band inside the CSC did hardly conform to the standard number of eight since there were already three placed above the entrance. Inside the chamber of stupa no.1 of the nearby triple stupa there is another stupa with eyes on the vase. This stupa drawing likewise reflects the Tibetan order of the stupa. See Kozicz (2007b:61–62). Eyes were frequently drawn on the harmikas of stupas in Nepal, but not on the domes. For a sculptural depiction of a stupa type from Kashmir see e.g. Pal (2003:106–107).

155

The Chorten with the Secret Chamber near Nyarma

0

Figure 5.19

10

50 cm

Comparison of stupa types from mural paintings: Left: Prototype of the classical Tibetan type of stupa. Centre and right: Two of the stupas from the CSC compared to the stupa from the centre of the eastern wall of the Alchi Dukhang which represents the Kashmir type of stupa architecture. All three are to scale.

Figures 5.20 and 5.21 Eyes on a dome inside the CSC (left) compared to a complete face on a stupa above Maitreya from Saspol Cave No.2.

the row of stupas shows that this section has to be considered separately—due to its transitional character from now on referred to as ‘level 1.2’. The murals of the upper level, level 2, are in rather poor condition. Since the eastern opening was clumsily enlarged in the recent past, the sun now shines inside. As a consequence, the western and northern walls have suffered severely from the sun’s rays, and the murals are almost gone. In addition, large proportions of all four walls have been covered with whitewash, particularly the southern and the western walls. Still, some of the murals can be identified. On the northern wall a Buddha figure was once flanked by two standing Bodhisattvas. The face of the left-hand Bodhisattva (from the viewer’s perspective) is very similar to the face of the flanking Bodhisattva inside stupa

156

Kozicz

no. 1 of the nearby Triple Stupa. The Bodhisattva is turning towards the central Buddha. He wears a dhoti which covers his knees. Most probably he holds a vajra, today difficult to attest securely, in his left hand right in front of his chest. The decoration of the southern wall is likewise difficult to identify. Interestingly there is no triad depicted there. Instead, a seated Buddha is flanked by another, smaller Buddha (?) figure, also seated. It is noteworthy that the left hand of the minor figure, performing the mudra of granting, extends into the mandorla of the larger Buddha. The larger Buddha is placed on a lower level than the accompanying figure. This higher position can be explained by the asymmetry of the architectural structure caused by the catwalk. On the entrance wall only the mural on the left of the opening—when facing it from inside—has survived. It shows a stupa almost identical to the stupa on the entrance wall of the nearby Stupa no.1 of the Triple Stupa. Again, eyes are depicted on the vase. It is noteworthy that the upper decorative section is similar to that of the ground chamber (level 1), but here the curtain is red only. The geese are outlined in red as well. As can be clearly seen from the elevation drawing, the structural order of the iconographic programme is completely different from the clear layout of the murals on the ground floor.

+300.00 Level 2

+200.00 Level 1.2

0.00 Level 1 0

Figure 5.22 Figure 5.23

4M

Complete interior elevations of the main wall (north wall) and the opposite entrance wall (south wall). Bodhisattva of the upper storey (south wall- left of the Buddha). Today, due to its low angle, the morning sunlight penetrates the chamber through the enlarged opening and is gradually destroying the murals.

The Chorten with the Secret Chamber near Nyarma



157

Hypotheses on the Purpose of the Stone Slabs

The projecting stone slabs are clearly part of level 1.2, and may be related to the murals of the stupas there. There is no firm evidence, but the Alchi Stupas may give some possible indication of their purpose. The Alchi Stupas in effect serve as shrines for votive ‘inner stupas’, which were placed in the centre of their inner chambers. The heights of these inner sanctums range from approximately 2.00 to 2.50 metres, and any similar votive stupa in Nyarma would have had to have been smaller because there is less space. If it had been any taller than 80 cm it would have extended into the upper storey of the catwalk level— from now on referred to as level 2—and clashed with the spatial limits of that level’s iconographic order. However, the construction of a stupa resting on the stone slabs would have necessitated at least two large cross beams. So far it has not been possible to identify traces of any stone beams, even among the debris on the ground. It is also possible that there was no single object in the centre but rather four votive stupas, one on each of the slabs. It is also possible that stupas may have been placed on the slabs temporarily for the duration of a particular ritual related to the objects sealed in the chamber. In this context it is worth recalling that a remarkable number of wooden votive stupas dated to the 10th to 12th centuries, the original contexts of which have not been discovered, have been found in Ladakh during recent years.18 The fact that the slabs were never plastered and painted rather contradicts the idea they might have served a ritual purpose. An alternative hypothesis suggests a change of the architectural and iconographical programme during the process of construction. It may be that the builders changed the conception of the monument after completing the ground floor i.e. when the slabs were already firmly connected to the structure of the walls and could no longer be removed. As the cross-section reveals, the back wall of only 20cm could hardly have sustained a dome construction similar to the Nyarma compound stupa. It is possible that the builders failed to take this into account at the outset, and were then forced to change the architectural concept. That would also explain the inconsistencies within the overall structure and the irregular order of the iconographic system. The implication is that the interior shape and spatial order are not the result of a plan laid out right from the beginning but rather that the design was developed and adjusted spontaneously in the course of construction in response to changing paradigms and functional purposes. It will be difficult to confirm any of these hypotheses unless further comparable structures come to light. 18

Snellgrove and Skorupski (1977:43, Figure 26) were the first to document a wooden votive stupa in Ladakh.

158

Kozicz

Figures 5.24 and 5.25 Inner stupa from the shrine stupa approximately 50 m from the Alchi compound (today in the back yard of the residence of the Alchi Lonpo). Note: This inner stupa has a face painted on its dome. Right: A wooden votive stupa from Saspotse.

Summary Comparisons of building materials used, and comparisons on stylistic grounds, provide clues for a dating of the Nyarma stupas to the 12th, perhaps even the late 11th century. Despite the stylistic similarities with the Alchi monuments, the order of the iconographic programmes and the applied construction technologies point at two different traditions of stupa architecture. The significance of the CSC is revealed through the division of its interior space into three sections and the unique interaction of these sub-spaces and their decorative patterns. The non-symmetric order and the interaction of the main axes of the various sub-spaces create an atmosphere which is completely different from that of the Alchi stupas. Its interior architecture represents a vertical arrangement of three separate spaces. Thus, the CSC is among the earliest surviving multi-storey structures with a multi-level iconographic system in the region, thereby testifying to the creative momentum of the Second Diffusion of Buddhism and its impact on the religious architecture of Ladakh. Acknowledgements My participation in the IALS conference at Rome 2007 including the work on this article has been generously supported by the Austrian Science Fund (FWF Projects P18336-G06 and P21139-G21). I am very grateful to John Bray and Neil Howard for their helpful and encouraging suggestions, comments and corrections.

The Dating of the Sumtsek Temple at Alchi Philip Denwood In 1980 I published most of the mural inscriptions in the Dukhang (’Du khang) and Sumtsek (gSum brtsegs) temples at Alchi (A lchi), Ladakh, and numbered them 1–12 (Denwood 1980:152). I concluded that the oldest temple on the site, the Dukhang, was founded in the early to mid-11th century. I further proposed, mainly on grounds of the palaeography, style and authorship of its inscriptions, that the three-storey Sumtsek temple alongside the Dukhang was founded (by Tshul khrims ’od) in the last third of the 11th century (Denwood 1980:152). Further important inscriptions from the top storey of the Sumtsek were subsequently published by Roger Goepper with Jaroslav Poncar (Figure 6.1) (Goepper (1996a:212; 216–7). See also Goepper 1990:159–176). The Sumtsek’s founder Tshul khrims ’od is mentioned in one of these (but see below), alongside paintings of a lineage of Kagyupa (bKa’ brgyud pa) lamas, the last of whom is ’Jig rten mgon po, the founder of the Drigungpa (’Bri gung pa) order (Figure 6.2). Along with art-historical arguments, this led Goepper to propose that the Sumtsek’s founder “must have been active around the end of the twelfth to the beginning of the thirteenth century”; thus over a hundred

Figure 6.1 Murals on the top storey of the Sumtsek temple, with members of a Kagyupa lineage in the left-hand panel, and rows of monks on the right-hand panel. Photo: Lionel Fournier.

© koninklijke brill nv, leiden, ���4 | doi ��.��63/9789004271807_�08

160

Denwood

Figure 6.2 Left-hand panel showing the Kagyupa lineage with accompanying dbu can inscriptions. Photo: Lionel Fournier.

years later than my dating. Goepper has continued to promote this hypothesis despite counter-arguments from, among others, Heather Stoddard and Lionel Fournier (Stoddard 2003:167; Segraves 2001). I feel it is worth restating and amplifying my own hypothesis in view of Goepper’s evidence and arguments. There are several places in the Sumtsek where inscriptions appear to have been obliterated (Fig. 6.3) and, in some cases, re-inscribed. Inscription 6, to the proper right of the statue of Maitreya, is written in an dbu med style like that of Inscription 2 in the Dukhang by the same author (a point to which I shall return below). The corresponding panel to Maitreya’s proper left has been overwritten in dbu can script and rather uneven lines in the reign of Tashi Namgyal (bKra shis rnam rgyal) in the 16th century (Inscription 10: Figure 6.4) (Denwood 1980:140–41; 149). There could presumably be two reasons (not

The Dating of the Sumtsek Temple at Alchi

Figure 6.3 Panel in the Sumtsek murals with obliterated inscription. Photo: Lionel Fournier.

Figure 6.4 16th century inscription in the Sumtsek (Denwood 1980: Inscription 10). Photo: Lionel Fournier.

161

162

Denwood

Figure 6.5 Rows of monks with obliterated inscriptions. Photo: Lionel Fournier.

mutually exclusive) for this: a) disapproval of the contents of the original; and b) a feeling that some later message was more important. Both obliteration and overwriting can, I claim, be seen on the second storey panels. On both the left (Kagyupa lineage: Figure 6.2) and the right (rows of monks: Figure 6.5) clear traces of former inscriptions can be seen even in Poncar’s photos (Goepper & Poncar 1996a:216–7). On the left the images are now accompanied by crude dbu can inscriptions with uneven lines; on the right the panels have not been reinscribed, leaving the figures anonymous, and so no threat to anyone. I concede that the lineage figures on the left do indeed represent the Kagyupa lineage down to ‘Jig rten mgon po. If the figures were original at the foundation of the temple and correctly identified in the original inscriptions, then why obliterate the latter and re-inscribe? The obvious conclusion is that the figures, like their present accompanying inscriptions, are not original.

The Dating of the Sumtsek Temple at Alchi

163

Is there any supporting evidence for this? I think there is. The Paintings The symmetry observed in the smaller figures of monks at the bottom, and in the monks on the right-hand panel (all facing in towards the doorway), has been destroyed, as the figures in the right-hand column of the left-hand panel (Fig. 6.2) are facing outwards. This panel now possesses its own internal form of symmetry. ii) Christian Luczanits (2003:28) has observed that the depictions in the lefthand panel “differ considerably from comparable portrayals at Alchi.” He also refers twice to what he calls their “unusually clumsy” execution. In this he includes the depiction of their clothing—“particularly the awkwardly drawn cape placed flat behind the body forming two pointed ends at the sides (as if attempting to represent one cape placed above another). Capes like this are found neither on any comparable painting of this lineage nor anywhere else at Alchi”. iii) The facial type—a white face, with the further cheek protruding beyond the nose, is not seen elsewhere at Alchi. iv) The bright blue tint of the background is similar to that in other acknowledged later additions in the Sumtsek; although this could perhaps be just an artefact of the photography. v) Goepper’s arguments about similarity of painting style in the second storey to the rest of the Sumtsek relate mostly to the parts of the wall which were not repainted, not to the lineage figures (Goepper 2003:15–24). a) i)

The Inscriptions The inscriptions in the lineage panel are in printing-style dbu can instead of cursive dbu med script like all other inscriptions in the Dukhang and Sumtsek except the 16th century Inscription 10. In both these dbu can inscriptions the execution is clumsy with wavering lines which sometimes wander outside the boundaries of the panels. ii) The syllable dge is spelt without the archaic final a chung which it consistently has in the dbu med inscriptions. iii) In the reference to the Sumtsek’s founder Tshul khrims ‘od (Figure 6.6) his name (if that is what it is: see below) has evidently been written first without the final syllable, which has been added later above the line. Moreover this syllable is in a form spelt not only wrongly (ba’od), but in a form hardly even possible within the rules of Tibetan orthography. In my experience the one thing any reader of a document can be guaranteed to spot is a misspelling of their own name. Would Tshul khrims ’od, b) i)

164

Denwood

Figure 6.6 Inscription including the name of Tshul khrims ’od. Photo: Lionel Fournier.

patron of the beautiful paintings and superbly executed inscriptions elsewhere in the temple, have countenanced such a botched job? This seems unlikely.1 vi) There are two inscriptions at Alchi composed by Grags ldan ’od—one in the Dukhang (Inscription 2) (Denwood 1980:27–8; 144–5) eulogizing sKal ldan shes rab, that temple’s founder, and the other in the Sumtsek (Inscription 6) (Denwood 1980:135–7; 147), eulogizing Tshul khrims ’od, both written in dbu med script. It is doubtful whether these could have been written more than a few years apart; certainly not the 150 years or so between my dating of the Dukhang and Goepper’s dating of the Sumtsek. c) Iconography The iconographic program of the temple is pre-Drigungpa. Whereas that order placed much emphasis on cycles of the Anuttarayogatantra, as Heather Stoddard has written, “it is the Yogatantra mandala that dominate [sic] the earlier period of Alchi”, and she adds in a footnote, “This of course goes against Prof. Goepper’s conclusion that A lci was founded by the ’Bri gung pa” (Stoddard 2003:67). Lionel Fournier has independently made similar observations (Segraves 2001). My conclusion is that the paintings and inscriptions of the left hand panel are a crude later forgery, designed to claim a Drigungpa inspiration behind the foundation of the Sumtsek.

1 It is even possible that the original passage bdag tshul khrims bgyi ba without the later interlinear addition just means “I, holder of monastic vows”, and did not refer to Tshul khrims ’od at all. I owe this idea to Amy Heller (personal communication).

The Dating of the Sumtsek Temple at Alchi



165

When Might This Have Happened? I Can Think of Three Possible Times

The spread of the Drigungpa order of Tibetan Buddhism into the western regions of the Tibetan world began in 1191 when ’Jig rten mgon po, founder of the order based in Central Tibet dispatched a group of monks to establish a hermitage in the Mount Kailash region (Vitali 1996:372). This was followed by a further expedition in 1208, and then by another in 1214–15 which according to the ’Bri gung ti se lo rgyus was patronised by the king of Purang and his son, and by dNgos grub mgon, king of Mar yul (i.e. some part of the Indus valley now included in Ladakh) (Vitali 1996:380). dNgos grub mgon seems to be the king of that name mentioned in the La dvags rgyal rabs as patronising the Drigungpas. Precisely what territory within Mar yul he ruled is not known, but it could well have included Alchi. In parallel with my research on all the inscriptions in Sumstek and Dukhang, including the inscriptional studies of the Drigung portraits, since 2007 Amy Heller’s historical research had led her to suggest the idea of Drigungpa reconsecration due to the historic parallels in Dolpo, Limi and Khojarnath in the first two to three decades of the 13th century (Heller forthcoming). The Three-storeyed temple at Wanla (Wan la), some 20 km west of Alchi, is similar in architectural style to the Alchi Sumtsek, although on the evidence of the woodwork and painting styles probably somewhat later. Roberto Vitali has tentatively dated it on art-historical grounds to “about 1240” (Vitali 1996:387–8). We learn from the inscription in this Wanla temple that its founder and local ruler ’Bhag dar skyabs had already asserted or reasserted Wanla’s control over Alchi; also that one of his sons had studied at Drigung; and that the title of khri dpon had been conferred on him (presumably by some Mongol ruler) after he had visited Kashmir (Tropper (2007): 116; 129). The likeliest times for him to have received this title would have been “the few years following 1235 and the few following 1252” (Howard 2008). Thus it may have been under these new rulers of Alchi in the mid-13th century that the “conversion” of Alchi to the Drigungpas and the painting of the top-storey lineage mural took place. In personal communication, Neil Howard has suggested an alternative possibility: My tentative interpretation of the contents of Inscription 10 is that King Tashi Namgyal was born in either a house at Alchi or in the monastery which is referred to figuratively as the local ‘bri gum pa [sic] monastery.

166

Denwood

If Petech’s date for his enthronement is correct: c. 1555, and he might have been born 20–25 years earlier, then the latter event took place in 1530–35—i.e. the time of Mirza Haidar’s invasion. At any rate, during the dark period when evidence suggests that Hor raids may have been frequent, Alchi would have been a safer place than most in Ladakh for a royal/noble birth. Tashi Namgyal’s father according to sources was named Bhagan, which must be true since when the La dwags rgyal rabs was begun about 100 years later there would have been many people alive whose fathers and grandfathers would have been well aware who Tashi Namgyal’s father was. There is a character named Baghan, a Jo of Ladakh, in Mirza Haidar2 who seems to have operated in lower Ladakh. Tashi Namgyal’s ancestry is from Lower Ladakh in the La dwags rgyal rabs. His family lived either in Basgo, or Temisgam or in Saspola from where they retreated to Alchi. Tashi Namgyal was a patron of the ‘Bri gung pa (Petech 1977:29–30) Out of all this came a local desire to restore the Alchi Chos khor and this pious fraud to enhance the Sumtsek’s ‘Bri gung pa credentials. Inscription 10 in the Sumtsek (Denwood 1980:152) was clearly written in the reign of Tashi Namgyal and dedicated to him. It is in a similar dbu can script to the top storey inscriptions with similarly poor spelling, and probably written over a blanked-out previous inscription. The connection with Drigung is clear, assuming that this is what “’Bri gum” refers to. These characteristics would be consistent with the hypothesis that Inscription 10 and the top story inscriptions are contemporary. 2 “Mirza Haidar” refers to Elias and Ross (1895): 443.

The Iconography and the Historical Context of the Drinking Scene in the Dukhang at Alchi, Ladakh1 Marjo Alafouzo The Dukhang (’Du khang) or Assembly Hall at the monastic complex of Alchi (A lchi) (Map 7.1) is dated to c. 1050 CE by inscriptional evidence and is thus the earliest temple on the site (Denwood 1980:151–52). The inscriptions are, however, largely religious in content and contain very little historical information

Map 7.1

Alchi, Mangyu, Tabo, Tholing and Dungkar (map not to scale) (Drawing: author).

1 This paper is based on an unpublished Ph.D. dissertation, The Iconography of the Drinking Scene in the Dukhang at Alchi, Ladakh, University of London, 2008. The author would like to acknowledge the help of Philip Denwood (Reader Emeritus, University of London), Dr Amy Heller, Neil Howard and Lionel Fournier during her doctoral research. Thanks are also due to the staff at The Western Himalaya Archive in the University of Vienna (WHAV), Austria, especially to Dr Verena Widorn and Verena Ziegler, for helping with image requests.

© koninklijke brill nv, leiden, ���4 | doi ��.��63/9789004271807_�09

168

Alafouzo

regarding the founding of the Dukhang. Furthermore, the written sources for Ladakh’s early medieval history (10th–12th centuries) are particularly fragmentary, with no certain dates or names for rulers, and can thus at best offer mere glimpses of possible historical scenarios in the region. Therefore the main approach in attempting to determine the Dukhang’s cultural background is through its secular murals, of which the best known is The Royal Drinking Scene as it has been published and subjected to art-historical discussion by several scholars. David Snellgrove and Tadeusz Skorupski did the pioneering work on the Alchi temples, including the Dukhang and The Royal Drinking Scene and their results were published in The Cultural Heritage of Ladakh (1977), Volume 1.2 The art-historical conclusions reached by scholars are, however, somewhat ambiguous and fail to determine the origin of the mural’s iconography. In the following, the drinking scene will be subjected to comparative iconographical analysis, together with available historical information to suggest an ethnic origin for the mural’s participants in historical context.

Figure 7.1 Mural painting The Royal Drinking Scene, c. mid-11th century, the Dukhang, Alchi. (Photograph courtesy: Jaroslav Poncar 1981, WHAV JP81.8.4.16). 2 David Snellgrove discussed the mural in Snellgrove and Skorupski (1977:31, 33). He also suggested the title for the mural. See also Pal and Fournier (1982). The most recent research on the mural is by Flood (2005) and Papa-Kalantari (2007).

royal Drinking Scene, Alchi

169

The Royal Drinking Scene (Fig. 7.1) is a foundation scene, painted at the time when the temple was built. The mural depicts a woman offering a drinking vessel to the centrally seated man. To his right, a kneeling man holds a cup, and standing subsidiary men and women flank the three central figures. Men carrying weapons and preparing drinks complete the lower part of the scene. Although the mural has been described as painted in a Yarlung dynasty (c. 630–850 CE) context rather than reflecting a political situation of 10th–11th century western Tibet (Snellgrove and Skorupski 1977:33), its classification as a foundation scene strongly suggests the artist was portraying actual people who were the patrons of the Dukhang. The representation of lay people amid purely Buddhist murals attests to their importance and thus to their royal status, and it may therefore be assumed that the mural is depicting an 11th century Ladakhi court. The secular subject matter of a cup offering is also extremely rare in surviving Tibetan temples, which normally have portrayals according to the Buddhist pantheon. A very similar mural of a woman offering a cup to a man, however, appears in a nearby, c.11th century Assembly Hall dedicated to Vairocana at Mangyu (Mang rgyu) (Fig. 7.2; and see also Map 7.1).3

Figure 7.2 Mural painting, 11th century, Mangyu, Ladakh. (Photograph courtesy: Christian Luczanits 1994, WHAV CL9439;2).

3 An early 16th century image of King Tashi Namgyal (bKra shis rnam rgyal) seated under a parasol and holding a long stemmed cup is portrayed in the Gonkhang (mGon khang) temple, near the Namgyal (rNam rgyal) Peak castle in Leh (see Snellgrove and Skorupski, 1977:99, 104). Historically and iconographically, however, the mural is different from the Dukhang scene.

170

Alafouzo

Therefore it seems that the iconography of the cup offering is unique to 11th century Ladakh, and this geographically limited artistic representation suggests a specific politico-historical situation in the region.

Textual Evidence for Cup Rites and Cupbearers

There is very little written evidence for cup rites and female cupbearers in Tibetan historical sources. The Tibetan legendary epic Gesar (Ge sar/ Ke sar) refers to a cup rite on the occasion of the enthronement of Gesar, when his fiancée ’Brug mo offered him an auspicious long-stemmed vessel full of pure nectar (Stein 1956:137–38). The epic was compiled in its present form in relatively modern times, and although its events and the personalities are not thought themselves to be historical, it is likely they hark back to a genuine historical situation in the 11th century (Stein 1959:230 ff.).4 Several references to women cupbearers and cup offerings can be found in the Turko-Islamic written sources, which reflect pre-Islamic Turkic traditions in the steppe.5 Dede Korkut is the epic of the Turkic tribal confederation of the Oghuz, which was compiled in the early 15th century but is based on much earlier 8th–11th century Turkic oral traditions from the steppe. The epic describes several events where drinking formed part of the festivities, such as banquets that took place before or after the hunt, and marriages (Lewis 1974:18).6 Female cupbearers, who were the captured women of the enemy, are mentioned in the role of serving drinks to the Oghuz noblemen during feasts (Lewis 1974:42, 88). It is worth noting that Gesar was in close proximity to Turkic populations, and his fiancée ’Brug mo is thought to have come from a region influenced or controlled by the Turks, who included the Uighurs of Ganzhou (Stein 1959:194, 196). This suggests that the cup rite reflects Turkic traditions, and entered Tibetan literature via the Gesar epic. The Turko-Islamic Qarakhanids (c. 992–1212) ruled a vast area, stretching from Transoxiana to Xinjiang (Golden 1992:222). Their customs are discussed in the Kutadgu Bilig, written in 1069 CE in Kashgar, which includes ceremonies where drinks were offered by male cupbearers (Dankoff 1983:135–36).7 4 The Ladakhi version of the Gesar epic has been shown to be a later import from eastern Tibet and is therefore irrelevant to 11th century Ladakh. 5 Esin (1969) discusses the cup rites in Central Asian Turkic context. 6 During the 9th to 11th centuries, the Oghuz confederation of nomadic Turks lived in western Central Asia and Kazakhstan (Agajanov 1998:66). 7 The Kutadgu Bilig was written for princes as advice on how to rule.

royal Drinking Scene, Alchi

171

Feasts took place at weddings, funerals and at events of granting a title or rank, and thus were similar in character to the Oghuz ceremonies (Dankoff 1983:189–90). When compared to the Oghuz tradition, however, the attitude of the Qarakhanids at least in formal court situations seemed to have been more restrained towards drinking. This was possibly due to the Qarakhanids’ being amongst the first Turkic tribes to convert to Islam, which might have encouraged their upper echelons to practise moderation in the consumption of alcoholic substances although, significantly, the cup rites were not banished altogether.8 The above textual evidence suggests the tradition of cup rites was established well before the Turks’ conversion to Islam, and originated in the nomadic customs of the Inner Asian steppe.9 The 13th–14th century written descriptions of the Mongol court ceremonies attest to drinking rites, which were performed according to a strict protocol and hierarchy, for example, in the order of drinking and the way the khan and his immediate family were positioned in court (DeWeese 1994:221–25). The textual sources also describe the Mongol enthronement ceremony, part of which was the elevation stage, where the khan was placed upon a felt rug and lifted off the ground by male court attendants (Sela 2003:1).10 In the final stage of the enthronement, the newly elected ruler is seated on the throne, and he is offered a cup of wine.11 The image that is most commonly portrayed in visual representation is the final stage rather than the action preceding it (Sela 2003:56). The above discussion illustrates the strength of ancient nomadic customs, which changed little in their basic functions over the centuries and continued to be performed in new socio-religious surroundings.

The Image of the Ruler Holding a Cup

S­cholars have concluded that the iconography in The Royal Drinking Scene is foreign, with hybrid Central Asian and Iranian influences (Snellgrove and Skorupski 1977:31, 33). More recently and specifically, it has been postulated 8 9 10

11

For a discussion on Qarakhanid iconography, see Karev (2005). Inner Asia throughout this article refers to Mongolia, Kazakhstan, Uzbekistan, Turk­ menistan, Tajikistan, Kyrgyzstan and Xinjiang. Sela (2003:25, 55) notes that although the Muslim authors traced the elevation ritual back to the Mongol conquest in the 13th century, there is a much earlier description of the ritual from 6th century Inner Asia. For example, in the 14th century inauguration of Haishan, the grandson of Qubilai Khan was given a goblet of wine once he was seated on the throne (Sela 2003:31–32).

172

Alafouzo

Figure 7.3 Gold medallion, 11th–12th century, Iran. (Photograph courtesy: Freer Gallery of Art, Smithsonian Institute, Washington DC. F 1943. 8).

that the cup offering in the Dukhang is derived from Islamic iconography (Flood 2005). The Islamic image typically portrays a seated ruler in a full frontal position, holding a cup in his hand and thus subscribing to the final stage of the enthronement ritual (Fig. 7.3). The ruler can also be depicted being offered a cup by one of the two flanking male attendants. The portrayal of the enthroned ruler appears frequently in several different media, including metalwork, ceramics and manuscripts from the 9th century onwards in the lands under various Islamic and Turko-Islamic dynasties, and thus encompassing a vast geographical area that included today’s Iran, Iraq and North Africa. The cup-holding image appeared also in the pre-Islamic Turkic funerary monuments of the Inner Asian steppe, and many of these stone statues depict male figures holding a cup in front of them (Fig. 7.4). In the 8th century Kök Turk stone stele from Mongolia the Turk leader Bilge Qaghan and his two attendants are shown holding drinking vessels. The Qaghan is portrayed frontally, seated with his legs folded and ankles crossed in the Turkic position known as bagdas (Esin 1969:232).12 12

According to Esin, this position formed the prototype for representing the monarch in Turkic art. For an illustration of the Bilge Qaghan stele, see Sinor and Klyashtorny (1996:342, Fig. 3).

royal Drinking Scene, Alchi

173

Figure 7.4 Stone statue, Turkic balbal with drinking cup from Lake Issyk-kul. (Photograph courtesy: Hilary Smith).

The Dukhang mural, however, does not show us the established image of the ruler holding a cup, but instead the central male figure is being offered a cup by the woman to his left (Fig. 7.5). Therefore the iconographic significance of

Figure 7.5 Detail of The Royal Drinking Scene. (Photograph courtesy: Jaroslav Poncar 1981, WHAV JP81. 8. 4. 17).

174

Alafouzo

the scene focuses on the female cupbearer. The absence of the cup in the hand of the central man is further emphasised by the man to his right, who does hold a cup. It seems therefore that the iconographic parallels for The Royal Drinking Scene have to be searched for beyond Islamic representations of a ruler, as they differ substantially from the Dukhang image. Furthermore, cup offerings by women in a secular context are hardly ever shown in surviving works of art. The strongest documentation for female cupbearers is in TurkoIslamic literary sources, which hark back to earlier pre-Islamic steppe customs. Turkic literary evidence will be further examined and applied in the following iconographic analysis.

The Iconography of the Drinking Scene

The Dukhang mural portrays the participants in a clear hierarchical order, with the men to the right and the women to the left of the central male figure, who, seated at the highest level, occupies the dominant position in the Dukhang composition (see Figs. 7.1 and 7.5). However, the man is not portrayed frontally but in a three-quarter view, which together with the absence of a cup contradicts the position reserved for a sovereign. The central female is seated only slightly lower than him, which suggests her status is nearly equal to his. In Mongol protocol, such a position would have been reserved for the khan’s wife (Jackson 1990:211). Among the most visible features of both the man and the woman are their haloes, which in Tibetan painting are depicted not only on Buddhist deities and important monks but also on secular people thought to be worthy of the divine (Singer 1997:128). The attendant next to the central man is kneeling on both knees, in an act of deep homage and the raised cup in his hand suggests he is about to perform a cup rite in honour of the couple.13 His lack of a halo further marks him out as a non-royal personage. The palms of the woman’s hands are painted red, a characteristic that can be found depicted in Tibetan paintings of both male and female Buddhist deities and of high-ranking lamas. The rarity of secular women in extant earlier Tibetan painting makes it difficult to draw any firm conclusions about whether painted hands were a commonly shown feature for women. The female, nonOghuz cupbearers in Dede Korkut are described as having their hands dyed with henna from the wrists down (Lewis 1974:42, 88). The epic also refers to the hennaed hands of the Oghuz prince’s daughters and daughters-in-law, who 13

Esin (1970–71:25) notes that the act of raising one’s cup to the chest before drinking denotes homage and the vow of fidelity.

royal Drinking Scene, Alchi

175

were participating in a marriage feast (Lewis 1974:69). There could be two possibilities for the woman’s red coloured palms in the Dukhang. Firstly, the red colour might signify the female as a devout Buddhist of royal rank and thereby could have a transcendental meaning. Alternatively, if her red palms have a profane rather than a sacred implication, then it is possible she is portrayed as a bride in the mural. In Dede Korkut both the captured non-Oghuz women and the aristocratic Oghuz women themselves have hennaed hands, suggesting the tradition could have originated amongst the Turkic people from the steppe. The Turko-Islamic textual sources attest to cup offerings by women on the occasions of feasts and marriages. The act of the cup offering in the Dukhang mural is performed strictly between the centrally seated man and woman while the onlookers are either serving or cheering the couple. The scene could thus be portraying a marriage ceremony, where the woman is showing allegiance to her new husband by offering him a cup. It must again be emphasised that marriage ceremonies or drinking rites are not depicted in extant Tibetan temples outside Ladakh, which suggests the scenes in the Dukhang at Alchi and Mangyu were painted to commemorate a historically and politically important event in the region. The lower scene of the mural depicts armed men, whose weaponry implies they are soldiers (Fig. 7.6). The centrally seated man’s supreme rank amongst

Figure 7.6 Detail of the lower scene of The Royal Drinking Scene. (Photograph courtesy: Jaraoslav Poncar 1981, WHAV JP81.8. 4. 18).

176

Alafouzo

these men is further denoted by his axe and dagger. In the foreground there are five narrow-necked drinking vessels. Two of the men are kneeling on both knees, and one of them holds in his hands a flask and what appears to be a long piece of fabric, which resembles the scarf hanging down his back. The flask designates the man in the role of the cupbearer, and thus the fabric he is carrying could be a ceremonial scarf that was used in connection with the cup rite. In the Gesar epic, the bride of Gesar offers him a drinking cup and scarves on his enthronement (Stein 1956:137–38). In more modern times, the marriage customs amongst the nomads in Rudok (Rut hog) involve a reception ceremony for the gnya’ bo (the bridegroom’s singer), who is received by a girl holding a kettle full of ‘white wine’ (apparently chang, ‘beer’) that is adorned with a scarf (Shastri 1992:758–59). The underlying theme in the above ceremony is reminiscent of the Dukhang scene, and thus modern day marriage customs in Rudok might derive from ancient traditions that have been adapted over the centuries. There is no definite historical evidence as to the identity of the participants in the mural. The two central men have been described as not Tibetan but of “Indo-Aryan type” (Pal and Fournier 1982:17).14 Snellgrove, relying on the list of Ladakhi kings compiled by A.H. Francke from the La dvags rgyal rabs, the Ladakhi royal chronicle, has postulated that the central man could be the anonymous father of king Utpala (Snellgrove and Skorupski 1977:81). The chronicle, however, is considered historically unreliable and therefore may not be depended on for accurate documentation. Despite this lack of written sources for early medieval Ladakh, it is still possible to suggest a cultural background for the characters of The Royal Drinking Scene by analysing their hairstyles, costume and physical features.

The Women’s Dress and Hairstyle

The central woman in the Dukhang wears a long brown robe with a square neckline, covered by a white cape. The dress is similar to today’s main female costume in Ladakh, called sul ma, which seemingly has retained its basic form for generations (Ahmed 2002:111). The shape of the Dukhang attire differs visibly from the 11th century Tibetan women’s outfits at Tabo in Spiti, Himachal Pradesh (see map 7.1). The earliest representation of female costume at Tabo is in the Entry Hall (dated to 996 CE), and shows women wearing roundnecked plain robes underneath long capes (Fig. 7.7). The capes are also plain 14

Snellgrove suggests that the central figure is Tibetan (Snellgrove and Skorupski 1977:31).

royal Drinking Scene, Alchi

177

Figure 7.7 Mural, late-10th century, Entry Hall, Tabo, Spiti, Himachal Pradesh, India. (Photograph courtesy: Christian Luczanits 1994, WHAV CL9475, 33).

but have narrow, colourful trimmings on the inner edges. Later female attire depicted in the Assembly Hall at Tabo (dated to c. 1042) consists of very ample, white capes with blue trimmings and wide sleeves (Fig. 7.8). A short shawl in

Figure 7.8 Mural depicting the Pilgrimage of Sudhana, mid-11th century, Assembly Hall, Tabo. (Photograph courtesy: Christian Luczanits 1991, WHAV CL9130, 20A).

178

Alafouzo

white, brown and blue covers the upper part of the cape. The few examples of women’s clothing from the ancient Guge kingdom, western Tibet include v-necked dresses, covered by collarless red or white capes with contrasting trimmings on the edges.15 The women at Alchi and Tabo are shown wearing short thick boots, the shape of which strongly resembles those worn in modern Ladakh. In fact, this type of footwear has continued to be worn across Tibet. The longevity of the design suggests that it was dictated by practical considerations such as the harsh climate and the availability of materials.16 The hair of the centrally seated woman in the Dukhang is depicted in multiple thin plaits, which are entwined with strands of blue jewellery (see Fig. 7.5), and similar strings of jewellery are also visible around her neck. The three women near her have identical hairstyles and jewellery. The blue stones are likely to be turquoises, which even today are highly valued and frequently worn by Tibetans and Ladakhis. The central female also has a big blue ornament placed on her forehead, which resembles the arrangement at the front of the perak (pe rag) headdresses in modern Ladakh (Clarke 2004:21). The women depicted in the Assembly Hall at Tabo also have plaited hairstyles and blue stones attached to their hair, and display a blue ornament on their foreheads (see Fig. 7.8).17 The tradition of plaited haired entwined with turquoises still continues in today’s eastern Tibet (Clarke 2004:91, Fig. 7.9). The women in the Dukhang mural can thus be deemed Tibetan by their costume, hairstyle and jewellery. Despite its strong regional character, the Dukhang dress has the cape in common with western Tibetan female costume. We can thus surmise that the cape was a cross-regional part of Tibetan ­women’s dress, and furthermore, that their attire, including hair ornaments, jewellery and footwear, shows historically remarkable unity and continuity.

15

16 17

Female donors wearing this type of dress are depicted on the 11th century north-western stupa at Tholing, illustrated in Jiang, Cheng’an, and Wenlei Zheng 2000:237, plate 142. See also Heller 2011. At the Buddhist cave site of Dungkar, north-east of Tholing, the woman’s upper shawl resembles very closely the one depicted at Tabo in Fig. 7.8. Ahmed (2002:109) has noted that the sole of the modern boots is made of yak leather and the upper part from coloured woollen cloth. Similar ornamentation is depicted on the female donors on the north-western Tholing stupa and on the woman portrayed in Dungkar.

royal Drinking Scene, Alchi



179

Men’s Physiognomy, Costume and Hairstyle

The central man in The Royal Drinking Scene has been described as a foreigner, whose physiognomic details include dark brown skin colour, a moustache and goatee (Fig. 7.9). In contrast, the Tibetan woman in the centre of the mural

Figure 7.9 Detail of The Royal Drinking Scene. (Photograph courtesy: Jaroslav Poncar 1983, WHAV JP83.8.4.20).

has a light, almost luminous skin tone. Thus the artist might have applied dark skin colour to denote a non-Tibetan. In addition the goatee is not frequently depicted on Tibetan men in painting and sculpture, which suggests it was not considered a common physiognomic characteristic. Even when a goatee or beard is implied, many portrait thangkas of Buddhist hierarchs show the men with only a faint outline of facial hair.18 The attempt to distinguish between different cultural backgrounds is also evident on the portrayal of the men’s hair and costume in The Royal Drinking Scene. The two centrally placed men and some of the attending figures have their hair painted in long and thick wavy strands (see Figs. 7.5, 7.6). In contrast, 18

See, for example, the portrait thangka of Taklung Thangpa chenpo (sTag lung Thang pa chen po) in Kossak and Singer (1998:91).

180

Alafouzo

the other male participants in the scene wear several thin straight plaits, and this type of hairstyle is also seen on the men in the murals of the Assembly Hall at Tabo (see Fig. 7.8). These men are dressed in traditional Tibetan outfits, which are loosely shaped with exceedingly long sleeves. The cuffs and collars are of contrasting material to the main fabric of the robe, which has wide lapels, and a belt tied around the waist. These characteristics were also the main features of the Yarlung dynasty men’s outfit.19 The pleated trousers that are integral to the men’s garb at Tabo, however, appear to be a post-Yarlung dynasty development. Both at Alchi and Tabo, the men with multiple thin plaits are dressed in a Tibetan costume, and thus can plausibly be deemed Tibetans. The long and thick wavy strands of hair of the centrally seated man (see Fig. 7.9) differ clearly from the 11th century hairstyle of the Tibetan men at Tabo, and they have been compared to Seljuq coiffure (Fig. 7.10).20

Figure 7.10

19 20

Seljuq bowl depicting seated and standing figures in a garden dated March–April 1187 CE (Muharram 583). Transparent cream coloured glaze overpainted in seven colours and then re-fired at lower temperature (mina’i). From Kashan, Iran. (Photograph © Trustees of the British Museum, London).

See Karmay (1976:69–74) for discussion on the Yarlung dynasty men’s costume. Flood 2005:80, fn. 57; Sims 2002:24. The Seljuqs (c. 1038–1198) were a Turko-Islamic dynasty, based in eastern Iran but who had nominal ruler ship over Turks in western Central Asia and Xinjiang.

royal Drinking Scene, Alchi

181

In 11th–12th century Islamic art, long braids are a characteristic reserved for people of Turkic origin and distinguish them from other cultural groups (Sauvaget 1951:31; Azarpay 1981:176).21 Twelfth century Seljuq figural style commonly portrays men’s braided hairstyle with just a few rather than many thick plaits, which curve slightly at the ends (see Fig. 7.10). The artists at Alchi were clearly attempting to portray a different hairstyle from the multiple thin plaits associated with Tibetan men. Although differing stylistically from the Seljuq plaited coiffure, it seems that the thick wavy strands of hair in the Dukhang were an artist’s interpretation of the Turkic hairstyle.22 The costume of the centrally seated man is very different from the 11th century Tibetan outfit. He wears a long-sleeved caftan with a deep v-neck, whose fabric has several roundel motifs and integrated armbands. The cuffs are folded back, and a wide red and white chequered belt is tied around the waist. Underneath the caftan, there is a round-necked undergarment. Light coloured trousers are tucked inside patterned short boots. The features of the Dukhang costume are therefore exactly the opposite of the traditional Tibetan outfit, which is made of plain fabric, is not decorated with designs of armbands and is loosely worn around the body. Scholars have noted that the Dukhang costume is Turkic, and that it was popular in Central Asia and regions under Turko-Islamic rule (Sims 2002:24). The Turkic outfit was developed from a Sasanian (c. 224–642 CE) prototype, and is well suited to horse riding. The armbands are particularly associated with Turko-Islamic peoples such as the Ghaznavids and the Seljuqs, although they were also popular in pre-Islamic Central Asia (Talbot Rice 1969:263).23 The murals in the Assembly Hall at Tabo depict male costumes with designs of armbands, and these types of outfits have been deemed non-Tibetan (Wandl:1997).24 The 11th– 12th century Turko-Islamic outfit was typically round-necked (see Fig. 7.10), while the v-neck, which seems to have originated from pre-Islamic steppe fashion, was depicted more rarely. The Dukhang costume does not therefore represent the popular Turko-Islamic fashion of the day but is somewhat archaic.

21 22 23 24

An early, c. mid-7th century mural from Afrasiab, in Samarqand in Uzbekistan depicts two Turkic delegates with long braids (see Grenet 2004:111). An identical hairstyle to the non-Tibetan one in the Dukhang is depicted in an 11th century mural on the north-eastern stupa at Tholing (see Namgyal 2001:122–23). Designs of armbands are depicted on Sogdian costume in the c. 500 CE murals of Panjikent in Tajikistan, and also on a foreign man in a 5th–6th century Cave 1 at Ajanta, India. See Klimburg-Salter (1997:Figs. 129, 130 and 202).

182

Alafouzo

The designs on the costume show strong parallels to block printed designs employed in western and north-western India, and thus the robe could be of Indian manufacture.25 The motifs on the fabric depict a four-legged creature with a long tail, set inside a roundel. Between the linked roundels there are red interstitial motifs, which resemble sketchy four-petalled flowers (see Fig. 7.9). The pattern of linked roundels, enclosing either floral or animal motifs, is common in depictions of textiles in 11th and 12th century Indian arts.26 Approximately contemporary with the Dukhang costume are 11th century Pala Buddhist manuscript illustrations, which show a variety of textile designs, including linked roundels.27 The creature inside the roundel on the Dukhang robe has strong stylistic similarities to a motif depicted on the back cushion of the throne in a late 11th century Buddhist manuscript from Nalanda, and also to a lion depicted on an early 12th century Jain manuscript cover from Rajasthan, north-western India.28 Closer to Alchi, linked roundels, which enclose floral motifs, are painted on a strip of actual cloth attached to the ceiling of the Assembly Hall at Tabo, and on the dhotis of Buddhist deities in the murals (Klimburg-Salter 1997:172, Figs. 113, 114 and 194). The pattern also appears on 11th century Buddhist metal sculpture from western Tibet, and the costumes of donors depicted on the north-eastern stupa at Tholing (mTho lding) have roundels enclosing either floral motifs or a fantastic creature that is very similar to the one on the Dukhang outfit.29 Thus the motif of linked roundels had a widespread popularity in the arts of the Himalayan region during the 11th century, and it seems possible that the design was copied from contemporary textiles. The above art-historical comparison suggests that the inspiration for the Dukhang textile motifs should be sought in the artistic traditions of India.

25

26 27

28 29

I am very grateful to Dr Ruth Barnes, textile curator in the Department of Eastern Art, Ashmolean Museum, Oxford, for discussing the Dukhang costume with me. For a full discussion of the man’s costume in the Dukhang, see Alafouzo (2008). John Guy (1998:57) notes that the linked roundels design appeared in Indian ornamentation from at least the first century BCE. For the design of linked roundels, see a manuscript illustration dated to c. 1025 from Bihar illustrated in Pal (1993:53, CAT 1B) and Pal (1993:57) for a c. 1075 manuscript from Nalanda, Bihar. See Huntington (1990:plate 58b) for the throne creature and Nawab (1980–1985:colour plate 73 for the lion). For an example of the pattern on western Tibetan metal sculpture, see Pal (2003:135, Fig. 85). For Tholing donors, see Namgyal (2001:122).

royal Drinking Scene, Alchi

183

The two centrally placed men and some of the soldiers in the Dukhang wear a narrow scarf with long ends around their heads (see Fig. 7.5). This has a striking similarity to the 7th century headgear of the Turkic qaghan, described as follows by the Chinese pilgrim Xuanzang: “[The Kaghan’s] head was bare and wrapped only in a silk band more than one zhang [3.2 metres] in length, with its ends hanging behind” (Baibakov 1998:221). The Dukhang headdress has been compared to that of the Seljuqs (see Fig. 7.10), as they have the long floating ends in common (Sims 2002:24). The Dukhang headband, however, leaves the top of the head bare while the Seljuq men wear a type of cap, which covers the head completely, demonstrating that the two headdresses are quite different in style. A remarkably similar but much earlier comparison to the Dukhang headgear is to be found in a c. late 7th century woollen hanging excavated in Antinöe, Egypt, in which we see a horse rider wearing a headband with long floating ends.30 The headband depicted at Mangyu is like the one in the Dukhang (see Fig. 7.2) but there are no other comparative 11th century Tibetan art-historical examples. The Tibetan men in The Royal Drinking Scene wear hats suited to the local climate rather than headbands (see Fig. 7.1). It seems that while the Dukhang headband had remained practically unchanged from the 7th century Turkic one described by Xuanzang, the Seljuq headdress had developed from the earlier style. This art-historical analysis serves to demonstrate that the origins of the costume, hairstyle and headdress of the centrally seated man in the Dukhang are to be found in Turkic fashion from Inner Asia.

The Ruler, the Cup and the Hunt

The above discussion can be reinforced by examining another mural in the Dukhang, which I shall refer to as The Ruler, the Cup and the Hunt (Fig. 7.11). The mural is clearly divided into three registers and is situated on the eastern wall of the Dukhang in a more prominent position than the drinking scene. The uppermost register portrays a centrally seated man who is flanked on his right by a seated male offering him a cup and to his left by a seated and a standing man (Fig. 7.12). The middle register depicts a group of nine men, two

30

Illustrated in Otavsky (1998:154, Fig. 86). The rider’s outfit and the other motifs on the tapestry suggest a Sasanian influence (Kitzinger 1946:41–42; Otavsky (1998:16).

184

Alafouzo

Figure 7.11

The Ruler, the Cup and the Hunt, the Dukhang, Alchi. (Photograph courtesy: Jaroslav Poncar 1989, WHAV Alchi_D_JP89_9.3.8).

Figure 7.12

Detail of The Ruler, the Cup and the Hunt. (Photograph courtesy: Jaroslav Poncar 1983, WHAV Alchi_D_JP83_9.3.12).

royal Drinking Scene, Alchi

185

of whom are kneeling and offering a cup to the centrally seated man. The rest of the men are standing, and many of them are armed (Fig. 7.13). The lower register shows men on horseback holding bows (Fig. 7.11).

Figure 7.13

Detail of The Ruler, the Cup and the Hunt. (Photograph courtesy: Jaroslav Poncar 1983, WHAV Alchi_D_JP83_9.5.1).

The centrally seated man is portrayed in the full frontal position, or bagdas. His costume, albeit less luxuriously patterned, headdress and hairstyle are practically identical to those of the central man in the drinking scene in Fig. 7.1. The men in the middle register are clearly engaged in a drinking ceremony, which involves two kneeling cupbearers wearing ceremonial scarves, a covered three-legged table and several drinking vessels on the ground (see Fig. 7.13). The accompanying men carry weapons and are dressed in v-necked caftans, and wear either headbands or flat, wide-brimmed hats. On the basis of the analysis applied to The Royal Drinking Scene, it would appear that the Ruler, the Cup and the Hunt mural also shows a drinking ceremony but with certain iconographical differences. The centrally placed man’s full frontal position strongly suggests that he is the ruler (see Fig. 7.12), and his rank is further emphasised by the halo around his head and the parasol placed directly above him. The two seated men flanking him are depicted

186

Alafouzo

in a three-quarter view, and one of them is offering the ruler a cup, thereby following a pre-defined royal protocol. The ruler and his men all wear nonTibetan costumes and headdresses and therefore it is likely that the mural portrays a foreign ruler with his soldiers. These armed men dominate the entire scene, which is devoid of any Tibetan men or women. The Ruler, the Cup and the Hunt painting illustrates three distinct themes, those of the ruler seated amongst his courtly attendants, the drinking ceremony and the hunt. Comparative 11th century art-historical evidence suggests that these themes are not depicted in Tibetan painting, except in the murals at Alchi and also at Mangyu, where horse riders with bows accompany the cup offering (see Fig. 7.2). Therefore, the above themes may be deemed foreign cultural components, which during the 11th century appeared in Ladakh, but seemingly not elsewhere in Tibet. The scenes in The Ruler, the Cup and the Hunt painting correspond to the princely imagery frequently portrayed in Persian painting, where each theme can be represented in conjunction with another or on its own. Princely imagery has a long history in the visual arts of Persia, making an early appearance during the Sasanian period (Sims 2002:40). The themes of the ruler at the centre of courtly adoration, the cup offering and the hunt, however, should not only be referred to in the context of Persian painting. The textual evidence discussed above makes clear that the cup offering formed an important part of Turkic customs in Inner Asia, where the rite was often performed in conjunction with the hunt or during tribal gatherings.

Historical Evidence

The above art-historical discussion suggests a strong Turkic connection in 11th century Ladakh. There is no direct reference to Turks in Ladakh in the La dvags rgyal rabs, which is the main source for the pre-15th century history of the region. The chronicle was largely compiled in the 17th century, and is thus of limited use in attempting to discuss 11th–12th century events in Ladakh. Regarding the early medieval history of western Tibet, a 15th century text, mNga’ ris rgyal rabs, translated and extensively commented on by Roberto Vitali under the title The Kingdoms of Gu. ge Pu.hrang, also includes earlier historical events in the region. As with the La dvags rgyal rabs, the accuracy of its historical data is fragmented because it is based on much later material from different sources. Nevertheless, an 11th century event in Guge (Gu ge) referred to in the mNga’ ris rgyal rabs will be re-examined below as it has possible implications for early medieval Ladakh.

royal Drinking Scene, Alchi

187

According to mNga’ ris rgyal rabs, the monastery of Tholing in Guge, founded in 996, was attacked in c. 1037 by the troops of the Hor (Vitali 1996:288–89).31 The text refers to the invader of Tholing as Bha ra dan dur of the lineage of Hor nag mo A lan. Vitali has argued that the latter is a Tibetan translation of the dynastic name Qarakhanid as Hor means Turk and nag mo is black, thereby implying a ‘Black Turk’ (Vitali 1996:287). The term Qarakhanid is, however, a modern scholarly conventional one derived from the most important titulature of the dynasty, Qara Khan, or ‘Black Ruler’ (Golden 1990:354; Golden 1992:214; Paul 2002:71).32 Islamic sources of the time called the dynasty ’Āl-i Afrāsiyāb’ and ’al-Qāqāniya’ (Golden 1992:214) and thus it seems improbable that on linguistic or historical grounds Hor nag mo A lan refers to the Qarakhanids. It is also unlikely that A lan is a Tibetan phonetic spelling for the Turkic Arslan Khan, a traditional title for the head of the Qarakhanid Eastern qaghanate, as the Tibetan word does not include the actual epithet for the ruler, Khan.33 The attack by the Hor on Tholing was seemingly short-lived and the actual text does not refer to any large-scale damage (Vitali 1996:289, fn.443). On the contrary, the Hor restored a temple and did not appear particularly antiBuddhist, which also suggests that the Islamic Qarakhanids were not behind the raid. Although it has traditionally been assumed that by the 11th century the Qarakhanids had conquered Khotan on the Southern Silk Road, there is no direct evidence for this in the historical records (Horlemann 2007:95–96). The Qarakhanids had established themselves in Kashgar from the late 10th century onwards, and thus, on geographical grounds, it seems unlikely that they would have undertaken a journey that entailed weeks of travelling over difficult terrain to accomplish a short-term raid on Tholing. There is evidence for the presence of Turks in today’s Northern Areas of Pakistan to the west of Ladakh during the 11th century. The Persian geographer and scientist Al-Bīrunī (973–1048), who worked for the Ghaznavid court, is still known today for his work translated into English as Albērūnī’s India, which discusses the geography and customs of India c. 1030. Al-Bīrunī noted that the north and the east of Kashmir belonged to the Turks of Khotan and Tibet 31 32 33

The text also refers to two further raids on Guge during the 12th century, which according to Vitali, have a foreign involvement (Vitali 1996:347, 350–51, 354–55). Sinor 1998:236 notes that the exact significance of the epithet black (‘kara’) in tribal and personal names has not been fully established. The title Arslan Khan may not have even been in use at the time of the Tholing raid in c. 1037, as the Qarakhanid dynasty did not divide into the Western and Eastern branches until 1040.

188

Alafouzo

(Sachau 2002:193). He also referred to the Turkish tribes called Bhaṭṭavāryan, who lived in Gilgit, Aswira and Shiltas, i.e. the modern Northern Areas of Pakistan (Sachau:195). Bhaṭṭavāryan is a Sanskrit word, where bhaṭṭa means ‘lord’ or ‘master’ and vāryan implies ‘best’ or ‘chosen’, and thus the name would be appropriate for a dynasty.34 This dynasty, because of its Sanskrit name, could have been Buddhist or Hindu or both, but not Muslim. There is some evidence that Buddhism continued to thrive in Gilgit at least until the late 10th century. Vitali has noted two passages on the Tibetan Buddhist master gNubs chen, which refer to his translation activities in Bru zha in the late 9th century (Vitali 1996:167, fn. 230). A late 10th century itinerary written in Khotanese describes a journey through Gilgit and Chilas to Kashmir, and the text notes eight stone sanghāramas, or vihāras, in Gilgit (Bailey 1936:262). It is worth noting that AlBīrunī does not identify the Turks of Tibet and Khotan. The Ghaznavids and the Qarakhanids enjoyed diplomatic and military relations with each other, and thus the latter would almost certainly have been known to Al- Bīrunī. His lack of knowledge of the Turks to the north and east of Kashmir might imply that they belonged to non-Muslim Turkic tribes that were different from the Bhaṭṭavāryan in Gilgit. The mNga’ ris rgyal rabs recounts the abduction of the Guge king ’Od lde in Bru zha during a military campaign, and his subsequent escape and death in Baltistan in 1037 (Vitali 1996:281–87).35 While the historical validity of these events cannot be confirmed, they nevertheless suggest that a king of Guge went through Ladakh and was captured in the Northern Areas of Pakistan. His captors may have been the Buddhist Turks mentioned by Al- Bīrunī, and it is possible these Turks ventured as far as Tholing in 1037 where they frightened the locals but did not damage the Buddhist temples. Significantly, Turkishsounding names begin to appear in the Ladakhi list of kings after ’Ol lde’s demise. According to Luciano Petech, names such as gZi De khyim and De mur are of Turkic origin, whereby gZi De khyim is a Tibetan rendering of the ancient Turkic title Tegin and De mur refers to Turkic Temur (Petech 1997b:110).36 In connection with the present discussion, the importance of the Ladakhi lineage rests less on the accuracy of its succession than on the presence of foreign names, some of which are Turkic. It is unlikely that the compilers of Tibetan royal genealogies would have included foreigners in their lists unless 34 35 36

Bhaṭṭa is known in Sanskrit Buddhist as in bhaṭṭaraka, which is Tibetan rje btsun (‘lord’). Philip Denwood, personal communication. Bru zha denotes Yasin, which is situated above Gilgit in the Northern Areas of Pakistan. There is also evidence of Turkic names in the 13th–14th century Wanla inscription (see Tropper 2007).

royal Drinking Scene, Alchi

189

there were good grounds for doing so. Significantly, Turkic names appear in the Ladakhi genealogy, which is corrupt in terms of accurate dates and names, but which nevertheless gives a strong hint of non-Tibetan rulers. There is a very clear reference to non-Muslim Turks in Tibet c. mid-11th century in the Islamic text, Annals of the Saljuq Turks. The entry is recorded as a miscellaneous event for the year 1046–47 and reads as follows: From the land of Tibet this year came a great horde of Turks beyond numbers. They made contact with Arslan Khan, the ruler of Balasaghun, thanking him for the good way he ruled his subjects. They caused no trouble or harm to his subjects but took up residence there. He sent to them and asked them to convert to Islam. They did not accept but remained friendly. (Richards 2002:62).37 These Turks could have been Buddhist, and although their exact identification and habitat remain unascertained, their mere presence in Tibet is significant in the context of the events discussed above. In conclusion, the portrayal of the ruler holding a cup denotes royal power and is frequently found in Islamic arts across a wide geographical area. Significantly this image is completely absent in the Dukhang artistic representation discussed above and instead the emphasis is on the cup offering to the male central figure. In The Royal Drinking Scene the action focuses on the Tibetan woman making the cup offering to a foreign man of royal rank. In contrast, in The Ruler, the Cup and the Hunt scene the non-Tibetan soldiers perform the act of offering the drinking vessel to the frontally seated ruler. In both murals, therefore, power is centred on a foreign man portrayed in the middle of the composition. The Royal Drinking Scene represents a rare depiction of a cup offering by a woman, which was attested to in Turkic literary records but hardly ever depicted in surviving works of art. The context and the iconography of the mural suggest it shows a marriage between a Tibetan woman and a foreign man. The Ruler, the Cup and the Hunt scene portrays the penultimate stage of the enthronement ritual, which is referred to in the 13th century written records of the Mongols, but the origins of which are drawn from ancient Inner Asian steppe customs. Therefore, these two secular murals in the Dukhang represent manifestations of political power within a Tibetan Buddhist context, accompanied by scenes of hunting and military grandeur influenced by a non-Tibetan courtly tradition. It is likely that Mangyu was part of the Alchi ruler’s domain, as the subject matter and costumes in the temple’s 37

Balasaghun is situated to the west of the Lake Issyk-Kul in Kyrgyzstan.

190

Alafouzo

mural are practically identical to those in the Dukhang. The issue of the exact identity of the Turks in Ladakh may remain unresolved until further documentation emerges. It seems certain, though, that these Turks, despite their hostile intention to gain new territory, were Buddhist and allowed Buddhism to flourish, as there was no adverse effect on the long-term continuation of Buddhism and Tibetan culture in Guge and Ladakh.

The Wood Carvings of Lachuse. A Hidden Jewel of Early Mediaeval Ladakhi Art Heinrich Poell

Introduction; The Setting

The small village temple of Lachuse (Figure 8.1)—located in a remote corner of Western Ladakh1—is a stone building with a single square room without windows, accessed through a door in the front wall; light is admitted through an opening in the roof. The temple is built into a steep slope slightly above the village, so that at the back of the building one can step directly onto the roof. This is a somewhat unusual setting for an early temple, all the more remarkable as there is plenty of level ground in and around the village. Originally the building had a recessed vestibule, but this is now largely obscured by the clumsy and unnatural addition of a monks’ house that has been built in front of the temple.2 At first sight the building appears to be in a good state, but closer inspection shows extensive repairs to all exterior walls, and added masonry for the support of the façade and the side walls. The sagging front beam supporting the roof is broken and rests now on the monks’ house built into the main façade. The masonry of the monks’ house and of the walls buttressing the original construction is of much lower quality than the original walls. On the inside (Figure 8.2) the walls are covered with modern ‘Thousand Buddha’ murals, interrupted on the back by a niche with two clay stupas; this is obviously a later addition to the original plan. The ceiling has undergone extensive renovations and alterations; the main beams and the rafters are now of rather irregular shapes and sizes, with much new material between reused old wood. Within this largely despoiled setting, wood art of exceptional quality and art historical significance has been preserved. The entrance to the temple has 1 Lachuse lies about two hours’ walk south–west from Kanji, the next major village, in a side valley off the ancient trade route from Lamayuru over the Kanji La (pass) to Ringdom in Zangskar. 2 The inside dimensions of the temple are approx. 6.2 × 5.0 m; the original vestibule would have measured about 4.8 × 2.1 m. Cf. the ground floor plan in Figure 8.1; the original walls of the building are shown in grey.

© koninklijke brill nv, leiden, ���4 | doi ��.��63/9789004271807_���

192

Figure 8.1 Lachuse temple, view from the South–East and floor plan (© John Harrison/ Achi Association).

Poell

The Wood Carvings of Lachuse

193

Figure 8.2 Lachuse temple, inside view.

a doorframe with scenes from the Buddha’s life story and other figural decoration. In the vestibule several carved pillars and lintels, loose architectural pieces and a large sculptural panel indicate that the temple must once have had an impressive wooden façade. On the inside there are four fluted columns, two of which still have magnificent capitals, and four bracket lions on the walls. Like most other wood art in Ladakh, Lachuse has been largely ignored in the growing body of research on Ladakh;3 the temple has been mentioned only rather recently (Luczanits 2005:80–82) in a review of artistic developments in early mediaeval Ladakh, where it was somewhat globally identified 3 For the wood art of this period, there are some useful remarks in Khosla 1979:56–60 and—focussing on the Alchi Sumtsek (gSum btstegs)—in Goepper (1996b) and Goepper (1996a:23–29). An overview of the extant carved temple doors in Ladakh is attempted in Poell (2004). The wood art of Kinnaur and Spiti (which has of course close links with Ladakhi creations) was discussed by Luczanits (1996) and more recently by Klimburg-Salter (2002) and Di Mattia (2002). Wooden temple doors in Western Tibet have been published by Tucci (1973:Figures 133–138), Klimburg-Salter (1988b:Figure 1–3), Aschoff (1987:46), Admin. Comm. Archaeology (1991:Figures 20, 24; plates XVII, XXX, XXXIV, XLVI, XLVII, LVIII, LXIV, CLV), Wei and Li (2001:Figures 148–165); most recently Neumann and Neumann (2008) have analyzed the important wooden temple door in Khojar in great detail. Di Mattia (2007) has traced the evolution and origins of the Kashmiri architectural niches.

194

Poell

as a “predecessor” to the wood art of Alchi and related monuments. However, the wood art of Lachuse merits a more detailed discussion; as will be shown in this paper, it dates to different periods before and up to the late 12th century, and draws on influences from several neighbouring regions. Tracing the relations of these woodcarvings with artistic creations of other areas in Ladakh and beyond will permit a glimpse into the evolution of Ladakhi art in the early mediaeval period.

The Main Door

The doorframe in the centre of the front wall (Figure 8.3) is not a proper door construction; it was assembled from available pieces of carved wood by simply balancing the lintel on the two pillars on the left and right, and fixing the whole arrangement to the threshold and the mud walls.

Figure 8.3 Portal of the Lachuse temple.

The Wood Carvings of Lachuse

195

The Lintel The present lintel consists of two panels; the upper one is decorated with a kīrtimukha frieze, where a string of diamonds separates a row of lions’ faces from a band of half lotuses. A similar design can be seen on the doorframes in Mangyu and Sumda Chung where the lions’ faces are joined by a garland through their mouths.4 This garland is missing on the present panel. The rather crude carving and clumsy design of this panel is in marked contrast to the highly accomplished carving of the other wood art in the temple. It is also markedly inferior to the workmanship of the kīrtimukhas on the doorframes in Mangyu and Sumda Chung. On the lower lintel five scenes5 from the life of the Buddha are set in Kashmiri-type architectural niches of columns supporting stepped roofs or foliated arches. All columns are identical, with elaborate multi-tiered bases and capitals, while the roofs all have different designs; there are additional leaves and garlands on the outside of all roofs, and all columns except those of the first frame have a decorative scroll on top. Beginning from the left, the first scene (Figure 8.4) shows an archer and three attendant figures, one crouching or kneeling, the others hovering in the air and gesturing towards the central figure. The archer wears a necklace and jewellery, and a mālā falls down over his arms and back. The auxiliary figures also wear jewellery and have mālās around their arms and shoulders. Only the upper torsos are shown of the two female figures that emerge from the background; this identifies them as celestial beings (devatas). On top of the right column there are several objects—resembling trees or plants—in two rows, possibly the target for the archer. In the second scene (Figure 8.5) the central figure is shown walking to the left in a somewhat exaggerated pose. Note the distinct post-Gupta period head and hairstyle, the jewellery on the ears and around the neck, the mālā and the fold of the dhoṭī, all similar to the first scene. As on the first panel, the three figures on the right and the one on the lower left are devatas, shown in reverential poses and gesturing towards the central figure. The upper figure on the left, now almost obliterated, was apparently much larger than the other 4 This design can be traced to the art of Kashmir; cf. for instance the kīrtimukha friezes on the temples in Avantisvami and Parihasapura illustrated in Kak (1933:pls. L, LXX, LXXIV). 5 There was (at least) one more scene on the left end of the panel, as can be seen on Figure 8.3. On the right the lintel is cut off at the end of the last visible scene; it is thus not clear if there were more scenes on this end of the panel as well. However, I would conjecture that the panel had originally two more scenes, as an odd number of frames places the central scene into the most prominent position.

196

Figure 8.4 Scene 1, Buddha life frieze, Lachuse doorframe.

Figure 8.5 Scene 2, Buddha life frieze, Lachuse doorframe.

Poell

The Wood Carvings of Lachuse

197

Figure 8.6 Scene 3, Buddha life frieze, Lachuse doorframe.

auxiliary figures on the panel, and seems to have been in some interaction with the central figure. The next scene (Figure 8.6) on the lintel shows the main figure in threequarter profile facing the viewer. The figure steps with one foot on a small figure on the ground in the centre of the panel, while at the same time engaging with the other four figures to the left and right. These auxiliary figures are all carved in three-quarter profile, but other than on the first and second panels here they have full bodies and legs. They all seem to be males, wear prominent necklaces and mālās, and are all in the same pose indicating movement and touching or clasping of the limbs of the central figure. On the left of the fourth scene (Figure 8.7) there is a building with two stories and an open pavilion with a pyramidal roof on top. The main figure in the centre is seated on a lotus base in the ‘royal ease’ pose, one hand is raised and holds an object, the other rests on its knee. A kneeling female on the right interacts with this figure; note the female’s jewellery, the dhoṭī pattern on the left leg, and the distinct hairstyle. Below there are two auxiliary figures whose heads protrude over the floor into the upper scene. On the last scene of the panel (Figure 8.8) a richly caparisoned horse with a rider is shown moving to the right. Three kneeling figures support the hoofs of the horse.

198

Figure 8.7 Scene 4, Buddha life frieze, Lachuse doorframe.

Figure 8.8 Scene 5, Buddha life frieze, Lachuse doorframe.

Poell

The Wood Carvings of Lachuse

199

This scene is of course easily recognized as the Great Departure, and thus gives a point of reference to the Buddha’s life story. The scene to the left can then be identified as Prince Siddhartha’s life in the palace, and the first scene on the panel should be the archery competition Siddhartha had to undergo in order to win his wife Gopa. It seems also sufficiently clear that the centre of the panel shows a wrestling scene, again part of Siddhartha’s competitions for his wife. Thus only the second scene remains unclear. A possible interpretation could be the Throwing the Elephant episode; the damaged figure on the left is too big to be just another auxiliary figure, and the main figure in the centre could be taken as clasping with its right hand the trunk or the bottom of the animal. Leaving aside for the moment the question whether the panel was part of a complete life story—with the other parts no longer extant—or a stand-alone piece, we can safely say that it was intended as a continuous narrative of the Bodhisattva’s life from his adolescence to the Great Departure. This would mean that the artist has relied on a textual or oral source that places the Four Encounters before the episodes relating to Siddhartha’s marriage; it is highly unlikely that the Encounters would have been omitted altogether, given their pivotal role in the Buddha’s biography. This is at variance with the canonical texts, which place the Four Encounters directly before the Great Departure.6 However, in most depictions of the Buddha life in Western Tibetan and Ladakh before the 14th century there are variations in the selection of scenes and in the sequence of the depicted episodes; it seems that at that time the iconography of the Buddha life had not yet been standardized into a universally applied model. The sequence on our panel corresponds with the life story that is painted on the Maitreya statue in the Alchi Sumtsek, where the Four Encounters are inserted between episodes of Siddhartha’s adolescence and the story of the prince’s marriage and life of pleasure; the slightly earlier Dukhang door in Alchi puts the Four Encounters immediately before the Great Departure, while on the even earlier temple

6 See Csoma de Kőrös (1863) for an abstract of the Buddha life from the Tibetan canon, Foucaux (1848) for a Tibetan translation of the Lalitavistara (found to be in virtually complete agreement with the Sanskrit original), Krom (1925) and Seidenstücker (1914) for the Lalitavistara as a model for sculptural cycles on the Borobudur (Ctrl. Java) and in Pagan (Burma), and Rockhill (1884) and Bu ston (1932) for versions of the Buddha life based on the Lalitavistara that were (presumably) known to Tibetans of this period. In all these sources, the Encounters take place just before the Great Departure, after the Contests and Marriage.

200

Poell

door at Nako (ca. mid-11th century) the Encounters are again shown before the contests for the marriage.7 Both the architectural settings and the overall composition of the scenes are clearly indebted to Kashmiri artistic conventions.8 However, the art of Kashmir did not provide iconographic models for the whole life story, as life story sequences are unknown in the extant corpus of Kashmiri Buddhist art.9 Neither could the Lachuse artists have relied on Gandharan iconography, as the archaeological evidence indicates that there were also no complete life cycle depictions in Gandhara.10 Rather, Gandharan stupas and other edifices were decorated with an eclectic mix of scenes from the Buddha life and from 7

8

9

10





See Luczanits (1999) for the Sumtsek paintings, Snellgrove (1977:19 and pl. III), Khosla (1979:Figure 30) and Poell (2004:15 and Figure 1a) for the Dukhang door, and Ziegler (2008) for Nako. Cf. the niches and group compositions of the figure panels of the 8th/9th century Avantipur temple in Kashmir (Kak 1933:pl. L; Fisher 1989:39 and Figure 17, 107 and Figures 1, 2, 3). Only a very small number of pieces of Kashmiri origin illustrate events from the Buddha life: a stone fragment of the birth scene in the Srinagar museum (Kak 1923:38), two bronze aureoles preserved in Tibet (Pal 2001:116–117; Schroeder (2001:113 and Figure 21), and about ten miniature ivory carvings (Czuma 1989) with single events from the Buddha’s biography. This small number clearly indicates that complete life stories were not a popular subject, as otherwise many more pieces should be extant. Despite over 150 years of extensive excavations, only a few extended sequences of episodes from the Buddha life have been found in the Gandhara area, whereas panels with one to three individual scenes number literally by the thousands. Most of these pieces show a slight curvature, indicating that they were intended to embellish stupa drums of varying dimensions—from small votive shrines to larger ones of several metres diameter.  The typical finds from these sites are several panels with one to three scenes that may or may not be related in time, and frequently mixing jātakas and life stories. Some of these excavated sets of stupa decorations can be arranged into larger ensembles (see Behrendt 2005; cf. also Behrendt 2004:120–121 and Figure 108, 116–117 and Figure 93), but of course there might have been more panels, and their exact sequence remains more or less conjectural. The sole undisturbed find of a Gandharan stupa with life scenes is the well-known Sikri stupa of the Lahore museum, where the complete set of 15 panels is preserved. (See Alam 1994; the controversy over the exact sequence of the Sikri panels is irrelevant for the present argument.)  There is also a small number of single pieces with extended sequences of life scenes: two votive stupas with eight scenes around their bases from the Indian Museum, Calcutta (see Grünwedel 1901:154 and Figure 106, 156 and Figure 107; a similar piece is in the British Museum collection), and Luczanits (2008:184 and Figure 1); a panel with six scenes from the Musée Guimet (Klimburg-Salter 1995:193 and Figure 175), a harmika block of a votive stupa with four life scenes (Luczanits 2008:203), a stela with six life

The Wood Carvings of Lachuse

201

the jātakas that served for the edification of believers who were presumed to be familiar with the life of the Master. I have shown elsewhere (Poell 2004:196; Poell 2011:31) that the complete depiction of the Buddha life is a Western Tibetan innovation, drawing on iconographic and stylistic models from Gandhara; these models were probably transmitted through Kashmir, where they had lingered on in various forms.11 These pictorial cycles evolved in response to the need for a complete life story for Buddhist education and religious propaganda. The earliest examples are the carved doorframes of Tholing, no longer extant, and Khojar in Western Tibet, dating to around the late 10th century; nearer to Ladakh, the Buddha life carvings on the Nako temple door date to the mid-11th century.12 The carving on the panel is highly expressive—one might say, at times even mannered—with all figures knit into coherent compositions and closely interacting with the central figure. However, the emphasis is always on the central figure which in all panels is perceptibly larger than the other figures. On the first three panels the auxiliary figures are arranged around the central figure.







11 12

scenes (Luczanits 2008:324), and the drum of a votive stupa with eight scenes (Luczanits 2008:204 and 193–195).  None of these finds comes anywhere near to showing a coherent life story of the Buddha, nor could they have been part of a complete life cycle. A typical set of panels (Luczanits 2008:204) shows a jātaka, the Birth, Asita, two childhood and adolescence scenes, the Grasscutter and Nirvana. The Sikri stupa, the votive stupas cited above and the larger sets discussed in Behrendt 2005 all show a similarly eclectic selection of episodes, jumping from Birth and pre-Departure events to post-Enlightenment miracles and Nirvana.  I have argued in Poell (2011:14ff.) that Gandharan stupa decorations were not intended as biographical stories. Rather, scenes were selected for edification during circumambulation, and to provide points of identification for laypeople—showing how ordinary persons gained merit through interaction with the Master, as an example for the present day laity. Behrendt (2004:391) comes to the similar conclusion that the “selective combination of episodes was probably intended to give stupas a particular religious emphasis.” (Cf. also Stoye 2008.)  Pivotal to this argument is the extreme paucity of the Four Encounters in Gandharan art. So far only two examples have been described (Fischer 1958; Dye 1976). In my mind this indicates an absence of complete life stories in Gandharan art, as it is highly unlikely that the Encounters would have been omitted from depictions of the Buddha’s life story, if indeed the artists had wanted to show a coherent biography. See Siudmak (1995) for an analysis of the transmission of stylistic features from Gandharan art into Kashmir. For the Tholing and Khojar doors, see Tucci (1973:134 and Figure 136), and Neumann and Neumann (2008); for Nako see Ziegler (2008).

202

Poell

Figure 8.9 Three scenes from the Buddha life on the Dukhang door in Alchi.

This composition can be traced back to late Gandharan models of hieratical groups with a towering Buddha surrounded by smaller figures, many of which are celestial beings celebrating the achievements of the Master. In the art of Kashmir the same compositional principle can be seen in portable ivory shrines and on a few sculptural panels in Kashmir temples of the eighth and ninth century.13 That is in clear contrast to earlier Gandharan narrative panels, which generally have figures of the same size interacting with each other, and no celestial beings. At Lachuse each scene focuses on a larger–than–life, central Buddha figure that is shown in a pose related to the particular scene, resulting in somewhat static compositions. By contrast the carvings on the Dukhang door in Alchi put the emphasis on the story and the depicted action; the figures there are all of the same size and interact naturally with each other, instead of being shown in a reverential assembly around a towering central Buddha. The three panels from the Dukhang portal—archery, married life and the Great Departure— shown in Figure 8.9 illustrate this point when compared to the corresponding scenes in Figures 8.4, 8.7 and 8.8. In all three scenes, the actors on the Dukhang panels have much more natural poses than at Lachuse. In the Dukhang panel illustrating the married life of the Bodhisattva the couple is shown seated on the same bed and in a conversation of equals, whereas at Lachuse we see the formal interaction of the wife and attendants with a static royal figure. In the Departure scene, the Dukhang panel shows both horse and rider in full motion; 13

See Czuma (1989) for the ivory carvings, Kak (1933: Pl. L) and Fisher (1989:39 and Figure 17, 107–108 and Figures 1, 2, 3 for Kashmiri figure compositions), and Zwalf (1996:135, Figure 221 and pl. X) for Gandharan examples. The wooden shrine of the British Museum ivory is discussed in Barrett (1969).

The Wood Carvings of Lachuse

203

Figure 8.10 Votive shrine; painted wood; Kashmir, 8th/9th century, British Museum.

at Lachuse the horse seems to be frozen in a pose of motion, supported by adoring devatas. All figures on the panel have typical post-Gupta period heads—broad, round faces with hairs curled and piled up on top and framing the face. This is rare in the art of Western Tibet and Ladakh; the only examples I know of are the Buddha life panel on the Nako doorframe and the lintel of the doorframe of the Dromtön (Brom ston) Lakhang in Tabo. In Kashmir, such post-Gupta period heads are only found on some singular stone sculptures and on the wooden votive shrines of the British Museum and Kanoria ivories (Figure 8.10; see Czuma 1989:Figures 1, 3 for the Kanoria ivory). Post-Gupta period faces and hairstyles are much more common in Himachal wood sculpture of the 7th to 9th centuries (such as the Brahmaur portal), which in turn depends upon earlier stone statuary such as can be seen in the Masrur rock temples. One should also note the careful carving of the niches on the Lachuse lintel, in particular the profiles of the roofs and the decorative leaves. The design itself is more complex than in Alchi and in Mangyu, and nearer to Kashmiri

204

Poell

models of the late 9th century.14 The niches in Lachuse are obviously important compositional elements, indicated by the careful carving of their decorative elements and by the rather static positioning of the figures within the (largely identical) frames. By contrast, on the Dukhang door in Alchi the niches are not important in themselves and are treated more as frames for the composition than as architectural devices; roof designs are simplified, and figures are frequently positioned in front of the columns, which at times are omitted altogether. The Door Jambs The two door jambs show a sequence of figural panels of varying heights; all figures are framed by Kashmiri–type architectural niches. These niches are severely damaged and almost obliterated in some parts, but it is clear that they are of somewhat simpler design than the ones on the lintel. On the left pillar (Figure 8.11) the top panel shows a figure that might be taken for a female, standing frontally with just a hint of tribhaṅga; a mālā falls over the arms down to below the knees and is passing between the legs. The figure wears some jewellery and an elaborate headdress, and there is a starshaped halo behind its head. Apparently there were six arms, but this is not entirely clear as the carving is very indistinct in this area. The niche has a trefoil arch within a simple sloping roof. The next panel shows the Buddha in meditation, seated in a closely fitting niche where the trilobed arch almost encircles the head and shoulders of the figure. The figure wears no jewellery and has no attributes. To the left and right there are two leaves or branches, suggesting an outdoor setting for this scene. The third panel is very similar to the first panel above. However, here the figure clearly is a female and is shown in a more animated pose than in the top panel. The figure has apparently only two arms, and holds an attribute, probably the stalk of a flower, in the left hand. Finally, the last panel on this pillar has been cut off at its end, leaving only the stepped roof and the head and upper torso of a crowned figure with a star-shaped halo as on the other panels above. Moving to the right pillar (Figure 8.12), we see a ferocious figure in a threatening pose on the top panel. The figure has two arms and wears jewellery, the fold of the dhoṭī and a mālā can be seen between its legs, and it has a garland 14

Cf. the niches with trefoil arches from Avantisvamin and Pandrethan temples, illustrated in Kak (1933:Pl. LXIV, LXIX) and Fisher (1989:34 and Figure 6, 36 and Figure 8, 113 and Figure 15). The origin of these niches can be traced to the architecture of Gandhara. See Behrendt (2004: Figures 11, 96), and Di Mattia (2007:55). Khosla (1979:73 and Figure 14) has shown the evolution of the Ladakhi niches from their Kashmiri models.

205

The Wood Carvings of Lachuse

Figure 8.11

Left pillar of the Lachuse portal.

Figure 8.12

Right pillar of the Lachuse portal.

206

Poell

with a string of pearls in its right hand, while the attribute in its left is now indistinct. The niche has two trefoil arches on top of each other, resting on columns with a deep groove along their face. This design is clearly different from the other architectural niches on the doorframe. The second panel shows a standing (female?) figure; the design is similar to the top and third panel on the left, but here the pose is more frontal and the mālā falls down in a wider circle. The figure has an elaborate headdress or crown and wears big earrings; streamers from the headdress fall down to its shoulders, and behind the head there is a star-shaped halo. The next panel shows a seated figure with one hand in front of its chest; the other hand is raised to the right, with a streamer falling down onto the right knee. The slanting line running behind the figure’s head from its raised hand to the top of the panel might be a sword, which would then identify the figure as Manjushri. The frame has a trefoil arch as roof, with additional decorative items on its outside; this design is more complex than on the second panel on the opposite pillar, where also a seated figure is shown. As on the opposite side, the bottom panel has been cut, leaving only the chest and the head of a figure that might have been seated or standing. The figure wears a crown and earrings, and streamers fall down from its head; it seems to have had only two arms, but this is not entirely clear. The niche consists of a trefoil arch supporting a sloping roof. Overall the panel seems to have been very similar to the second panel above. The iconography of the two pillars clearly refers to a Buddhist pantheon, but—apart from the Buddha on the left, and Manjushri and the dvārāpala on the right—a precise identification of the depicted figures is impossible due to the decay of the wood and the resulting disappearance of most details and attributes. However, it seems that the two jambs were conceived as a pair where standing female deities alternate with seated Buddhas/Bodhisattvas, with a protective guardian on top of the two columns (now missing on the left one). Figure 8.13 shows a conjectural reconstruction of this assembly of Buddhas and goddesses. The reconstructed jambs would measure about 200 cm, assuming an undecorated part at the bottom (as on all other Ladakhi doorframes from the period) and some additional decoration or a simulated capital on top. A similar arrangement can be seen on the main pillars of the Dukhang door in Alchi, where two male, multi-armed Bodhisattvas alternate with two two-armed goddesses.15

15

Illustrations in Snellgrove and Skorupski (1977:43 and Figure 25), Poell (2004: Figure 4), Di Mattia (2007:75–76 and Pls. 4, 5, 6).

207

The Wood Carvings of Lachuse

Panel heights (cm)

(capital) ?

(capital) ?

Panel heights (cm)

(?)

dvārāpala (?)

dvārāpala

29

38

Female standing 6 arms (?)

Female standing 2 arms

40

26

Male seated 2 arms

Male seated 2 arms

29

41

Female standing 2 arms

21+

Male 2 arms seated (?)

Male 2 arms seated

(empty base) ??

(empty base) ??

25 (?)

+ Panel partially cut Figure 8.13

Female standing (?) 25+ 2 arms (?)

Missing panels

Reconstruction of the pillar pair of the Lachuse portal.

(?)

25 (?)

208

Poell

However, there is some difference in the height of corresponding panels on the left and right, the design and the decorative details of corresponding niches on the left and right differ, and the figures on the right jamb are more slender and fill less space of the niches than on the left. Finally, the right pillar is about one to two cm wider and deeper than the left one. If the two pillars were indeed intended as a pair—and I do not see any other possibility—one would expect the same dimensions and overall design for corresponding elements. I would conclude that in all probability the two pillars were not carved by the same artist or workshop. While the door jambs of the Dukhang portal in Alchi have a similar layout and iconography, their carving style is very different from the Lachuse door. The standing figures of the two Lachuse pillars have no equivalents anywhere in Ladakhi art, and the seated figures are also quite different from the ones we see on the Dukhang door jambs. Also, the architectural niches in Alchi—both on the Sumtsek façade and on the Dukhang doors—are much less elaborate than at Lachuse. Only the dvārāpala figure on top of the right jamb at Lachuse is rather similar to the one on the Sumtsek façade. Reconstruction of the Portal Figure 8.14 attempts a reconstruction of the Lachuse portal on the basis of the extant fragments (lintel and jambs of the present portal, and a panel fragment from inside the temple). The reconstruction assumes that the door had a complete life story, shown in 24 or 26 scenes on two horizontal panels and two vertical pillars. As already mentioned, depictions of the Buddha life emerged in Western Tibet in the late 10th century. If we accept a mid-11th to late 12th century date for Lachuse (see the Conclusions below), complete pictorial life cycles should have been known to the Lachuse artists, as the Tholing, Khojar and Nako portals were then already extant.16 It is therefore reasonable to assume that the Lachuse door had a complete life story, and not only the present panel with selected scenes from the pre-Enlightenment period in the Buddha’s hagiography. A coherent depiction of the Buddha biography as recounted in the Tibetan sources17 requires about 25 scenes. The reconstruction proposed here can 16

17

The figures on the Nako life story panel have post-Gupta period heads and hairstyles that are quite similar to those in Lachuse. As such heads occur nowhere else in Ladakh, this might indicate some connection between the Nako and the Lachuse carvings. Tibetan biographies of the Buddha divide the life story into ‘Twelve Deeds’, starting with the Bodhisattva’s sojourn in the Tushita heaven and ending with the distribution of the

The Wood Carvings of Lachuse

Figure 8.14

209

Reconstruction of the Lachuse portal (© G. Kozicz).

accommodate 24 or 26 scenes as follows: seven on the present lintel (which probably had two more scenes in its original state; see note 5), six or seven on each of the two additional jambs, and five on an additional horizontal panel. This would be consistent with the other extant life cycles in woodcarvings in the Alchi Dukhang and at Nako.18

18

relics. For a succinct outline see Tucci (1988:134); an overview of the various episodes comprised in each of the ‘deeds’ is given in Csoma de Kőrös (1863). The Buddha life on the Khojar portal has 27 scenes (Neumann and Neumann 2008), while the Alchi Dukhang portal shows 25 scenes (Poell 2004); if, as I believe, the Nako portal had a complete life story, there were 24 or 26 scenes in a similar arrangement as in the Lachuse reconstruction: the seven presently visible scenes on the upper horizontal panel, 5 scenes on the lower horizontal panel (now completely eroded), and six or seven scenes on each of the outer doorjambs (of which only the topmost frame on the left jamb is presently discernible). (Cf. Ziegler 2008.)

210

Poell

Figure 8.15

Portal of the Chigtan temple (after De Filippi 1924).

Without this additional horizontal panel, there would have been 19 or 21 scenes, which would also provide enough space for a somewhat abbreviated life story. However, I think that the patrons of the Lachuse temple would have wanted to place the doctrinally important Enlightenment scene, and not the much less important Contests and Married Life, over the centre of the entrance.19 I assume therefore that the Lachuse portal had a second, smaller panel with the concluding life scenes (Maravijaya, First Sermon and Nirvana, preceded by one or two miracle scenes and/or the Descent from the Trayastrimsha Heaven); this panel would have been mounted below the extant panel. Models for this kind of layout in Ladakh and adjacent areas are the (no longer extant) door of a small temple at Chigtan (Figure 8.15), the carved doorframe at Nako, and the portal of the Alchi Dukhang.20 All these doors have an inner and an outer frame, two or three pairs of jambs, three lintels, and decorative panels between the lintels. Such a layout is clearly derived from North and Central Indian Gupta and post-Gupta period temple portals, of which the earliest examples can be seen in Ajanta and Deogarh (late 5th and 5th to 6th century, respectively). That design was not used in Kashmir, but can be frequently 19

20

This is the case both on the Khojar portal (see Neumann and Neumann 2008:Figure 1) and on the Alchi Dukhang door, and possibly also on the Tholing portal documented by G. Tucci (see Klimburg-Salter 1988b: Figure 1). See Francke (1914:100 and Pl. XLIII) and (De Filippi 1924:258) for Chigtan, Ziegler (2008) for Nako, and Khosla (1979:Figure 30) and Poell (2004:Figure 1a) for the Dukhang portal.

The Wood Carvings of Lachuse

211

seen in Himachal temples, such as on the portal of the small temple in Ribba (late 10th century) and in the wood art of Chamba, particularly in Brahmaur (8th to 9th century). Those creations in turn draw on the stone art of Himachal, as in the rock temples of Masrur (7th to 8th century). Ribba in particular was an important model for the development of Western Tibetan temple architecture, transferring the North Indian Gupta period stone portal into a woodcarving idiom.21 The single-figure jambs of this design were then converted into narrative panels, such as those carved on the Tholing and Khojar doors, which may have served as models for Nako and the Alchi Dukhang. In Ladakh only the doors of the Alchi Dukhang and, as far as it can be judged from the available poor photographs, the lost Chigtan temple have/had a similar layout of the main door jambs.22 The jambs of the later temple doors in Alchi—on the upper floor of the Sumtsek façade, and on the Lotsawa and Manjushri temples—and in Mangyu and Sumda have all a different decoration, such as geometric scrolls and/or repeated small Buddha figures. That puts the Lachuse doorframe into a certain timeline close to the Alchi Dukhang; the distinct differences and the somewhat closer connection of some elements with Kashmiri idioms might be taken as clues for dating the Lachuse jambs to a period before the major artwork in Alchi. The reconstruction proposed in Figure 8.14 is of course highly conjectural, and I am aware that it leaves open a number of questions, such as for instance the unclear iconographic role of the ferocious guardian figure on top of the extant (right) jamb. I would also conjecture that one of the now lost Lachuse lintels was probably showing the Five Jinas. This is suggested by a comparison with the Alchi Dukhang door, where an assembly of four Buddhas and four goddesses on the jambs is complemented by a horizontal panel with the Five Jinas, presumably in a reference to the Vajradhatu mandala.23 Finally, the panel with the kīrtimukha frieze was probably not part of the original doorframe, as the quality of the wood, the rather coarse carving and the width of the panel do not accord with the other pieces of the portal. A comparison with the Chigtan and Nako portals suggests that the piece could have been a reinforcement of the wall above the portal. 21 22

23

For Ribba (now sadly lost due to fire) see Klimburg-Salter (2002) and Di Mattia (2002). On the doorframe of the Wanla Sumtsek (bKra shis gsum brtsegs) there is a somewhat truncated Buddha life in 16 scenes, combined with geometric décor as on the later doorframes. See Luczanits (2002) for a description and dating of this monument. This convention could be traced to the Khojar portal, where all thirty-seven Buddhas of the Vajradhatu mandala are shown in a complex arrangement, interspersed with twelve goddesses. (See Neumann and Neumann 2008.)

212

Poell

Remains of the Façade and Architectural Pieces

Remains of the Façade The remains of the original façade construction of the temple (Figure 8.16) consist of two pairs of pillars, a pillar with sculptural decoration, a pilaster with scrollwork decoration, three short lintels, and a massive carved capital. These are clearly only fragments of the original façade, reassembled into a rather irregular arrangement between two ancient, but now broken, crossbeams. The upper crossbeam supports the roof and runs all across the front of the building, while the lower one now only extends from the left of the building to the right edge of the roof of the monks’ house. The two pillar pairs between the crossbeams carry each a short lintel that is decorated with simple garlands, similar to designs used in Alchi. All four pillars are carved with capitals and somewhat simpler bases; the only other decoration is a scroll under the capitals. This scroll is similar to the carvings on the façades in the Alchi complex and also on the balcony of the Wanla fort. The pillars are carved only on three sides, with the original back left plain; however, they have been positioned in such a way that the best preserved faces, originally their left sides, are now on the front.

Figure 8.16

Remains of the temple façade in Lachuse.

The Wood Carvings of Lachuse

Figure 8.17

213

Central sculptural panel of the Lachuse façade.

In the centre of the façade there is a massive pillar (Figure 8.17) with figural decoration on its front, and some decorative carvings around its sides24 indicating a capital and a base. The sculpture on the front shows a standing figure holding an object (a lotus stem?) in the left hand. The figure has rather exaggerated waist and hips, but seems to be a male one; the uṣṇīṣa on its head and the elongated earlobes without jewellery would identify it as the Buddha, but this is far from clear due to the decay of the wood. It is framed by an elaborate, Kashmiri–type architectural niche, with a double roof supported on single columns on the left and right. On top of the roof, there are an āmalaka and some additional decorations, but this is much abraded and now impossible to identify in detail. Stylistically the figure on this panel is similar to the standing figures on the left door jamb below. The piece seems a complete and self– contained composition, but it is of course possible that it is the fragment of a larger sculptural element. 24

As with the other façade elements, the back of the pillar is left undecorated.

214

Figure 8.18

Poell

Tripartite façade of the Alchi period (after Neuwirth 2009).

From these observations it seems clear that the temple had originally a tripartite façade of the type shown in Figure 8.18, similar to the three-storied temples in Alchi and Wanla, with three pairs of pillars inserted between decorated crossbeams which in turn were supported by two columns. The pillars were probably flanking decorated panels or wooden sculptures as can still be seen in Alchi. However, the Lachuse panel is stylistically not compatible with the Alchi-type façade and woodcarvings, as it does not follow Kashmiri models, but is more related to the woodcarving styles of the Brahmaur temple and other Chamba wood art. It seems that the façade at Lachuse was overall of a simpler design and execution than the one in Alchi; note in particular the absence of all decorative elements, such as wooden garlands, checkerboard panels, etc., that can be seen on the Sumtsek façade in Alchi. It is also noteworthy that the pillars and the sculptural panel have been left undecorated on their backs. In Alchi, façade elements are generally carved on all four sides. Two Loose Capitals Two large carved capitals have also been preserved at Lachuse, one on top of the monks’ house (seen in the lower right of Figure 8.16) where it now supports the sagging roof structure of the main building, and the other inside the prayer hall in front of the stupa niche (cf. Figure 8.2), where it serves as a low table for lamps and offerings.

The Wood Carvings of Lachuse

Figure 8.19

215

Sculptured face of the capital in the Lachuse façade.

Both blocks have a carved figure in an architectural niche on the two ‘short’ faces, while the ‘long’ faces are left undecorated except for a scroll or flower garland along the bottom. The bottom shows a stepped design that reduces the bottom cross section to a square (lost on the piece in the façade). On both blocks the sculptured faces are almost obliterated on one side, while the weather damage on the other side is much less severe. There is even less weather damage to the ‘long’ sides and to the (undecorated) top and bottom faces of the pieces. Figure 8.19 shows the best preserved of these figures. Its architectural niche consists of an elaborate scroll supported by square pillars with a groove on their front. The figure in the niche is crowned, wears earrings and a necklace, and sits on a lotus cushion with a cloth valance. One hand is in front of its chest, the other is slightly raised. To the left of the figure there is a rather prominent flower (or perhaps a conch on a staff). A garland (or streamers) falls down from the top of the niche, and there is also a mālā around the shoulders of the figure and on its seat. The figure is inclined forward and has a half lotus carved on its underside; this indicates that it was supposed to be viewed from below. On the

216

Poell

other side of the block there was a niche similar to the one on the front with a seated figure on a lotus throne, and with the same garland decoration on the left and right of the niche. The other capital has on its better preserved side a similar niche with the Buddha in bhūmisparśa mudrā, shown with a halo around the head, but without jewellery and other attributes. The opposite side of the block shows an almost obliterated figure seated in rājalīlā with one leg hanging down from the seat and with a lotus in its left hand. These figures are carved in high relievo, showing an accomplished line and carefully carved details, such as crowns and jewellery, and are set in elaborate frames of unusual design. Overall, these compositions are very close to classical Kashmiri sculpture, while the remaining decoration of the capitals is very similar to the column capitals inside the temple (see below), and in turn to the architectural decoration one can see in the Alchi chos ’khor. I think it is safe to assume that these two blocks were column capitals in the façade, but it is difficult to say what their exact function was. In their original position, the sides and the top and bottom faces were evidently protected from rain and snow, while the short faces must have been exposed to the elements, with one side suffering perceptibly more damage than the other. I do not think that the two pieces were capitals for single columns, because that would result in an awkward relation between the large capital and the relatively small column, and the presence of the mortises on the sides would then be difficult to explain. More likely, these blocks were the centre pieces of three-column arrangements similar to those on the façade of the Alchi Sumtsek or in the Wanla fort (Figure 8.20). The Column Capitals Inside the Temple Inside the temple the roof is supported by two undecorated main beams which rest each on two columns and on lion brackets that protrude from the walls. The two back columns have elaborate ancient capitals, while on the front one column carries a fragment of an ancient capital of the same design, and the other one has a modern, crudely carved replacement. Figure 8.21 shows the back left capital, which consists of a lotus ‘cushion’ and three rectangular blocks. The lowest block has scrollwork decoration around all sides, while the middle and top blocks have on their ‘long’ faces decorative scrollwork terminating in volutes; the ‘short’ faces of these two blocks have been left undecorated. On the top block a Kashmiri-type architectural niche frames a central figure on both sides of the capital. Figure 8.22 shows three of the figures in the central niches of the two complete capitals. Viewed from the entrance, the left capital has on its front the

The Wood Carvings of Lachuse

Figure 8.20

Three-column arrangements in Alchi (above) and Wanla (below).

217

218

Poell

Figure 8.21

Column capital in the Lachuse temple.

Figure 8.22

Central niches of the Lachuse column capitals.

Buddha in bhūmisparśa mudrā, clad in a monk’s robe leaving the right shoulder bare. Note the distinct uṣṇīṣa, the curly hairstyle, and the prominent halo behind the head of the figure. The seat is a double cushion without decorative carving. The niche framing this figure has a stepped, flat roof, supported by double columns with elaborate bases and capitals. The other capital shows on its front side roughly the same design, albeit with some slight variations: here the columns have a profiled outer rim, and the seat is a square pedestal on a lotus protruding from the niche. The star-shaped

The Wood Carvings of Lachuse

219

halo of the central figure is much more prominent than on the left capital. The figure on the back of this capital is shown in the rājalīlā (‘royal ease’) position, with the right hand raised and the left in its lap. These capitals are clearly of the same type as those that can be seen in the Alchi Dukhang and in the Manjushri and Lotsawa temples. However, the design here is more elaborate than in the Alchi temples, and the workmanship of these pieces is markedly superior to that of all other Ladakhi examples of this type; this pertains both to the complex and highly accomplished composition of the niches and central figures, and to the decorative carvings. In particular, the figures in the central niches have a sculpture-like quality, resembling Kashmiri bronzes in design and proportions.25 Note in particular how the knees of the Buddha figure protrude over the lotus seat—a typical feature of Kashmiri bronzes. Overall, these figures are far superior to the rather stylized (and sometimes carelessly outlined) relief figures on the Dukhang and Sumtsek capitals. We have thus the same situation as with the loose capitals described above: the overall design and the geometric decoration correspond to what one can see in Alchi and on related monuments, such as Mangyu and Sumda Chung, while the sculptural elements are much closer to Kashmiri sculptural styles than in Alchi. The Capital Over the Side Door The door to the monks’ house is decorated with an architectural piece of woodcarving (Figure 8.23); originally that was a pilaster capital, as can be seen by comparison with the façade of the Sumtsek and Manjushri temples in Alchi (Figure 8.24). The piece shows a seated figure, possibly Manjushri (if the slanting line behind the figure is taken for a sword) in the centre, set in a Kashmiri-style architectural niche with a complex trefoil arch with a garland on the inside and a decorative banner on the outside. The proportions and the carving style of both the figure and the niche differ from those of the carvings on the door jambs of the main entrance. At the bottom the capital has two oval side pieces with auxiliary figures on a lotus seat with raised hands. Exactly the same device is found on the pilaster capitals on the Sumtsek and Manjushri temples at Alchi. It is thus highly likely that this carving has been produced by the same workshop as the column capitals of the Alchi vestibules.

25

See the numerous examples in Pal (1975:Figures 24, 30, 39, 48 and others).

220

Poell

Figure 8.23

Pilaster capital over the entrance to the monks’ house.

Figure 8.24

Pilaster capital from the vestibule of the Lotsawa temple, Alchi.

The Wood Carvings of Lachuse

221

Conclusions Regarding the dating of the monument, I would agree with Luczanits (2005:80) that Lachuse seems to be a “predecessor” of the Alchi Sumtsek in terms of its wood architecture, though that statement requires some qualification in view of the distinct differences between the portal and the wood architecture of the vestibule and inside the temple. The façade in Lachuse is obviously closely related to the wood architecture in the Alchi complex. Its overall design follows the Sumtsek model, and individual pieces, such as the capital over the side door, the pilaster in the vestibule wall and the bracket lions inside the temple, are very similar to corresponding architectural and decorative elements in the Alchi temple complex. For the capital over the door to the monks’ house (Figure 8.23) the similarities are so close that the same workshop must be assumed. But for some other pieces (in particular the column capitals in the façade construction and inside the temple) a close examination reveals subtle, yet distinct differences; most importantly, the sculptural decoration of the Lachuse capitals is closely modelled after classical Buddhist sculpture of Kashmir—in fact, much closer than anything seen at Alchi or elsewhere in Ladakh. I would conclude from this that these parts of the Lachuse wood architecture illustrate an earlier stage in the fusion of Kashmiri styles with local innovations than the artwork at Alchi.26 The fact that the Lachuse façade was 26



The tripartite façade itself is also a fusion of Kashmiri elements with local innovations. The “presumed Kashmiri origins” (Luczanits 2005:82) of this type of façade was mentioned passim or was implicitly assumed by many authors (a.o. Khosla 1979:57–58; Snellgrove and Skorupski 1977:45; Goepper 1996b:91), but was never convincingly demonstrated. Kashmiri classical architecture knows peristyles, architectural niches as frames for doors and sculptural compositions, and trefoil arches for doorways, but no two-tier façades (cf. Fisher 1989). The wooden shrine of the British Museum ivory (see Figure 8.10) is a good example for the translation of a Kashmiri temple façade into wood architecture, but that was clearly not a model for the Sumtsek and Lachuse façades (though this style has certainly influenced the design of the temple portals in Mangyu). Depictions of Kashmiri architecture on the clay statues of the Sumtsek (see Goepper 1996a:46–47, 51, 55, 59–65) show several temples and palaces, but none of these could have been the models for a Sumtsek-type façade.  To my mind this indicates that tripartite temple façades are a Ladakhi innovation, responding to requirements resulting from the purpose of the buildings and the available building materials (stone, mud bricks, wood) and artisans’ skills. The design is certainly based on architectural elements from Kashmir, but these were combined with architectural devices and sculptural decoration that drew on models from other areas. Similarly to the development of the depiction of the Buddha life, those influences were

222

Poell

overall much simpler than the one of the Sumtsek may also point to an earlier date. Yet how much earlier is impossible to say in the absence of dated pieces or other supporting (epigraphical, historical) evidence; the simpler decoration might also have been the result of a more austere design and/or of a commission by less munificent patrons. And the ‘earlier’ column capitals might have been produced by a different (more conservative?) workshop working besides the one that produced the other, ‘later’ carvings. While this establishes some sort of relative chronology for Lachuse in its relationship with the art of Alchi, it is hardly possible to derive a specific date for the Lachuse façade and wood architecture from that argument. There is first the still unresolved question of the date of the Alchi Sumtsek (to which the Lachuse façade is most closely related). Snellgrove and Skorupski (1977:79–80) proposed an “11th / 12th century” date for the “oldest parts” of the Alchi Dukhang and Sumtsek. Goepper (1990) has—on the basis of the inscription and decoration on the topmost level—dated the Sumtsek (and also the “Great Stupa”, i.e. the entrance chorten to the chos ‘khor) to the period of 1200 to 1220, with the oldest temple on the site, the Dukhang, about half a century earlier; this argument is expanded and reinforced in Goepper (2003) and Luczanits (2003) by an analysis of the iconography of the lineage paintings. However, Denwood (in Snellgrove and Skorupski 1980:152, and in the present volume) “proposed, mainly on grounds of the palaeography, style and authorship of its inscriptions, that the three-storey Sumtsek temple alongside the Dukhang was founded by Tshul khrims ’od in the last third of the 11th century”, confirming thus the dating of Snellgrove and Skorupski (1977). A second complication results from the building history of the Alchi chos ’khor, which is much more complex than what the usually assumed straight timeline of Dukhang—Sumtsek / Great Stupa—Manjushri / Lotsawa— Lakhang Soma implies. Luczanits and Neuwirth (2010)27 have shown that all buildings in the Alchi chos ’khor have undergone extensive changes before and up to the 14th century, and it seems that this has resulted in different timelines for the architecture, for the wood art and architectural elements, and for

27

fused into a novel and unique architectural idiom. (See also Di Mattia 2007 for similar arguments for the development of Western Tibetan architecture in general. In regard of the Gupta period layout of temple portals Di Mattea writes (2007:69) that this “was absorbed, then mediated, elaborated and finally exported” by artists in Himachal to Western Tibet and Ladakh.) The paper presents a rather condensed summary of the building history of the Alchi chos ‘khor. The authors are preparing a comprehensive publication on the architecture and building history of early Ladakhi temples.

The Wood Carvings of Lachuse

223

the other artwork (painting, clay sculptures). That explains that the vestibule of the Lotsawa and Manjushri temples is very similar to the Sumtsek façade, while their interior is clearly later (probably mid- to late-13th century) than the decoration inside the Sumtsek. Moreover, the Dukhang originally had a façade and vestibule that was very similar to the Sumtsek,28 while the extant wood sculptures and architectural pieces in the vestibule are probably earlier than the wood art in the Sumtsek.29 This leaves the question of an absolute date for the Lachuse façade rather open; all we can say is that it dates to somewhere between the mid-11th and the late 12th century, depending upon which date—from late 11th to early 13th century—one accepts for the Alchi Sumtsek. The situation is different for the portal. The (reconstructed) layout of the Lachuse portal corresponds to the Chigtan and Dukhang portals in Ladakh, but its decoration has no connection with the doorframe of the Alchi Dukhang (which also has a Buddha life) or with the other extant doorframes in Ladakh (of the Alchi Sumtsek, Lotsawa and Manjushri temples, and in Mangyu, Sumda Chung and Wanla). The architectural niches at Lachuse are much closer to Kashmiri models than in Alchi, and the figures on the portal jambs are reminiscent of Himachal wood art such as in Brahmaur.30 For the depiction of the life story, no development line from the Lachuse portal to the carvings on the Dukhang doorframe in Alchi can be constructed; the rather static compositions of the life scenes at Lachuse are clearly not related to the animated designs on the Dukhang portal, where there are also no post-Gupta period heads and hairstyles. More plausible seems a connection with Nako in Kinnaur, where the doorframe of the dKar chung lha khang has a very similar layout and where we see also post-Gupta period hairstyles on the sculptural panels. The Lachuse portal appears thus as a combination of Kashmiri niches with sculptural styles from Himachal that were transmitted to Ladakh through Ribba and Nako. In view of the ‘purer’ Kashmiri influences and the stylistic elements from Himachal (based on 8th to 10th century examples), I would 28 29

30

Remains of which can still be seen high up on the wall to the left and right of the portal. I owe this information to Prof. Holger Neuwirth of Graz University. A solution to the vexed question of the dating of the Alchi complex is suggested by Snellgrove’s conjecture that the clay figures and interior decoration of the Sumtsek could be later than the (original) building and the exterior wood architecture (Snellgrove and Skorupski 1977:80). As already noted, the central sculptural panel in the Lachuse façade (Figure 8.17) belongs stylistically to the carvings on the doorframe, and has also no relations with the wood carving style of Alchi.

224

Poell

argue for an earlier date than the other wood art (i.e., the façade and architectural elements) in Lachuse, and thus also for a date prior to the Alchi Sumtsek. A comparison with the woodcarvings of the Dukhang portal indicates that Lachuse was probably also earlier than the Alchi Dukhang (or more precisely, earlier than the Dukhang portal, which might or might not be coeval with the building in its present state). This suggests an early 11th century date for the portal (i.e., earlier than the Dukhang portal in Alchi), but of course that ought to be treated as a conjecture as long as there is no corroborating evidence from an established timeline of comparison pieces. It seems thus that the wood art at Lachuse dates from two distinct periods, possibly covering a time from the early 11th to the late 12th century (where the upper time limit corresponds with the date of the Alchi Sumtsek); that might indicate a long period of construction, reuse of older wood pieces from a different site, or changes and additions after the completion of the original building. It is of course tempting to speculate about a refitting of an earlier building in the Alchi style—analogous to the changes that happened in Tabo about a hundred years after its founding31—, but without additional evidence that remains speculation. It is perfectly possible that the Lachuse portal was carved at the same time as the façade and the columns inside the temple, being the work of artists that worked in different, earlier traditions. At any rate, the wood art at Lachuse illustrates the complexity of the art history of Ladakh in the early mediaeval period, when Ladakhi artists had close links not only with Kashmir (resulting from the work of Kashmiri artists recruited by Rin chen bzang po for the construction of the temples he founded in Western Tibet and Ladakh in the early 11th century),32 but also with other, more distant regions in the Western Himalayan area. Of particular importance in this respect were Ribba in the Sutlej valley and Brahmaur in the Chamba region of present-day Himachal; there, influences from the North Indian plains and from Kashmir were fused into a distinct idiom, which in turn provided inspiration to the Lachuse artists. The often-quoted ‘Kashmiri connection’ of the art of Ladakh33 thus turns out once more as an oversimplification and should be replaced with a more nuanced picture: Ladakhi artists have been influenced by Kashmiri and non-Kashmiri styles and idioms, and the artistic

31 32 33

For the history of the Tabo complex, see Klimburg-Salter et al. (1997). See the biography of Rin chen bzang po in Snellgrove and Skorupski (1980:90–92). A phrase coined by P. Pal some 20 years back in his seminal paper “Kashmir and the Tibetan Connection”; see Pal (1989:117).

The Wood Carvings of Lachuse

225

production of this period has to be understood as local creations that were inspired by the art of several neighbouring regions and beyond. Acknowledgements I thank Christian Luczanits for bringing Lachuse first to my attention and for helping me with the identification of the Throwing the Elephant scene. I have also benefited from discussions with André Alexander of the Tibet Heritage Foundation/Leh Old Town Initiative and with Prof. Holger Neuwirth of the Technical University of Graz/Austria. Gerald Kozicz prepared the beautiful reconstruction drawing in Figure 8.14 and helped with the diagram in Figure 8.13. John Harrison has allowed me to use the plan in Figure 8.1 that he has prepared for the Achi Association, and Prof. Neuwirth provided the drawing in Figure 8.18. Prof. Rob Linrothe, Christian Luczanits, Christiane Kalantari, Verena Ziegler and Kurt Tropper gave valuable comments on particular issues and/or earlier drafts of this paper. I also want to thank Prof. Lo Bue for his comments on the dating of Lachuse, and for his excellent editorial advice. All photographs are by the author, except where otherwise stated. Pictures of Lachuse were taken in September 2004 and July 2009.

The mGon khang of dPe thub (Spituk):1 A Rare Example of 15th Century Tibetan Painting from Ladakh Chiara Bellini

Historical Background

According to the Chronicles of Ladakh,2 dPe thub (Spituk) monastery (Figure 9.1) was founded by ’Od lde, the ruler of Gu ge, in a Rat year in the 11th century, which in all likelihood corresponds to 1048 CE.3 It seems that the monastery subsequently went into decline. All that remains of this first monastery are the two caves known as Brag khung kha ba chen, ‘Snowy Caves’, which lie near the present monastery in the village on the far bank of the Indus, and the outer shell of the mGon khang. The contents of the Chronicles are also extremely useful in reconstructing the political history of Ladakh at this time: the founding by ’Od lde of a monastery in the middle of the region, very near Leh, which at the time was an important entrepôt, is a sign of the strong political influence exercised by the kingdom of Gu ge over Ladakh. The fortunes of the monastery of dPe thub were later restored under the ruler Grags ’bum lde, a builder of temples who was renowned for having contributed greatly to the spiritual renaissance of Ladakh after a period of complex political and religious fluctuations. His reign lasted roughly from 1450 to 1490 (Lo Bue 2007:184), and he therefore probably ordered dPe thub to be rebuilt in the late 15th century. The project was undertaken in commemoration of a visit to the kingdom of Ladakh made by two followers of Tsong kha pa in 1461.4

1 My warmest thanks go to my friends in the monastery of dPe thub, who kindly allowed me to photograph the paintings in the temple. 2 La dvags rgyal rabs, recorded from the 17th century onwards (Petech 1977:1). 3 As pointed out by Erberto Lo Bue (Lo Bue & Bellini, forthcoming). 4 Sending emissaries in his place must have been normal practice for Tsong kha pa. He is known to have been invited to China in 1408 by the third emperor of the Ming Dynasty, Chengzu (generally known as Yongle, from the name of his period), but once again declined the offer, sending a pupil in his stead (Snellgrove 1987:181).

© koninklijke brill nv, leiden, ���4 | doi ��.��63/9789004271807_��1

The mGon khang of dPe thub (Spituk)

227

Figure 9.1 The Monastery of dPe thub.

The 15th century in Ladakh was marked by the rise to political power of the dGe lugs pa, due above all to the actions of the charismatic figure of Shes rab bzang po (Figure 9.2), the master who was associated with the very arrival of the order in Ladakh.5 A native of Mar yul and follower of Tsong kha pa, he 5 On the top portion of the southwest wall of cave 2 at Sa spo la (Saspol) are depicted three historical figures: a master and two pupils. In the middle is portrayed the bla ma Shes rab bzang po, flanked on his right by his grandson and pupil rJe dPal ldan shes rab and on his left by slob dpon ’Do sde. Information regarding Shes rab bzang po is provided by the bKa’ gdams kyi rnam par thar pa bka’ gdams chos ’byung gsal ba’i sgron me written in 1494 by Kun dga’ rgyal mtshan, in the bKa’ gdams gsar rnying gi chos ’byung yid kyi mdzes rgyan, penned in 1529 by bSod nams grags pa, and in the Vaiḍūrya ser po, written in 1698 by the regent of the Fifth and Sixth Dalai Lamas, Sangs rgyas rgya mtsho (Petech 1977:178).  According to the sources mentioned above, he came from Ladakh and became a follower of Tsong kha pa before founding the sTag mo lha khang to the north of Khrig se (Thikse) and the monastery of the same name. He is portrayed in a statue in the hermitage of sTag mo, to which E. Lo Bue kindly drew my attention. He is depicted in a cave as a master accompanied by two pupils. One of these is rJe dPal ldan Shes rab, corresponding to rJe drung dPal ldan Shes rab, the nephew of Shes rab bzang po, who became the first dGe lugs pa abbot of the monastery of Phug tal in Zanskar (Snellgrove and Skorupski 1980:42).  The follower depicted next to Shes rab with the name of slob dpon ’Do sde, might be slob dpon mDo sde rin chen, who is said to have founded the dGe lugs pa monastery of Chos sde dKar rgyas, in Zangs dkar, and added extensions to Phug tal and dKar sha, where dPal ldan Shes rab was also active (Tucci 1971:485).

228

Bellini

Figure 9.2 The master Shes rab bzang po and his disciples, portrayed in cave No.2 at Sa spo la.

was given the title of byams sems (Thubstan Paldan 1997:28). He founded the sTag mo lha khang to the north of the village of Khrig se (Thikse), where his nephew rJe dPal ldan Shes rab grags pa, later founded the monastery of the same name,6 as well as some important dGe lugs pa temples in Ladakh and Zanskar. It seems fair to assume that the rebuilding of the monastery of dPe thub was prompted by a process of religious and political transformation triggered by the fame of Shes rab bzang po, exactly as had occurred in the case of some temples in Zanskar (Snellgrove and Skorupski 1980:42).

 The presence of Shes rab bzang po, dPal ldan Shes rab in cave 2 proves the paintings cannot have been executed before the beginning of the 15th century. The period when the caves were decorated perhaps coincided with a growing interest at court in the dGe lugs pa order (Petech 1977:170; Petech 1978:361). Another important clue is the presence of Vajrabhairava in cave 3. This is possibly the first example in Ladakh of a painted representation of this deity, the subject of particular veneration among the dGe lugs pa. A painting of Vajrabhairava dating from just after the one in Sa spo la can be seen in the small temple inside the chos ’khor of Ba sgo, which has been recently restored by a team of Indian experts and the Italian restorer Anna Triberti. 6 Petech maintains that the monastery was the seat of Kushok Bakula rin po che. In fact, his private seat was the monastery of gSang mkhar (Sangkar), which was subordinate to dPe thub (Snellgrove and Skorupski 1979:109).

The mGon khang of dPe thub (Spituk)

229

The available sources are rather discordant on this rebuilding of the monastery of dPe thub and its transfer to the dGe lugs pa order. This conflicting information is due to the fact that the sources come from locations which are very far apart and the authors pursue diverging aims. The main sources referring to the putative rebuilding of dPe thub are: the La dvags rgyal rabs, chronicles of Ladakh probably first compiled in the 17th century; the Vaiḍūrya ser po, a guide to the dGe lugs pa monasteries and their founders and subsequent abbots, compiled in around 1697 by Sangs rgyas rgya mtsho (Tucci 1971:475), the regent of the Fifth and Sixth Dalai Lamas;7 the bKa’ gdams gsar rnying gi chos ’byung yid kyi mdzes rgyan, a historical text on the development of the Karma pa and dGe lugs pa schools, compiled by bSod nams grags pa in 1529;8 and the oral tradition of Ladakh, also confirmed by The Guide of the Buddhist Monasteries and Royal Castles of Ladakh by the Ladakhi dGe lugs pa monk Thubstan Paldan (Thub btsan dpal ldan), whom interviewed in 2002 and 2004. Looking at this list, from which emerges quite clearly the geographical and chronological distance between the authors, it becomes possible to envisage the different purposes these texts might have had. Each furnishes the same or similar historical facts seen from a different point of view. This makes it difficult to reconstruct the history both of this monastery and of Ladakh as whole accurately. Examining the sources chronologically, to judge by what bSod nams grags pa writes in the bKa’ gdams gsar rnying, dPe thub is supposed to have been restored by Nam mkha’ ba or Nam mkha’ pa (Tucci 1971:485; Petech 1977:167), who is perhaps depicted in the caves of Sa spo la (Figure 9.3).9 This figure is likely to have been Nam mkha’ dpal ba (1373–1447), a bKa’ gdams pa master and direct disciple of Tsong kha pa, who lived in dGa’ ldan byang rtse and founded the monastery of dGa’ ldan grva tshang. The same source claims he also founded Klu dkyil (Likir), adding that Nam mkha’ ba became its abbot

7 The work’s full title is dPal mnyam med ri bo dga’ ldan pa’i bstan pa zwa ser cod pan ‘chan pa’i ring lus chos thams cad kyi rtsa ba gsal bar byed pa vai dûr ya ser po’i me long. The section of this text dealing with the founding of the dGe lugs pa temples in Western Tibet was translated by Tucci (1971:472–488). 8 Petech regards this text as particularly authoritative (Petech 1977:174). 9 In the corner between the southwest and northwest walls of cave 2 in Sa spo la are portrayed three masters: in the centre Nam mkha’ ga(. . .) bzang, to his right dBang phyug chos rje and to his left (. . .) bSod nams. The person who wrote the inscription, rather crudely, might have made a mistake in adding a ‘ga’ instead of a ‘ba’. If this were so, the central figure might be a master called Nam mkha’i ba (Petech 1977:167), who is mentioned in the bKa’ gdams gsar rnying gi chos ’byung yid kyi mdzes rgyan by bSod nams grags pa.

230

Bellini

Figure 9.3 The master Nam mkha’ ba and his disciples, portrayed in cave No.2 at Sa spo la.

after the master lHa dbang blo gros, who will be discussed below, although Nam mkha’ ba was the elder of the two.10 However, no other source confirms this information as far as the founding of dPe thub is concerned. The La dvags rgyal rabs says only that the monastery was rebuilt at the wish of King Grags ‘bum lde after two followers of Tsong kha pa visited the kingdom of Ladakh in 1461, without mentioning any specific name. The text of the Vaiḍūrya ser po is sometimes inaccurate and incomplete in spite of the fame and shrewdness of its author Sangs rgyas rgya mtsho, who was perhaps not particularly interested in what had really happened in Ladakh. At all events, in this work it is written that the person responsible for the ‘repairs’ to dPe thub was the above-mentioned gSang phu pa lHa dbang blo gros, who was the direct disciple of mKhas grub rje (1385–1438) (Tucci 1971:484). What is most interesting here is that the oral tradition of Ladakh, and especially that of Thub bstan dPal ldan (Thubstan Paldan 1997:14), corroborates this version, maintaining that dPe thub was refounded by lHa dbang

10

It seems that this information is backed up by an eighteenth-century inscription in Klu dkyil, which was copied by Francke (F.182) (Petech 1977:167).

The mGon khang of dPe thub (Spituk)

231

blo gros,11 supported first by Grags ’bum lde and later by Blo gros mchog ldan (Tucci 1971:485). lHa dbang blo gros seems a rather interesting figure. He was active in Ladakh and Zanskar in the 15th century and also restored the gSer khang in Tholing (Vitali 1990:38). These details are to be found in an inscription in the lower right-hand corner of the south wall of the dKyil khang in Tabo, where his name appears several times (Petech and Luczanits 1999:6; Vitali 1990:38). The west wall of the same temple in Tabo also has a portrait of him. The work of lHa dbang blo gros is associated with the endeavours of his older contemporary Ngag dbang grags pa, a direct disciple of Tsong kha pa and abbot of the chief monasteries in Gu ge: Tholing, rTsa hrang blo stang, Dung dkar and Mang nang. In 1424 he held the most important spiritual post in all Gu ge, presiding at the coronation of King Nam mkha’i dbang po phun tshogs lde. He built the new monastery of Tholing, with the help of lHa dbang blo gros (Vitali 1990:38). After following Tsong kha pa and his novel teachings, Ngag dbang grags pa, a devout missionary from mNga’ ris, made his way back to Western Tibet. His intention was to spread the teachings of the Tibetan reformer and found temples and monasteries to provide a firm base for the newly-formed order (Klimburg-Salter 1997:243). Thanks to the energy of this figure, aided by other masters and scholars like lHa dbang blo gros, Tholing became an important outpost of proto-dGe lugs pa in the 15th century. It was from here that monks set off to spread the teachings of their master Tsong kha pa, a process which occurred very quickly. Shes rab bzang po, a direct disciple of Tsong kha pa, had himself already begun this missionary activity in Ladakh. It is not hard to imagine the reasons why the two followers of Tsong kha pa decided to visit the kingdom (an episode mentioned in the Chronicles of Ladakh), in view of what was happening in Western Tibet during the 15th century—a period of great turmoil for the newly-established dGe lugs pa school. It is reasonable to suppose that one of the two disciples mentioned in the Chronicles might be lHa dbang blo gros, sent to Ladakh by Tholing. The paintings in the mGon khang probably belong to this stage in the development of the dGe lugs pa in Ladakh and are part of the refurbishing of the monastery complex. Their distinctiveness lies chiefly in their beauty: these works are exquisite examples of their style (Figure 9.4). Moreover, they are the sole examples of their kind in the whole kingdom of Ladakh. The style of these paintings stands 11

The name Nam mkha’ dpal ba appears in the Vaiḍūrya ser po ma, where he is described as the second abbot in dPe thub (Tucci 1971:484).

232

Bellini

Figure 9.4 Deity belonging to the cycle of ‘Direction Deities’, painted on the east wall of the mGon khang of dPe thub.

out immediately when compared with that of contemporary temples and monasteries, such as those in the lHa khang so ma, in A lci, those in the Sa spo la caves, those in the Gu ru lha khang, in Phyi dbang (Phyang), in the chos khor of Tsha tsha Phu ri, in A lci, those in a mchod rten in Nyar ma12 and those which are all distinguished by their coarser and less sophisticated style. Taking as a specific example the cycle of paintings in the Sa spo la caves, datable to around the late 15th century, the differences between the two styles are glaring and cannot be explained away as the result of gradual changes occurring over a decade or two. In all probability, groups of artists arrived in the area from the most important dGe lugs pa centres in Tibet, with the founding of the first dGe lugs pa monasteries in Ladakh. It is very likely that the founding of dGe lugs pa monasteries in Ladakh attracted groups of artists from more important dGe lugs pa centres in Tibet. The paintings in the mGon khang in dPe thub might thus have been executed by artists from Western Tibet. On close examination, certain paintings in the 12

To this list should be added the bCu gcig zhal, in Wanla, the Seng ge lha khang, in Lamayuru, and the lha khang in A lci Shang rong.

The mGon khang of dPe thub (Spituk)

233

Figure 9.5 Pictorial cycle painted on the north wall of the mGon khang of Khrig se.

Gu ge area, for instance those in the Vajrabhairava lHa khang, in the lHa khang dkar po and in the lHa khang dmar po in rTsa brang (cf. Tucci 1989), display similarities in style and iconography with the paintings in the mGon khang in dPe thub. This style in turn seems to betray the influence of models found in the sKu ’bum in Gyantse, where mKhas grub rje was involved in the iconographic layout.13 This is the master lHa dbang blo gros followed and it is highly probable that he travelled to Gyantse, where he would undoubtedly have viewed the paintings within the various buildings. Jamspal maintains that the very reason for building the Te’u ser po stupa in Leh was to erect a structure in imitation of the stupa in Gyantse (Jamspal 1997:142). Another example of this unusual style can be seen in the mGon khang in Khrig se, where a number of wonderful figures (Figure 9.5) can be made out in spite of the pervasive darkness and the layers of oily black soot deposited by centuries of burning candles. The temple belongs to the dGe lugs pa order and probably dates from the monastery’s foundation by Shes rab bzang po’s nephew, rJe dPal ldan Shes rab, portrayed together with his uncle in cave 2 in Sa spo la. In all likelihood, Shes rab bzang po also played a part in the foundation, which probably dates from the early 15th century, since it is known that in 13

mKhas grub rje helped Rab brtan Kun bzang found the dPal ’khor chos sde in Gyantse (Ricca and Lo Bue 1993:25).

234

Bellini

1447 certain beams in Nyar ma were used in the construction of the ’Du khang dKar po in Khrig se. The paintings in the mGon khang in dPe thub also display certain similarities with paintings executed inside the sTon pa lha khang in Phug tal, in Zanskar (Snellgrove and Skorupski 1980:52), whose first abbot was precisely dPal ldan Shes rab (Tucci 1971:485). The monastery of Phug tal is part of a group of monasteries won over to the dGe lugs pa order by the efforts and charisma of Shes rab bzang po. The resemblances among the paintings in these three sites (dPe thub, Khrig se and Phug tal) reveal the existence of a school of painting, very probably of Tibetan origin and associated with the dGe lugs pa order.

The Cycle of Paintings

Access to the rectangular temple is by a side entrance set into the south wall. On the west wall there are nine statues of Protectors of the Dharma, including Mahākāla, mGon dkar, the ‘White Protector’, the lCam sring group, ‘Brother and Sister’, Khyi tra, riding a dog, and dPal ldan lha mo on a mule. From these emerges the figure of Vajrabhairava, the largest of the statues present, which constitutes its parivāra. Two other statues, dPal ldan lha mo riding a mule and an attendant, are set in the corner between the west and north walls, opposite the entrance wall. The statues are framed by large halos of flames painted red and decorated with fine spirals painted in shades of the same colour, creating a chiaroscuro effect. The outlines of the halos have a serrated ‘bird’s crest’ appearance, originally an Indo-Newar stylistic feature found in several other temples in Ladakh.14 The leaping flames reach up to the ceiling, covering the wall behind the statues completely and lending a powerful atmosphere to the temple as a whole. The iconographic cycle is dedicated to Vajrabhairava, rDo rje ‘jigs byed, a deity worshipped with particular fervour by the dGe lugs pa. He is the god of death and the netherworld, as well as ‘Lord of the Law’ (Chos kyi rgyal po), the supreme judge who decides the fate of the deceased. He owes the devotion he attracts, especially among ordinary people, to his role as judge and executioner (Tucci 1989:77).

14

Halos of flames depicted in a similar fashion are present in the paintings in the Gyantse stupa (Ricca and Lo Bue 1993:166, Figure 51).

The mGon khang of dPe thub (Spituk)

235

Vajrabhairava is the terrifying emanation of Mañjuśrī,15 constituting the fusion of his benevolent and wrathful aspects (zhi khro). He symbolizes the ability to overcome and destroy all forces pitted against the Dharma and its practice and the power to eliminate anything standing in the way of a person achieving ultimate salvation. This characteristic in particular is embodied by Yamāntaka,16 an emanation of Vajrabhairava;17 the very name of this deity is the symbolic expression of the concept expressed above: he is the ‘Remover of the Obstacle Yama’, namely death (Māra, tib. bDud). Yamāntaka is regarded as a god of redemption. The iconographic programme echoes the cycle painted in the Vajrabhairava lha khang of rTsa brang.

The North Wall

There are five registers of paintings on the north wall, covering a variety of different images (Figure 9.6). The paintings cover three-quarters of the surface area of the wall, leaving a broad band completely image-free along the bottom. In the upper register are depicted seven masters in monk’s habits, who are part of the tantric cycle concerning the mandala of Mahāvajrabhairava. 15 16

17



In his triumphal manifestation, Vajrabhairava is represented with 34 arms, 16 legs and nine heads. The face above the other two central ones is the bodhisattva Mañjuśrī’s. Vajrabhairava’ is often mistaken for Yamāntaka, especially by the Tibetans themselves. Yamāntaka is in fact a manifestation of Vajrabhairava and has his own characteristics. He has a different iconography and does not have an animal aspect, but a human face instead. He is the symbolic representation of the perfect union between pairs which achieve the quintessence of tantric asceticism: ‘father and mother’, ‘meaning and mystic gnosis’, ‘sun and moon’, ‘iḍā and piṅgalā’ (Tucci 1989:84). A considerable tantric literature has developed on the subject of Vajrabhairava and can be broken down into three groups of Tantras, each distinct but closely correlated with the others. On the one hand there is the group devoted to Vajrabhairava, namely the Śrīvajramahābhairavatantra and the Śrīvajrabhairavakalpatantrarāja; on the other there is the series of Tantras dedicated to Yamāri: Kṛṣṇayamārisarvatathāgatakāya­ vākcittakṛṣnayamārināmatantra, Kṛṣṇayamārikarmasarvasiddhikaraṇanāmatantra and dPal gShin rje gshed nag po’i rgyud rtog pa gsum pa (Tucci 1989:78–79). All these ‘revelations’— the Tantras being esoteric revelations of the Buddha, according to Mahāyāna beliefs—are closely associated with one another.  The first master to spread the teachings regarding Vajrabhairava was Lalitavajra, who lived in around the tenth century; according to tradition, he brought the Mahāvajrabhairava from Uḍḍiyāna to Tibet, although philological analysis has shown that these texts were of Nepalese origin.

236

Bellini

Figure 9.6 The North wall of the mGon khang of dPe thub.

They embody the continuity of the initiatory teachings begun by Vajrabhairava himself. There are 16 masters of the lineage depicted here: some visible on the north wall—one seems to have been destroyed—and others on the south, diametrically opposite. According to tradition they should correspond to Vajrabhairava, Ye shes mkha’ ’gro ma, Lalitavajra, Amoghavajra, Jñānakaragupta, Padmavajra, Dīpaṅkararakṣita, rDo rje grags pa (the lo tsa ba of Rva), Chos rab of Rva, Ye shes seng ge of Rva, ’Bum seng ge of Rva, Rong pa Shes rab seng ge, bLa ma Ye shes dPal, Chos rje Don grub rin chen, Tsong kha pa and mKhas grub (Tucci 1989:111). Special care has been taken to convey the expressions on their faces, as if the artist was intent on portraying their individual features. Mutilated parts of the human body (legs, arms, heads, etc.) stand out against the dark background behind them. This gruesome sight, which includes predators, such as big cats and vultures, reproduces a typical cremation or so-called ‘sky funeral’ site, an iconographic subject typical of this Indo-Tibetan cultural context. The figures are heavily outlined in black and it is worth mentioning here that this is a feature found in other cycles of paintings in Ladakhi temples. It is

The mGon khang of dPe thub (Spituk)

237

already to be seen in the temple of Mang rgyu, for example, and is still present in the 16th century mGon khang in the rNam rgyal rtse mo in Leh.18 Above the figures of the masters, human and animal skins are arranged in the form of a sort of long canopy, a decorative element often present in subjects of this kind. The register beneath contains six dharmapāla with multiple arms, once again part of the cycle associated with the mandala of Vajrabhairava. The third and fourth registers are devoted to an iconographic group revolving around the figure of Yama,19 Lord of Death and the Netherworld. He is also known as ‘Lord of the Dharma’, dharmarāja in Sanskrit.20 Yama takes several different forms, each with its own epithet, but the most common are phyi sgrub, nang sgrub, or phyi nang gsang sgrub. Like many tantric deities, this dharmapāla also has ‘outer’, ‘inner’ and ‘secret’ forms. The dPe thub version is the ‘inner’ Yama, Nang sgrub Srin gi gdong can, with the human, wrathful, face of a rākṣaṇa. His attributes are a flaying knife and a skullcap. This variant of Yama is often portrayed together with four other emanations of the same divinity, namely Zhi ba’i gShin rje, rGyas pa’i gShin rje, dBang gi gShin rje and Drag gi gShin rje (De Nebesky-Wojkowitz 1996:83). As mentioned above, between the third and fourth registers there is a series of five Yamarāja, comprising Yama, the chief figure, in his ‘inner’ aspect with a human face, and four images of Yama with a bull’s head. A smaller figure portraying Yamī, Yama’s sister, is added to this group of five Yamarāja (Figure 9.7). The first figure in the Yamarāja cycle, running from west to east in the third register, rappresents gShin rje ser po (Pīta Yamarāja, ‘Yellow Yama’). Next to him is depicted gShin rje dmar po (Rakta Yamarāja, ‘Red Yama’), followed by gShin rje dkar po (Sita Yamarāja, ‘White Yama’), whose attributes are a doubleheaded hourglass drum and an arrow, and gShin rje sngon po (Nīla Yamarāja, ‘Blue Yama’), who holds a spear and a lasso. All are depicted with an erect penis, a feature typical of the emanations of Vajrabhairava, Yamāntaka and Yama, symbolizing the mystic bliss achieved through the alchemical union of bodily fluids, according to traditional Indian physiology. This supreme bliss is experienced the instant these fluids in the two channels either side of the spinal column (iḍā and piṅgalā) unite in the middle channel (susumṇā).

18 19 20

My thanks go to Christian Luczanits for his remarks regarding this stylistic feature. The correlation between Vajrabhairava and Yama is depicted graphically in the maṇḍala of Vajrabhairava, which faces south, where Yama’s kingdom reliably lies. At times he is also known by the epithet gShin rje chos kyi rgyal po.

238

Bellini

Figure 9.7 Yamī painted on the north wall of the mGon khang of dPe thub.

The four animal-faced aspects of Yama preside over the four points of the compass and are considered tutelary deities of the four chief magical actions associated with the tantric liturgy (Tucci 1989:85). The fourth register is partially occupied by the red halo of the two statues on this side of the temple, dPal ldan lha mo and one of her attendants. Cemeteries and cremation sites, with mutilated bodies being eaten by vultures, are painted over the remaining area. The next figure on the wall is the terrifying image of Yamī (Figure 9.7), with a human face and a martial stance, holding a kapāla in her left hand and a flaying knife in her right. Her long red hair, framed by a diadem made of human skulls, is done up in a bun, with just a few flickering, flame-like locks loose. The goddess’s body is adorned by various types of jewels: earrings, anklets, garlands and a long necklace of freshly hacked-off human heads reaching down to her ankles. In spite of her aggressive appearance and bulging muscles, she still retains a certain femininity, highlighted by her slim waist and full, round breasts. A long piece of cloth billows on Yamī’s shoulders as if tossed by an imaginary wind swirling about her body. Her wrathful expression is made all the fiercer by the third eye placed vertically between her flaming

The mGon khang of dPe thub (Spituk)

239

eyebrows and by her fleshy red lips drawn back to reveal her sharp teeth set in a snarl (Figure 9.7). Next to Yamī is painted the most important image on the wall, the large figure of Yama gShin rje Nang grub (Antarasādhana Yamarāja), coloured blue and with red hair and wrathful expression. He wields the same attributes as his sister and assumes the same stance. His face is more carefully detailed, making this image if anything even fiercer and more terrifying than the previous one. Nag sgrub is the tutelary deity of Tsong kha pa and might therefore have been depicted in the mGon khang of dPe thub to reinforce the commemorative purpose of the rebuilding of the temple. gShin rje Nang grub dances over a remarkably well-observed corpse, in which the expression of agony is very skilfully conveyed (Figure 9.8). The detailing in this figure in general is of a notably high standard. For instance, the hair has been painted strand by strand with an extremely fine brush. Figures with a human and animal aspect, with long hair and fierce expressions, are depicted horizontally under the corpse. A similar group is visible also at rTsa brang where the figures are identified by captions written under them. These are Yama’s acolytes going about their funerary operations with their spears, staffs and other weapons of the period. A red, bull-faced figure wields a trident, while a black crow with outspread wings carries a female figure (Figure 9.9), which is found also at rTsa brang, where the inscription identifies it as ‘Bebs ma (Tucci 1989:93). Another female figure, with the face of an animal and long dark hair, dances next to another with a beard and moustache bristling with flames, hair standing and a fierce expression (Figure 9.10). An ash-covered demon holds a big white sack in his hand, while the figure by his side plays with a ball of thread. These weird and grotesque creatures are swathed in plumes of smoke and surrounded by swollen rivers of blood flowing among them (Figure 9.11). The pleasure the artist derived in painting these particular scenes is palpable.

The East Wall

The east wall is divided up into registers in the same way as the previous, with the upper register containing a row of thang kas. The yi dam cycle beneath is considerably blackened and damaged, making it hard to achieve an accurate iconographical reading, and is interspersed with images of Herukas. Each Heruka is surrounded by a halo of red flames, similar

240

Bellini

Figure 9.8 An “enemy of the Doctrine” painted on the north wall of the mGon khang of dPe thub.

Figure 9.9 Acolytes of Yama. North wall of the mGon khang of dPe thub.

The mGon khang of dPe thub (Spituk)

Figure 9.10

Acolytes of Yama. North wall of the mGon khang of dPe thub.

Figure 9.11

Acolytes of Yama. North wall of the mGon khang of dPe thub.

241

242

Bellini

Figure 9.12

Vajradhāra. East wall of the mGon khang of dPe thub.

in style to those behind the statues, creating a harmonious continuity between painting and sculpture.21 The black background to the figures is decorated with human skeletons depicted with quick strokes of white paint. The third register shows four finely drawn peaceful figures. The first is Vajradhara (Figure 9.12), who holds two vajras across his chest in simulation of an embrace with his tantric consort. The second is a particularly elegant figure seated with one knee raised (Figure 9.13). Her sinuous body is showed 21

This interplay can also be seen in the gSum brtsegs in A lci or in the Ta bo monastery, where painting and sculpture are perfectly matched, and indeed blend seamlessly into one another (as in the paintings on the dhoṭīs of the statues in the gSum brtsegs in A lci and those on the statues in the Śākyamuni and Maitreya chapels in Mang rgyu).

The mGon khang of dPe thub (Spituk)

Figure 9.13

243

Deity. East wall of the mGon khang of dPe thub.

up by her loose red trousers, which leave her belly uncovered, displaying her slim waist. Her delicately-featured face is seen in profile and her long hair is partly arranged loosely on her shoulders and partly done up in a small bun on the nape of her neck. This figure is followed by Nāro Ḍākinī (Figure 9.14), depicted with a wrathful expression and in her typical iconographic pose, as if performing a dance step. In her left hand she holds a kapāla full of the blood and fat of enemies of the doctrine, which she raises towards her lips. In her right she holds the ritual knife used for the symbolic flaying of the enemies of the doctrine, while a khaṭvāṅga,22 a trident with three heads at different stages of decomposition impaled on the end, rests in the crook of her arm. This image is notable for its sexually explicit details, including muscular legs, with full thighs and fine ankles, large round breasts, a huge sexual organ— symbolizing fertility—and supple hips. The fourth figure is equally beautiful, being in the same pose and performing the same gestures as the second. The draughtsmanship in these images is especially admirable. The fine line seems deliberately to highlight the litheness 22

These particularly macabre objects and symbolic instruments originated in Indian origin and are associated with the worship of Śiva.

244

Figure 9.14

Bellini

Nāro Ḍākinī. East wall of the mGon khang of dPe thub.

of these gorgeous female deities and to convey the eroticism of their rounded bellies and the sensuality of their gestures. The accurately drawn light and dark Herukas which follow the peaceful figures might correspond to Krodharāja sMe brtsegs, a special form of Jambhala (Tucci 1989:104–105). This protector exists in both white and green versions and is normally invoked during purification rites performed with consecrated water. These rituals generally act as exorcisms and are carried out in order to obtain relief from illness or infection or to counteract the harmful influence of demons (gnod) (Tucci 1989:106). On the other hand, the two protectors could be the two versions of the six-armed Mahākāla: Nātha Mahākāla and Sita Mahākāla. The sixteenth figure in this register does indeed portray Nātha Mahākāla, whose face is shown in three-quarter profile, with three heads and six arms and red hair held in place by a diadem of human skulls. The short beard of flames rendered along the jaw line with small orange curls immediately catches the eye. On this same row can also be seen a representation of dPal ldan lha mo, Rematī, astride a mule (Figure 9.15). She is depicted in her secret manifestation,

The mGon khang of dPe thub (Spituk)

Figure 9.15

245

dPal ldan lha mo. East wall of the mGon khang of dPe thub.

whose symbolic meaning is only accessible through the profoundest meditation. dPal ldan lha mo is the Buddhist version of Kālī, the ‘Great Mother’, the wrathful aspect of Śiva’s consort, symbolizing the ceaseless cycle of life and death. According to Indian demonology, Rematī was originally an ogress haunting the lives of children. In the earliest and most renowned Indian treatise on

246

Bellini

paediatrics, the Kumāratantra, her influence became active on the seventh day of the seventh month of the seventh year of a child’s life, and gave rise to bouts of high fever. When she began to exercise her power, the child would begin to display clear signs of illness: a general weakness of the body and a loss of hair, appetite and voice. Then, after the ogress was converted and incorporated in the Buddhist pantheon, she adopted these features as part of her iconography (Tucci 1989:97). All these details of unequivocal Indian origin are clearly to be seen in the Rematī painted in the mGon khang in dPe thub. The artist has obviously attempted to depict the extremely emaciated body as realistically as possible, dwelling on the muscle sinews and protruding bones. The deity’s dark skin—traditionally painted blue—is covered with the ash of cremated human bodies. The wrathful face includes a third, horizontal eye between the flaming eyebrows. The red hair is framed by a diadem of human skeletons, from which a few curls rendered with the tip of a fine brush break free and hang down over her broad forehead. Rematī traditionally wears a dhoṭī made from a tiger skin, while in this case she is actually wearing a short skirt decorated with a pattern of interlaced stripes and a snake belt. The mule she is riding is depicting at a walk and the realism of the representation reveals the eye of the naturalist in the artist and his knowledge of anatomy, a skill frequently repeated in the depiction of birds, dogs and other animals. The fourth register contains sixteen images of exceptional beauty, two of which are covered by a cupboard, painted within elegant, trefoil-arched niches with a red background. These figures belong to the cycle of ‘Direction Deities’, which includes the deities in the Hindu pantheon. The image of Brahmā (Tib. Tshangs pa) (Figure 9.16) opening the cycle shows him with three heads and four arms seated on a splendid, finely-detailed peacock. The middle face is peaceful and benevolent, with long eyebrows and finely outlined almond-shaped eyes. The small, even lips are poised in a slight smile. The other two faces are exact opposites of one another: one is explicitly feminine the one, while the other is plainly male. His bottom right hand draws a small noose to his chest, while the other points to the sky; the bottom left hand holds a rosary, while the other seems to be tendering a lotus flower out of the frame. Indeed, the originality of the composition lies precisely in the way the fingers of this hand extend beyond the edge of the niche holding this image of Brahmā. The next deity, Vemacitrin, (Tib. Thags bzang ris), is partially covered. The face is damaged, although its beautiful oval shape and slightly parted red lips can still be admired, and the detailing in general has almost disappeared. The hair is drawn back in a bun on the nape of the neck, leaving the broad forehead free except for a few delightful locks that have broken loose, framing the face.

The mGon khang of dPe thub (Spituk)

Figure 9.16

247

Brahmā. East wall of the mGon khang of dPe thub.

Thags bzang ris sits on a throne supported by two large wheels, from one of which, bizarrely, protrudes a horse’s tail. The third figure, Sita Candra, is portrayed with her eyes half-closed and turned downwards in ecstatic meditation. Her beautiful red lips are parted in a delicate smile. The deity mimes the gesture of stretching a thread across her

248

Bellini

Figure 9.17

The “enemy of the Doctrine”. Detail of painting on the east wall of the mGon khang of dPe thub.

chest, holding a long-stemmed lotus flower in her right hand. She is seated cross-legged on a lotus bloom and wears loose, emerald-green trousers draped around her in heavy pleats. The fourth figure, Nirṛti, is the first of only two wrathful figures in this cycle. He holds a flaying knife and a kapāla in his hands.23 His bright red hair is done up in a bun adorned with a diadem of human skeletons. He is shown riding an enemy of the doctrine, although he actually looks as though he is dancing over him (Figure 9.17). The expressiveness of the face and precise draughtsmanship make this secondary figure perhaps one of the most beautiful in the entire cycle of paintings. Nirṛti appears to step out of the frame, with one foot alighting on the script in lan tsa characters running along the edge of the wall. Śiva as Īśāna (Tib. gNod sbyin dbang ldan), the fifth figure, is depicted embracing Pārvatī (Figure 9.18), who is considerably smaller and painted blue. In his right hand he wields a trident and in his left, held low, he holds a kapāla, while encircling his consort’s waist with his arm. The pair are sitting astride a red bull. 23

Given the attributes, the figure depicted might be Yama, sometimes also present in this cycle (Chandra 1999:173).

The mGon khang of dPe thub (Spituk)

Figure 9.18

249

Śiva and Pārvatī. East wall of the mGon khang of dPe thub.

The sixth figure to be represented is Pṛthivī Devī, (Tib. Sa’i lha mo). The central portion of this painting is badly damaged and it is impossible to make out any attributes or the gestures of the hands. The seventh figure, with a wrathful expression and red hair unusually gathered up at the top of the head, wields a sword while riding a white bull. He wears a short red dhoṭī and a crown of human skulls. The face of Indra (Tib. brGya byin), the eighth figure, is damaged. He is riding an elephant, whose rather strange appearance suggests that although the

250

Figure 9.19

Bellini

Sita Varuṇa. East wall of the mGon khang of dPe thub.

artist had a good knowledge of animal anatomy, he had never actually seen an elephant, inevitably leading to the deduction that he was not of Indian origin. The ninth image depicts the four-armed Viṣṇu, shown riding Garuḍa and holding his sceptre in one of his middle hands. Viṣṇu is followed by Sita Varuṇa (Tib. Chu la dkar po), Lord of the Nāgas (Figure 9.19). This exceptionally beautiful figure holds a sort of snake lasso in his hand, while riding a sea serpent. The dynamic intensity of the whole composition is remarkable: the legs gripping the flanks of the sea serpent are

The mGon khang of dPe thub (Spituk)

251

turned to the left, the arms move to the right, while the god turns his sultry gaze again to the left. Indeed, the handling of the face is especially exquisite. The eyes are sharply defined, the evenly curved lips slightly parted in the merest smile, and the perfectly straight nose ends in small, graceful nostrils. The result conveys an aesthetic sensibility completely new and unique in Ladakh. The cycle also includes a representation of Sita Gaṇapati (Tib. Thsogs dbag dkar po), a manifestation of Gaṇeśa converted to Buddhism. Although the twelfth figure is not immediately recognizable owing to the severe damage to its face, it is clear from the red skin, hand gestures and the fact that his mount is a yak that this is a representation of Agni (Tib. Me lha). The thirteenth figure is Vāyu, (Tib. Rlung lha), who is depicted holding a long lasso in her hands. He wears blue trousers with red edging and a white belt. She is depicted riding a goat, whose head is turned towards her as he turns her gaze upwards, in a sort of ecstatic rapture (Figure 9.20). Sūrya, the ‘Sun’, the fourteenth figure, is seriously damaged. His right hand is turned towards his lap and holds a lotus, while the left hand is in the gesture of abhaya mudrā, warding off danger. The figure is seated on six horses’ heads, arranged in an original way to symbolize his chariot. This group of deities has been painted with particular care compared with other similar iconographic cycles, such as the one found in the mGon khang of Phyi dbang or in the temple of Maitreya in Ba sgo. The figures are depicted in a variety of poses and performing different gestures and include the features, such as facial expressions, clothing and hairstyles, that make each unique. There is a clear attempt at providing an individual ‘portrait’ of the deity represented. Even the halos that surround them are of different colours.

The South Wall

The south wall, in which the entrance is set, is undoubtedly the one most seriously damaged: the plaster is in a terrible condition and the paintings are severely blackened by the smoke from oil lamps. Two sections of the wall have been rebuilt and plastered and are thus undecorated. This means the iconographic programme is incomplete, although there are enough details remaining to be able to reconstruct the full cycle with some confidence. In the western portion of the upper register can be seen mirror images of masters wearing monks’ habits found on the opposite north wall, completing the cycle of the Mahāvajrabhairavatantra lineage. The unusual appearance of these figures’ faces makes them extremely interesting: some have slightly pop eyes and wrathful expressions.

252

Figure 9.20

Bellini

Vāyu. East wall of the mGon khang of dPe thub.

In the register below are represented four figures on horseback divided into pairs, one beneath the other (Figure 9.21). Temples, both in Ladakh and in Tibet, frequently contain images of horsemen in which the symbolic significance can vary from one to the other. Each of these four figures is depicted with his own individual features and clothing, and thus enjoys a distinct identity. They all carry arms, whether spears or bows and arrows, and are dressed in what are probably the fashionable Tibetan clothing and headwear of the period. The knots in their

The mGon khang of dPe thub (Spituk)

Figure 9.21

253

The “Eight Masters on Horseback”. South wall of the mGon khang of dPe thub.

horses’ tails symbolize royalty and are quite common features in this type of image. There seems to be an equal number of horsemen, albeit considerably damaged, on the opposite side of the same wall. The full set constitutes the parivāra of Vaiśravaṇa, known as rTa dbag brgyad, the ‘Eight Masters on Horseback’ (De Nebesky-Wojkowitz 1996:69). An enormous Tibetan lion facing its rider adorns the entrance door. Vaiśravaṇa, so damaged as to render him virtually unrecognizable, is the figure directly above. The only clues to his identity are the three features that can still be made out: Tibetan boots, the treatment of the weave of the clothing making it resemble the chain mail of a suit of armour, and a kind of pennant rippling in the wind like a victory standard. It is common to find one of the Four Guardian Kings, particularly Vaiśravaṇa, at the entrance to a temple. He protects the north and guards the treasures. Vaiśravaṇa, rNam thos sras or simply rNam sras in Tibetan, is also known as Kubera or Jambhala (De NebeskyWojkowitz 1996:68). To Vaiśravaṇa’s right there are a number of badly damaged protectors with multiple arms. The register below is also lost—a peaceful, smiling face being all that remains.

Chigtan Castle and Mosque: A Preliminary Historical and Architectural Analysis Kacho Mumtaz Ali Khan, John Bray, Quentin Devers and Martin Vernier Until the early 19th century Chigtan (Cig gtan or Cig ldan) in the Purig (Pu rig/ Bu rig) region of western Ladakh was a semi-independent principality ruled by a local chief or Jo. Like the rest of Purig, it was originally Buddhist, but the ruling family adopted Islam in the late 16th or early 17th century CE. The region’s oral tradition and its architectural inheritance are therefore of particular interest as a meeting point of Buddhist and Islamic cultural influences. Today, Chigtan has long since been eclipsed as a regional centre by the modern town of Kargil. However, it is still widely known as a focal point of Ladakhi culture, famous for the local version of the Kesar epic and for a wider repertoire of folk stories and songs.1 Chigtan castle is now sadly dilapidated but, even in its present ruinous state, it remains a particularly important example of Ladakhi architecture. The mosque below the castle has hitherto been overlooked almost completely by scholars but, as will be seen below, it contains an important set of wood carvings. The whole complex deserves to be better known. In two earlier articles Neil Howard (1989, 1995) presented a survey of fortresses scattered across the Indus valley and Zangskar, but he was not able to include Chigtan. This essay expands on this earlier work by presenting a preliminary study of Chigtan’s history and the remains of the castle’s architecture. It draws on photographs and oral evidence collected in the early 20th century and more recently, together with a preliminary inspection of the site as it stands today.

The Chigtan Foundation Myth

Chigtan’s foundation myth was recorded by the Moravian missionary scholar A.H. Francke in the early 20th century, and is still part of the region’s oral tradition today. 1 Herrmann (1991) includes a German rendering of the saga as recited by Rahimulla Takarpa from Chigtan.

© koninklijke brill nv, leiden, ���4 | doi ��.��63/9789004271807_��2

Chigtan Castle and Mosque

255

Francke (1926:172–175) reports than in 1906 he heard that Chigtan had had its own historical chronicle. It emerged that the original document no longer survived but that the former Jo, who now lived in Kargil, had learnt it by heart and could recite it from memory. Francke duly sent one of his assistants to record the text. According to the chronicle, the founder of the Chigtan principality was called Tsangkhan Malik (lTsang mkhan ma lig), and originally came from Gilgit in what is now northern Pakistan. He carried a walnut stick and planted it in the village of Dargo (Dar go) before going to sleep. When he awoke, the stick had sprouted leaves and, taking this as a good omen, he built a castle there. He had a similar experience in Kuksho (Kug sho), this time with a birch stick, before finally coming to the nearby village of Chigtan. He also built a castle there although, as Francke (1926:174) remarks, this must have been a distant predecessor to the building whose ruins survive today. Francke comments that the founder’s original name was probably simply ‘Tsangkhan’, and that ‘Malik’ might have been added later to make his name sound more Islamic. He records a similar foundation myth for the rulers of Sod (bSod) near Kargil who belonged to the same family and also held a small castle at Pashkyum (Pas kyum/Pas skyum).

Figure 10.1

Chigtan castle in 1909. Photo: Babu Pindi Lal. Courtesy Kern Institute, Leiden University, ms nr.XI. fol.31.

256

Khan, Bray, Devers and Vernier

Figure 10.2



Chigtan castle in 2009. Photo: Kacho Mumtaz Ali Khan.

Tsering Malik

According to local tradition,2 the present castle was built on the orders of Tsering Malik (Tshe ring ma lig), who ruled Chigtan at the turn of the 16th and 17th centuries. His hybrid Buddhist/Muslim name suggests that he may have been one of the first members of the ruling dynasty to embrace Islam, although it is likely that many of the ordinary inhabitants of the region would have remained Buddhist for some generations to come. Even today there is still one Buddhist household in the village, and hybrid Buddhist/Muslim names have been common in the nearby village of Kuksho until recent times (Sheikh 1995:164). The early 20th century historian Hashmatullah Khan (1939) states that Tsering Malik was the younger son of Kho Kho Bazam, the ruler of the combined principality of Sod, Pashkyum and Chigtan.3 Kho Kho Bazam’s elder son 2 The references to the oral tradition in this paper area are as recorded by Kacho Mumtaz Ali Khan. 3 Hashmatullah Khan served in Ladakh as an official of the Jammu & Kashmir government in the early 20th century. His work draws heavily on oral traditions, many of which have since been lost. We are grateful to Abdul Ghani Sheikh for this reference.

Chigtan Castle and Mosque

257

Urgyal Malik (Ur rgyal ma lig?) was the heir apparent, but he appointed Tsering Malik as the governor of Chigtan. Tsering Malik declared independence with the help of his father-in-law, the Ladakhi King Jamyang Namgyal (’Jam dbyangs rnam rgyal, r. c. 1595–1616), and captured Pashkyum as well as Chigtan, while his brother retained control of Sod. Tsering Malik’s construction—or reconstruction—of the castle may have been intended to provide him with a secure base when making this bid for greater regional power. A somewhat similar account—though differing in important details— appears in the La dvags rgyal rabs, the Ladakh royal chronicle (Francke 1926:106–107; Petech 1977:33–34). This records that Jamyang Namgyal took Tsering Malik’s side in a conflict with an unnamed neighbouring chief, possibly Urgyal Malik.4 Jamyang Namgyal’s decision proved disastrous because it led to the intervention of Ali Mir, the Makpon (dMag dpon) of Skardu in Baltistan, who defeated him in battle and overran the whole of Ladakh. Jamyang Namgyal was captured and taken to Skardu where, according to legend, he fell in love with the Makpon’s daughter Gyal Khatun (rGyal Khatun). Later he was allowed to marry the princess and to return peacefully to Ladakh.5 Tsering Malik presumably retained or resumed control of Chigtan, although the chronicle does not state this specifically. Tsering Malik also appears in two historical songs collected by Francke (1909:64–66). The first of these refers to the castle and in Francke’s translation runs as follows: On the boundary of heaven and earth, There is a castle raised by [a] lion If you ask where that is, where that is. It is the youths of middle age in our godly land. If you ask where that is, where that is. It is all the gravel-plains of the beautiful [castle] Shag-mkhar. Let us dance, let us laugh, O companions. Cry out ‘bravo’ before our good lord. Call out ‘bravo’ before our good Lord Tshering malig. The second of the two songs refers to an alliance between Tsering Malik and a king called Dzeldan Namgyal (mDzes ldan rnam rgyal). ‘Dzeldan’ means 4 Francke suggests that the opposing chief was Khri Sultan of Kartse (dKar tse) in the Suru valley. 5 For a recent study of intermarriages between the ruling families of Ladakh and Baltistan see Halkias (2011).

258

Khan, Bray, Devers and Vernier

‘possessing beauty’, and Francke (1909:66) suggests that it is an epithet for Jamyang Namgyal.

Local Legends Concerning the Castle’s Construction

The castle formerly had nine storeys and according to legend was crowned by a revolving wooden cabin at the top. Its main architect is said to have been Chandan, a highly skilled mason and carpenter from Chorbat in Baltistan. Tradition relates that, once the castle was completed, Tsering Malik’s councillors advised him to chop off Chandan’s hands so that he could not build a similar castle elsewhere. When Chandan heard of the plot, he removed some vital working parts of the revolving cabin so that it tilted to one side. He explained that he could not rectify the problem without a special tool that he had left at home, and he asked Tsering Malik to send 20 young men to fetch it. Tsering Malik duly sent the young men, and they carried a message from Chandan to his daughter-in-law, a particularly intelligent woman whom Chandan had selected after she had passed a series of tests. The message was: “The sky is clear, but there are only a few stars. It is a full moon, but the moon is only half.” The first part of the message meant that Chandan was in captivity, and was receiving only a thin broth by way of nourishment, without any meat or fat, and the second part meant that he was only receiving a half piece of bread. When Chandan’s daughter-in-law heard the message, she understood that he was in trouble. She asked the youths to fell some poplar trees in order to make the special tool, and managed to get them to stand in such a position that they were all trapped when the trees came down. She then freed one of them and told him to return to Chigtan with a message that the remaining 19 would all be killed unless Tsering Malik released her father-in-law with due honour and reward. Tsering Malik duly complied. Abdul Ghani Sheikh (2005:31–43) notes that similar legend is told in Leh, where the construction of the palace is likewise attributed to a Balti mason called Chandan. According to the Leh version, King Sengge Namgyal (Seng ge rnam rgyal, r. 1616–1642) chopped off Chandan’s right hand with a view to ending his architectural career. However, it seems that even this did not deter him because the legend relates he used his left hand to build further palaces at Hanle (Wam le) and Shey (Shel) in Ladakh and at Rudok (Rud hogs) in what is now Western Tibet.

Chigtan Castle and Mosque



259

Later History

Conflicts between Ladakh and Baltistan, as well as dynastic rivalries within Purig itself, were recurrent features of the region’s history in the 17th and 18th centuries (see Schuh 2008). Since Chigtan lies between the two main regional centres of power in Leh and Skardu, it is easy to imagine that it was caught up in these conflicts on more than one occasion, and that the defensive fortifications that we can still see in the castle were essential in the fluid and often violent political environment of the time. However, there are few documentary records. One of the few literary references that survives comes from the biography of Stagtshang Raspa (sTag tshang ras pa 1574–1651), the founder of Hemis monastery (Petech 1977:49). According to this text, King Sengge Namgyal invaded Chigtan soon after coming to the throne in succession to his father Jamyang Namgyal. However, the attack was unsuccessful: the Ladakhi leader Gaga Tenpa (Ga ga bTsan pa) was captured with some eighty men, while the nephew and niece of the Chigtan Jo were held hostage in Ladakh. Stagtshang Raspa was asked to mediate, and negotiated an exchange of the captives and a one-year truce. Apparently drawing on the oral tradition, Hashmatullah Khan records what appears to be a separate but related incident from the same period. It seems that Sultan Malik, who had succeeded Urgyal Malik as ruler of Sod, invaded Chigtan, and killed the now elderly Tsering Malik and his son in battle. He then annexed Chigtan and imprisoned Tsering Malik’s grandsons, Adam Malik and Chhozang (Chos bzang?) Malik. The majority of the people of Chigtan were Buddhist and opposed the rule of Sultan Malik: the implication is that he was more ardent in his promotion of Islam than Tsering Malik had been. They therefore appealed to Sengge Namgyal for assistance, but he was preoccupied with his domestic affairs and refused to intervene. However, during the same period a Chigtan am chi (practitioner of Tibetan medicine) managed to cure the queen of Ali Sher Khan in Skardu. When she offered him a reward, he asked for her husband’s help in restoring Tsering Malik’s descendants to their Chigtan inheritance. Ali Sher Khan duly sent a delegation to Sultan Khan who agreed to restore Adam Malik to Chigtan. Adam died childless and was succeeded by Chozang whose descendants continued to rule Chigtan until the early 19th century.6 A final reference comes from the last days of the Ladakhi kingdom. In 1834, the Dogra general Zorawar Singh conquered Ladakh on behalf of his master, 6 For further details of the lineage see Francke (1926:173) and Hashmatullah Khan (1939).

260

Khan, Bray, Devers and Vernier

Raja Gulab Singh of Jammu. In 1840 Zorawar invaded the region a second time to suppress a Ladakhi rebellion. Some 60 years later, Tsetan (Tshe brtan), a Khalatse villager who had served in the war, dictated his reminiscences (Francke 1926:253). He mentions Rahim Khan, the chief of Chigtan, as one of the rebels who was subsequently captured and taken to Skardu. Tsetan relates that Zorawar had Rahim’s right hand, tongue and ears cut off in a public mutilation in front of his army: he took two days to die of his wounds. Hashmatullah Khan identifies Rahim as the younger brother of Mohammed Ali Khan, the last ruler of Chigtan, who had likewise been killed while fighting the Dogras. Zorawar’s invasions led to the final defeat of the Ladakhi kingdom and its incorporation first into the dominions of Jammu and then, after 1846, into the combined princely state of Jammu & Kashmir. The descendants of the Chigtan chiefs retained high social status and in recent times one family member, Kacho Sikandar Khan, achieved renown as a historian and writer.7 However, they lost all formal political power. Their castle now served no practical purpose and was allowed to fall into disrepair. A particularly bleak period came in the 1960s when stones from the building were taken to construct a nearby hospital. The surviving ruins therefore constitute no more than a shell of the original building.

A Preliminary Analysis of the Castle Architecture

The best evidence for the castle’s original appearance now comes from photographs taken in the early 20th century. As recorded in his subsequent report, Francke (1914:99–101) visited Chigtan on behalf of the Archaeological Survey of India in 1909, and he was accompanied by a skilled photographer, Babu Pindi Lal. A comparison between Pindi Lal’s image with a second photograph taken from the same place a century later (Figures 10.1 and 10.2) shows how far the building has deteriorated. In 1909 the roofs and verandas were still in place. Now, they have fallen away completely. The site of the castle is particularly impressive, at the top of a cliff overlooking the Chigtan river, which is a tributary of the Indus some 15 miles to the north. Below the main building, a series of fortifications (in the left of the picture in Figures 10.1 and 10.2) follows the ridge down to the valley. The bottom of this ridge and the buildings that stood on top of it were blasted away in the 1970s to permit the construction of a motor road. Behind the main castle 7 See Kacho Sikandar Khan (1987). Kacho Mumtaz Ali Khan, one of the present authors, is Kacho Sikandar’s son.

Chigtan Castle and Mosque

Figure 10.3

261

Sketch plan of Chigtan fortifications by Martin Vernier and Quentin Devers. The main castle is to the lower left.

and out of sight in these pictures (but see the plan above) one can still see the remains of a wall connecting a series of smaller fortifications. As noted above, local legend has it that Tsangkhan Khan, the founder of Chigtan, was the first to build a castle on the site, and his successors no doubt expanded and strengthened the original fortifications. As they did so, they would have incorporated earlier structures into later buildings rather than replacing them entirely. The legend implying that Chandan built the entire castle in the course of a single building programme is therefore misleading: the castle must be understood as a complex of interconnected buildings, each with its own architectural style. What follows is a preliminary set of observations of the different components, pending more detailed archaeological examination.

The Ruins Behind the Main Castle

At the top of the mountain above the main castle there are the decayed ruins of a large round tower, and this appears to be the oldest remaining part of the fortification complex (Figure 10.4). From this high position there is no more than a limited view into the valley, which curves away on both sides, but there is a clear view of the arid mountains behind through which passes the

262

Figure 10.4

Khan, Bray, Devers and Vernier

Ruins of the upper tower. Photo: Quentin Devers.

trail leading to Kuksho and Dargo, the other two villages traditionally linked with the foundation of Chigtan. From the tower a decayed wall runs downhill towards a second cluster of buildings. This layout is reminiscent of the fortress at Basgo (Ba sgo) where there is a similar configuration consisting of a round tower at the top of a ridge linked to the rest of the complex by a wall, and this too appears to be the earliest part of the complex (Howard 1989:227–237). The second tower (proceeding downhill) is in rather better condition. Its original core, one of whose walls is circular, is made of stone masonry. The structure bears the marks of several successive repairs in stones, bricks and shuttered mud: more detailed on-ground observation will be required to confirm the precise order in which these repairs and additions took place. The original tower has been partly enclosed by walls made of shuttered mud (Figures 10.5 and 10.6). These features are somewhat reminiscent of the castle at Temisgang (gTing mo sgang) where two outer towers are enclosed on three sides by tall shuttered mud walls (Howard 1989:251–254). Below this tower there are several buildings variously constructed with both masonry and shuttered-mud, in which several different phases of repair both with stones and mud bricks can be observed. A gorge and a steep cliff cut off the lowest of these buildings from the main castle. However, some of the older inhabitants of Chigtan remember seeing a footbridge linking the castle to the upper buildings.

Chigtan Castle and Mosque

263

Figure 10.5

Partly enclosed tower. the original stone tower is just visible inside the shuttered mud walls. Photo: Quentin Devers.

Figure 10.6

The stone tower behind the walls of shuttered mud. Photo: Gerald Kozicz.

264

Khan, Bray, Devers and Vernier

Figure 10.7



Rear of the main castle. Photo: Quentin Devers.

The Main Castle

As discussed, the main castle as we see it today is the result of successive phases of constructions, reconstructions and repairs. The main access used to be via a set of stairs that have now collapsed: these were built against the site’s southern flank and led upwards from a gatehouse. Inside the gatehouse there was a tunnel leading to an underground spring, making it possible for the castle to withstand a siege. Such underground access to water is quite common in Ladakh, and is found in places as varied as Bod Kharbu (Bod mkhar bu—see Francke 1914:99), Nyoma (Nyo ma), Stok Monkhar (sTog [s?]mon mkhar) and Padum (dPal ’dum, sPa dum). Babu Pindi Lal’s photographs show that one of the inner buildings—the second from the front—included a striking set of timber lacing: a framework of wooden beams laid out in parallel lines with stones filling the spaces in between. Timber lacing is found all over Ladakh, but usually with fewer beams set between floors. The lacing shown in the photograph, with perhaps five or six beams per floor, more closely resembles the style common to the older mosques and castles of Baltistan (Hughes 2007). Geographically, one of the closest examples is the castle in Kharmang (mKhar mang) on the banks of the river Indus only 60 miles from Chigtan, but now on the far side of the Line of Control between Indian and Pakistani-controlled territory.8 8 For an early 20th century photograph of Kharmang, see de Filippi (1932:32).

Chigtan Castle and Mosque

Figure 10.8

265

Timber lacing as seen in 1909. Photo: Babu Pindi Lal. Courtesy Kern Institute, Leiden University, MS No.XI. fol.31.

Timber lacing has both practical and aesthetic appeal: the wooden framework makes the structure more resistant to both earthquakes and decay over time. At the same time, as Babu Pindi Lal’s photographs also show, the beams provide scope for intricate decorative wood carving. Again, the style of carving is reminiscent of similar decorative patterns found in the palaces and mosques of Baltistan (cf. Klimburg 2007:158–164). By contrast the front building with the balconies is more characteristic of Central Ladakh and Tibet. Most of the building is constructed with stone masonry, except for the upper storey which is made with mud bricks. Its slightly tapering walls and more spread-out timber lacings are firmly rooted in Tibetan architectural tradition. It is conceivable that early examples of such constructions might have been built already in the eighth and ninth centuries by the armies of the Tibetan Empire in locations such as Alchi (A lci) and Balu khar (Balu mkhar, near Khalatse) in the Indus valley. One of the earliest such buildings in the wider region that can be dated provisionally with a degree of confidence is the ruined castle of Malakartse (Ma lag(?) mkhar rtse) in Zangskar whose configuration seems to indicate it was constructed in the same period as the 11th century chorten (mchod rten) on the same site (Linrothe 2010). Similar types of

266

Khan, Bray, Devers and Vernier

construction can also be seen in Wanla (Wan la) and in Hankar (Hang? mkhar) in the Markha (dMar kha) valley, where the wood carving makes it possible to date the building to the same period as the Alchi group of monuments.9 As discussed by Philip Denwood elsewhere in this volume, there is a strong case for dating the Alchi Sumtsek (gSum brtsegs) to the late 11th century. However, the balconies built on the top floor of the Chigtan front building point to a much later date. They are reminiscent of those in the mid-17th century Leh palace, or, to choose an example more comparable in size, at the Zimskhang (gZims khang) palace at Khyagar in the Nubra (Nub ra) Valley. A careful examination of Babu Pindi Lal’s photographs also shows that the building with the closely-laid ‘Balti’ timber lacings seems to have been extended vertically with a different masonry. From the photograph it looks similar to the style of masonry found in most of the walls still visible today at the sides and rear of the main site (Figure 10.7). This points to an extensive reconstruction of the complex in a relatively short time period, and this would be compatible with the view that the castle was substantially reinforced—though not built for the first time—during the time of Tsering Malik.

A Tentative Chronology

As there is no direct stratigraphic relationship among the buildings in the site as a whole, an exact chronological sequence cannot be drawn out. However, it is possible to discern at least five phases of construction.

• • •

The first and oldest corresponds to the very decayed round tower at the top of the hill behind the main site and the wall descending from it. The second phase is probably that which saw the construction of the stone core of the tower immediately below the previous one. A third stage of construction, not necessarily immediately following in sequence, occurred when that tower was supplemented with shuttered mud walls. A fourth stage took place when the building with closely-laid timber lacings was erected in the main castle complex. The final stage seems to have included the construction of the front building with the tapering walls and more spread-out timber lacings, probably together with the walls at the sides and rear of the main complex.

• •

9 For Wanla see Howard (1989:261–264) and for Hankar, see Devers & Vernier, (2011:82–83).

Chigtan Castle and Mosque

267

To add to the complexity of this chronology, one has to keep in mind the many repairs of the walls in shuttered mud, using both stone masonry and bricks. A definitive chronology would require a more thorough examination of the site and dating of organic or ceramic materials enclosed in the walls. It is to be hoped that this task will be undertaken in the future.

Chigtan Castle Today

Chigtan castle retains its central place in the local cultural imagination. One example is a competitive game of riddles, where the winning team members imagine themselves taking possession of a progressively grander set of castles, culminating in Chigtan as the grandest castle of all. The winners tease the losers by comparing themselves with the king sitting at the top of the castle with a golden hookah, while the losers resemble dirty men with filthy clothes and hair full of lice. However, while the castle still holds its status as a symbol of power and status, the contemporary reality is very different. Writing about his visit in 1913–1914, the Italian traveller Giotto Dainelli (1932:242) reported that inside the castle “all is desolation and ruin.” He added that one could only guess at the purpose of the various rooms inside the castle because “everything is tumbling down, walls, stairs and ceilings, and threatens to fall on your head.” Today the situation is far worse. The main entrance path has fallen away, so that even entering the main castle complex is a hazardous endeavour. In 2009, the US-based World Monuments Fund placed the castle on its worldwide list of endangered monuments. In 2011 the Leh chapter of the India National Trust for Art and Cultural Heritage (INTACH) led by Tara Sharma started work on a preliminary survey of the castle with a view to stabilizing it and conducting essential repairs to prevent further deterioration. The aim is to rehabilitate the site to make it a platform to promote historical conservation. It is to be hoped that this is the first stage towards the preservation of one of Ladakh’s most important historical monuments.

The Mosque

Below the castle there is a small mosque consisting of a simple rectangular single-storey building in traditional Ladakhi style, about 15 metres long by 10 metres wide. The mosque has a carved wooden doorway at one end, and a mihrab niche to indicate the direction of Mecca at the other. In between, there

268

Khan, Bray, Devers and Vernier

are two wooden pillars. There is no tower or minaret, and the building’s modest style means that it is easy to overlook it. The simple design of the Chigtan mosque recalls the early 17th century Sharif Masjid in Leh, which is in the Tsa Soma (Tshas so ma—‘new garden’) area of Leh, below and to the west of Leh Palace. The Sharif Masjid, which has recently been restored by the Tibet Heritage Fund (THF), is the earliest mosque to be built in Leh (Alexander & Catanese 2008). It is thought that it was constructed during the reign of King Senge Namgyal, the son of King Jamyang Namgyal and his Balti wife Gyal Khatun. The mosque is rather larger than its Chigtan counterpart, and it has a small dome, but it is built in the same style, using traditional Ladakhi building techniques.10 However, we as yet have no secure dating for the Chigtan mosque and it seems that, like the castle itself, it is a composite construction incorporating parts of earlier buildings. The carvings on the door frame and the two pillars are its most striking feature. The frame of the doorway is carved with a ‘lotus’ design which is more usually associated with Tibetan Buddhist architecture. The capitals on the two pillars have even more intricate carvings. The holes below the capitals indicate that they once topped a squarish shaft, and this shows that they have been re-used from an earlier building, possibly the castle itself. In both cases, one side of the capital is more weathered than the other, and this suggests that they might once have been located within a courtyard. The first capital facing the door (see Figures 10.9, 10.10) shows two birds facing each other on one side, and a floral design on the other. On the bottom left there is a set of star-shaped patterns, and an interlaced pattern with a lily on the right. On the front of the second capital (Figures 10.11, 10.12) there are two hybrid dragon-like creatures of roughly equal size facing each other. On the other side there is a large hybrid dragon with a smaller one to its left. On the bottom there is an interlaced pattern topped by a trident on the left, and a similar interlaced pattern with lily flowers on the right. The two birds in the first capital might well represent peacocks, their long tails being turned into decorative scrolls to fill the sides of the surface. The floral design on the other side differs stylistically from the other three fronts but is close to the interlaced patterns in three of the four designs on the bottom of the capitals. Three of the four capital fronts are carved according to a central symmetry: the fourth one, with the larger ‘dragon’, is asymmetrical but compensates for this with the addition of the smaller ‘dragon’ on the left.

10

The original dome was later incorporated into a mosque in the nearby village of Shey, and has now been reconstructed as part of the restoration.

Chigtan Castle and Mosque

269

The designs on the bottoms of the capitals follow a similar mixed programme rather than a uniform one. Three sides have an interlaced flower style with a kind of ‘fleur de lys’ as a central motif. The fourth is more geometrically designed upon a juxtaposed star-like pattern. The execution of the floral designs in the first capital is also similar to what can still be observed through binoculars of the woodwork of Chigtan castle (see also the woodwork in Figure 10.8). It is likewise reminiscent of the floral interlaced design in the wooden lintel at Hankar in the Markha valley (Devers & Vernier 2011:Figure 39). Similar designs can also be found throughout Ladakh in paintings of various periods, for example in the early 17th century gSer zangs lha khang in Basgo. A similar pair of peacocks is engraved on a wooden capital of the gSer lha khang at Tabo in Spiti (Thakur 2001:133). The hybrid dragon motifs are more surprising as they show features that are reminiscent of the Central Asian steppe cultures’ animal style. These include the spiral lines on the thighs and shoulders as well as the way that the animal’s head is turned and looking over its back. Variations of these stylistic features have been found across Central Asia from ancient Scythia bordering on the Black Sea to western China.

Figure 10.9

Drawings of the first capital in the mosque by Martin Vernier. The rectangular panels show the bottom of the capital.

270

Khan, Bray, Devers and Vernier

Figure 10.10 Simplified line drawing by Martin Vernier showing the patterns in the first capital.

Figure 10.11

Drawings by Martin Vernier showing the second capital in the mosque.

As discussed in Tashi Ldawa Tshangspa’s paper in this volume, elements of this style can be seen in some of the oldest examples of Ladakhi rock art, notably the petroglyphs in the ‘Domkhar sanctuary’ (Figs 1.11, 1.12), which is a day or a day and a half’s walk from Chigtan, as well as the ‘Tangtse chase’ (Figs 1.13, 1.14) and the Renmudong petroglyphs in Tibet (Figures 1.15, 1.16). In Chigtan itself,

Chigtan Castle and Mosque

271

Figure 10.12 Simplified line drawing showing the patterns in the second capital.

Figure 10.13 Carving from the Kammand temple in Himachal Pradesh, dated approximately to the late 14th/early 15th centuries. Photo by Heinrich Poell.

272

Khan, Bray, Devers and Vernier

Figure 10.14 Carving from Kammand temple in Himachal Pradesh. Note the deer with its head turned backwards in the bottom centre. Photo by Heinrich Poell.

the closest example that comes to mind is the “large wooden board on which is carved a Naga-devouring Garuda” that Francke (1914:99) found in the castle a century ago. However, as Heinrich Poell points out, carvings incorporating a wide range of fantastic animals are also to be found in the wooden temple art of what is now Himachal Pradesh from the 12th/13th centuries onwards.11 As can be seen in Figs 10.13 and 10.14, these include dragons, composite creatures and fire-spewing deer, and they are portrayed with spiral lines, stars and other ornamentation on their bodies as well as the backward-bent head that is characteristic of the Central Asian style. In the absence of documentary evidence, we may never be able to identify the precise significance of these creatures in the mediaeval carvings of Himachal Pradesh and Ladakh. However, in both regions one may infer that they reflect local beliefs in a supernatural world inhabited by a wide range of deities and spirits, some of which may be portrayed in animal form. In contemporary Ladakh one thinks of the ‘lu’ (klu): these are water spirits that are often depicted in human or semi-human form but who may also take on the appearance of fish, lizards or snakes (Kaplanian 1981:211–213; 1995:101–108; Dollfus 2003). Another set of insights into traditional Ladakhi cosmology comes from 11

E-mail communication from Heinrich Poell, 13 November 2012. See Poell’s own paper in this volume for a discussion of similar themes in relation to the village temple in Lhachuse.

Chigtan Castle and Mosque

273

the wedding songs recorded by Francke (1923:33–35) which refer to snow lions, the king of the birds, a fish with a golden eye and a giant tiger. Similar mythical creatures appear elsewhere in the Tibetan cultural world. At the same time the comparison with mediaeval wood carvings from Himachal Pradesh suggests that Ladakhi artists were capable of incorporating—or perhaps in their turn contributing to—artistic styles and motifs from south of the Himalaya as well as Tibet and Kashmir. Acknowledgements The original core of this paper was a paper presented by Kacho Mumtaz Ali Khan at the 13th International Association of Ladakh Studies in Rome in 2007. Quentin Devers shared his archaeological knowledge and findings. Martin Vernier analysed and prepared the drawings of the wooden capitals in the Chigtan mosque. John Bray contributed much of the historical analysis and pulled the whole paper together. We gratefully acknowledge advice, information, suggestions and comments from Neil Howard, Tara Sharma, Abdul Ghani Sheikh and Heinrich Poell, while retaining collective responsibility for all errors of fact and interpretation.

Lamayuru (Ladakh)—Chenrezik Lhakhang: The Bar Do Thos Grol Illustrated As A Mural Painting Kristin Blancke

Figure 11.1

Statue of Bakula Rangdröl Nyima Rinpoche placed ad the centre of the Chenrezik Lhakhang.

According to a chronicle written by Bakula Rangdröl Nyima Rinpoche (Ba ku la Rang grol nyi ma) in 1862,1 Lamayuru monastery, built after Lotsawa Rinchen Zangpo passed through the region in the 11th century, originally consisted of 1 This unpublished chronicle, entitled g.Yung drung dgon dang po ji ltar chags rabs dang da ltar ji ltar gnas tshul gyi rnam dbye bi dza har tisma is mentioned in Vets & Van Quaille (1998:87). It is being translated by K.H. Everding. Bakula Rangdröl Nyima was the abbot of Lamayuru monastery. The dates of his birth and death are unknown, but he was a contemporary of Tsültrim Nyima (1796–1872) of Rizong monastery, and the two worked closely together. © koninklijke brill nv, leiden, ���4 | doi ��.��63/9789004271807_��3

Lamayuru (Ladakh)—Chenrezik Lhakang

275

five temples, one in each of the four directions and one at the centre, with statues and images belonging to the four classes of tantra, and countless paintings. After the Dogra invasions of Ladakh between 1834 and 1842, the original buildings were all but destroyed, and all the artefacts looted or, if impossible to carry away, smashed to pieces. In his chronicle Bakula Rangdröl Nyima describes his anguish and utter incredulity at the destruction of a formerly blessed and thriving monastery; he narrates how he had to go begging to accumulate enough money to rebuild the monastery and re-establish the monastic community. Here I would like to draw the attention to one rarely visited but very interesting temple in Lamayuru: the Chenrezik Lhakhang (sPyan ras gzigs lha khang), rising by the side of a thicket of willows in the northern section of the monastic compound, between the monastery kitchen and the small monks’ school. According to local sources, the location of this hall corresponds to the original site of the northernmost of the five temples.2 The reconstruction and decoration of this temple seems to have been a project particularly dear to Rangdröl Nyima. The temple dedicated to Chenrezik (Avalokiteshvara) is used mainly when people congregate there to recite one hundred million Chenrezik mantras,3 especially in the first 15 days of the Tibetan year and for one week during summer. The main feature in the hall is a large eleven-headed and thousand-armed Chenrezik statue. A local informant told me that Bakula Rangdröl Nyima commissioned the paintings in the hall so as to instruct his people about what it takes to become a Buddha (the south-east wall at the sides and above the main entrance is covered with detailed mural paintings of the Jataka tales and the life of the Buddha) and the way to get there through the devotional practice of Avalokiteshvara.4 On the north-west wall, on the left near the door to the inner sanctum, there is a beautiful although very damaged painting of 2 Lama Könchok Rigzin of the Central Institute of Buddhist Studies (CIBS) in Choglamsar has provided me with some of the stories related to the Chenrezik Lhakhang. 3 Maṇi dung sgrub, a collective practice done inside a temple, in which each participant recites the six syllable mantra oṃ maṇi padme hūṃ from morning till evening for consecutive days; every evening the number of mantras recited by all the people is counted and the practice goes on until one hundred million mantras have been accumulated. 4 From an interview with Lama Könchok Gyatso in Lamayuru (June 2009) I learned that Rangdröl Nyima, while devoting all his efforts to rebuild the entire monastery, was particularly interested in the Chenrezik Temple because he had a very strong personal connection with Avalokiteshvara. From Lama Könchok Rigzin I heard that, while staying at Tritapuri during his pilgrimage to Mount Kailash, Rangdröl Nyima had apparently received an object with the Chenrezik mantra from a naga, and this has been put inside the main statue in the temple.

276

Blancke

Avalokiteshvara, with rays of light emanating from his heart to each of the six realms of samsara in order to eliminate the suffering of all beings. The mural paintings on the right hand side (north-eastern direction), representing the visions one has during the intermediate state (bar do) between death and rebirth—which form the subject matter of this paper—follow the same rationale: they show how one can manage not to fall back into one of the six realms after death, and become enlightened instead. My reasons for writing this paper are twofold: first, these murals depicting the intermediate state between death and rebirth seem to be unique in the Tibetan world. Nowhere else have I seen a complete representation of the imagery related to the Bar do thos grol literature, including all the details of the visions experienced by a deceased person before he arrives at the point of choosing his next incarnation. Mural paintings of the peaceful (zhi) and wrathful (khro) deities, especially of the main zhi khro deities, are found in many temples, either as a cycle of independent images, or, as for example in the Lukhang Temple behind the Potala in Lhasa,5 in a particular context referring to the bar do practice of the inner yogas. Professor Tucci also speaks of one small temple at Chang in Upper Kinnaur, where the cycle of the naraka— infernal deities—is painted.6 But nowhere have I seen explicit illustrations of the Bar do thos grol chen mo (‘Great Liberation through Hearing’). That this teaching cycle was known in Ladakh is not surprising considering that by the 18th century the cycle of Kar gling zhi khro texts7 had spread all over the

5 See Winkler (2002): ‘Dividing the paintings on the northern wall on the second floor of the temple in 5 sections, section 3 is called “Peaceful and Wrathful Deities” ’. In this section is painted the whole cycle of peaceful and wrathful deities, while one yogin is seen meditating, which would suggest that he is meditating on the zhi khro deities as part of the practice on the bar do as contained in the rDzogs chen kun bzang dgongs ’dus “The Great Perfection, the Gathering of Samantabhadra’s Intention”, a tantra discovered by tertön O rgyan Padma gling pa that seems to be a source for the decoration of the northern section of the kLu khang.” Some of the images in the Lükhang murals may be seen in Baker & Laird (2001:73, 93–94). 6 See Tucci (1988:122–140). Tucci (1988:136) remarks that in Western Tibet he came across two thangkas depicting this cycle of deities, and he supposes that the fact that the region was under the influence of the Drugpa Kagyü school, in which this cycle of the narakas seemed to enjoy special diffusion, may explain its popularity in Western Tibet and Upper Kinnaur. I think that may hold true also for Ladakh. 7 A collection of texts started with a gter ma text hidden by Padmasambhava and revealed by the treasure discoverer (gter ston) Karma Lingpa in the 14th century, called Zab chos zhi khro dgongs pa rang grol (The Peaceful and Wrathful Deities: a profound teaching, natural liberation through [recognition of] enlightened intention).

Lamayuru (Ladakh)—Chenrezik Lhakang

277

Tibetan world,8 especially in the Nyingma and Kagyü lineages. These teachings were very popular in Bhutan, and we can imagine that from there they could have easily reached Ladakh through the Drukpa Kagyü lineages, or they could have spread from the Drigung Kagyü monasteries in Western Tibet. But the popularity of these teachings is not enough to explain why exactly they have been represented on a wall in Lamayuru. It took a special and unusual teacher such as Bakula Rangdröl Nyima to have them painted as a mural. If it is true that these paintings are unique, they are very precious and well worth illustrating, even if they are relatively recent and not so interesting from an art-historical point of view. Secondly, the wall paintings are in a bad state of conservation. Originally the paper was based on images received from Hilde Vets (dated 1995), Chiara Bellini and Laura Jokisaari, but for this publication a new set of images have been used made by Kaya Dorje Angdus. In the last few years there has been a minimal intervention by the Archaeological Survey of India to secure the building: some earth has been put on the roof to avoid leakage, the walls have been secured, the cracks and holes have been filled with mud. For the time being, the paintings are safe. I hope that an analysis of the significance of these murals, and the originality with which the subject matter has been approached, may eventually encourage a more durable restoration. Regarding the time of construction and decoration of the temple, from one inscription in the temple we learn that the Chenrezik Lhakhang was built and decorated in the ‘year of the horse’ a favourable year to go on pilgrimage to Mount Kailash. Most probably this must have been either 1846/1847 or 1858/1859.9 8 See Cuevas (2003:24). 9 On the right hand side near the entrance door to the Chenrezik Lhakhang (southeast direction) there is a small image of Mount Kailash, with three lamas admiring the sacred mountain. According to Lama Könchok Gyatso, this picture was made because the caretaker of the temple wanted to go on a pilgrimage to the Mount Kailash in the year of the horse (most probably either 1846/1847 or 1858/1859), but he did not receive permission to do so because he had to take care of the ongoing decoration works in the temple. As a reward for his efforts he received permission to have an image of the Kailash painted on the wall, as explained in the inscription near the mountain. This reads as follows: “gtsug lag khang ’di’i lha tshogs bzhengs skabs dbyar thog lnga’i ring la bza’ btung long spyod kyi bzhes pa’i do(ngo?) ’dzin khur len du seng ge sgangs pa’i tshe ring rgya mtsho bsdad de| lhag bsam rnam par dag pa’i sgo nas dge ba’i ’brel bar lus ngag yid gsum bkol ba ni| dal ’byor gyi mi lus la snying po len pas skal pa che| rta lor gangs ti se’i gnas sgo’i mjal kha sleb nas dad ldan pho mo mang po gnas skor du song yang tshe ring rgya mtsho’i rta lo’i chos skal gang ri chen po ti se’i sku’i bkod pa ’di ru bzhengs pa yin| de hi ma la ya zhes

278

Blancke

As to the painter(s) of the images, we unfortunately have no clues.10 The Bar do Paintings on the North-East Wall As mentioned above, the subject matter of the paintings on the northeast wall concerns the experiences that a deceased person goes through during the different stages between the moment of death and the start of a new life. From the Bar do thos grol chen mo,11 known in the West as ‘The Tibetan Book of the Dead’,12 one learns that from the moment of death the deceased will have to



10

11 12

yul nyi shu rtsa bshi’i ya gyal dang yang ‘khor lo sdom pa’i pho brang rgyal bas lung bstan pa’i yongs grags kyi gnas khyad par can gang gyi rna bar thos tshad ngan song med pa sogs kyi phan yon che ba dang|sku’i bkod pa ’di mig gi mthong tshad gangs ri chen po dngos su mjal ba khyad par med pa mdo nas gsungs so|” While the [images of the] deities in this temple were being created, for the duration of five summers,Tsering Gyatso from Sengé Sgangspé’s (Seng ge sgangs pa’i) [family] remained there, taking responsibility as caretaker for providing food and drinks. Engaging body, speech and mind in the practice of virtue in the purest possible way, this gave him the great fortune to use his precious human body with all its possibilities in its most meaningful way. When in the year of the horse the ideal opportunity arose for meeting the holy place of the Tise Snow Mountain, many faithful men and women went on pilgrimage; yet for Tsering Gyatso his portion of Dharma in the year of the horse was seeing to this composition of the image of the great Tise snow mountain right here. It is the foremost of the 24 holy places in the so-called Himalayas, and as prophesized by the Buddha it is the palace of Chakrasamvara. Whoever hears about this very special, universally famous holy place will obtain great benefits, such as not taking rebirth in the lower realms. Seeing this image with one’s own eyes is no different from actually meeting the great Snow Mountain itself; thus it is said in the sutras (translation by the author). Pallis (1995:241) suggests that the figure compositions are “probably mostly by Rigzin”, a very famous painter from Khalatse identified by Lo Bue (2005:354–58) as Tsewang Rindzin (Tshe dbang Rig ’dzin), who was active in Lamayuru between the late 1920s and early 1930s. However, Pallis’s and Lo Bue’s suggestion that this painter may have been responsible for the decoration in this temple should be discarded on chronological grounds, though Tsewang Rindzin may well have worked in Lamayuru after the decoration of the Avalokiteshvara shrine had been completed. The Tibetan title actually means ‘Great Liberation through Hearing in the Intermediate States’; this text is a guidebook for people who are dying. The various Western editions of the so-called ‘Tibetan book of the Dead’ are in fact translations of a few chapters of the Bar do thos grol chen mo, which belongs to a greater cycle of texts known as Zab chos zhi khro dgongspa rang grol, namely ‘The Peaceful and Wrathful Deities: a profound teaching, natural liberation through [recognition of]

Lamayuru (Ladakh)—Chenrezik Lhakang

279

pass through three intermediate states (bar do in Tibetan), during which he will have a series of experiences that will bring him, after 49 days, to a new incarnation in one of six possible existences: as a god, a demi-god, a human being, an animal, a hungry ghost, or a hell-being. If a spiritual guide reads the text to the deceased person day by day, pointing out the kinds of visions that he will experience, and if the deceased is able to follow the indications given to him, there is a possibility that during this intermediate period he will be liberated from the cycle of conditioned existence, or that he will be able to choose to reincarnate in a place auspicious for his further progress on the path towards liberation. Hence the second part of the title: thos grol, liberation through hearing. The text describes the three bar do states experienced after death: the intermediate state of the moment of death (’chi kha’i bar do), which starts at the time of death and ends with the manifestation of the ultimate nature of the mind; the intermediate state of ultimate reality (chos nyid bar do), in which the deceased experiences visions of deities, at first peaceful and, as time goes by, more and more wrathful and terrifying; and the intermediate state of becoming (srid pa bar do), with visions that will bring him to experience a new incarnation. On our wall in Lamayuru all three phases are illustrated, and every detail is explained by an appropriate inscription.

Structure of the bar do Mural Painting

The bar do mural, which is approximately 12 metres long and one and a half metres high, occupies the upper half of the wall; the lower section is painted in a plain brown colour. Structurally, the wall painting is divided into two rows: the upper row, starting with the image of Küntuzangpo, shows all the peaceful deities, first one by one, then all together, and finally the five Vidyadharas with their entourage. The lower row, starting from the left, first shows each of the main wrathful deities, one by one, then the row divides into two. Above, we can see the host

enlightened intention’. There have been several editions and partial translations of the Bar do thos grol chen mo, but its first complete translation was by Gyurme Dorje (2005). There the ‘Great Liberation through Hearing’ corresponds to the 11th chapter.

280

Blancke

Figure 11.2

Overview of the mural painting. Photo: Kaya Dorjay Angdus.

of wrathful deities with their entourage and a first image belonging to the srid pa bar do, and below various images of the srid pa bar do.

Detailed Analysis of the Mural Painting

1 The Intermediate State of the Moment of Death (’chi kha’i bar do)13 The first image in the upper left corner of the wall painting shows Küntuzangpo/Samantabhadra in sexual union with his consort, Küntuzangmo/ Samantabhadri. They symbolize the union of awareness and emptiness, the characteristics of the nature of mind. According to the Kar gling zhi khro, at the moment of death each of the physical elements of the body dissolves into one of the other elements (earth into water, water into fire, fire into air, air 13

In the Bar do thos grol chen mo text which I have followed this first couple of deities is included in the bar do of the moment of death, and the chos nyid bar do starts after the mind regains consciousness (see below). In some Dzogchen lineages the appearance of these two deities is considered already as part of the chos nyid bar do; they represent the peaceful manifestation of the indivisibility of space and awareness, whereas Chemchog Heruka and his consort represent the wrathful aspect of the same wisdom. See Dzogchen Ponlop (2008:175).

Lamayuru (Ladakh)—Chenrezik Lhakang

Figure 11.3

281

Küntuzangpo/Samantabhadra with Küntuzangmo/Samatabhadri. Photo: Kaya Dorjay Angdus.

into consciousness) and, once this process is completed, a state of pure light, the true nature of mind—the union of awareness and emptiness symbolized in the image of Samantabhadra and his consort—appears for a very short moment and, if recognized, brings one to liberation. This first image has no inscription. 2

Figure 11.4

The Intermediate State of Ultimate Reality (chos nyid bar do)

Main Peaceful Deities: the Tathagatas Vairochana, Akshobhya-Vajrasattva, Ratnasambhava, Amitabha, Amoghasiddhi. Photo: Kaya Dorjay Angdus.

But this recognition is quite difficult; so, after the moment of clear light experience has passed, if it has not been recognized the deceased falls unconscious for three and a half days. When he regains consciousness he will have

282

Blancke

to continue and roam in the intermediate state of ultimate reality passing through a series of visual and auditory experiences. In a first phase he will be confronted with peaceful deities: day after day he will meet one of the five peaceful Tathagatas in sexual union with his consort. From the heart of each Tathagata a ray of pure, brilliant, dazzling primordial wisdom-light will reach the heart of the deceased, together with a ray of dull light from one of the six realms. In fact below each of the peaceful and wrathful deities on our mural there is the image of a deceased person confronted with two rays of light, one emanating from the heart of the deity, and the other from one of the six realms.

Figures 11.5 and 11.6

14

The inscription below the Vairochana Buddha (appendix, inscription 1)14 reads, ‘On the first day after regaining consciousness, from the central pure land [called] Pervasive Seminal Point appears the Tathagata Vairochana, in yab yum [with his consort].’ The inscription for the wisdom light reads: ‘light of dharmadhatu wisdom’; the light from the six realms is the ‘white light of the god[s’ realm]’. Photo: Kaya Dorjay Angdus.

The transliteration of the Tibetan inscriptions will be found in the Appendix.

Lamayuru (Ladakh)—Chenrezik Lhakang

283

The inscriptions afford the name of each pure land and of the corresponding Buddha, together with the name of the particular wisdom aspect generating from his heart and the colour of the light of the particular realm of conditioned existence associated with it. In the Bar do thos grol text the spiritual guide will read out the characteristics of each deity and encourage the deceased to recognize the vision as an aspect of his own nature. Instead of fleeing from it, he should unite with the light, thus obtaining liberation. If he instead follows the softer and less fearful light of one of the six existences simultaneously manifesting with each of the Peaceful Buddhas, he will end up reincarnating in one of the six conditioned worlds of existence.

Figure 11.7

On the sixth day of the chos nyid bar do the five peaceful Buddhas appear together, each with his consort. On top: Küntuzangpo with Küntuzangmo and the Buddhas of the six realms; around: four female and four male gatekeepers. Photo: Kaya Dorjay Angdus.

However, becoming liberated at this point is not easy, and it is more likely that the deceased will let the opportunities pass without doing anything, thus arriving at the sixth day of the bar do of ‘ultimate reality’, in which the five peaceful deities with their consorts appear together, surrounded by male and female guardians, simultaneously with the lights of the six conditioned worlds. If the

284

Blancke

deceased recognizes these visions as aspects of his own mind and unites with them he will be liberated. Otherwise, on the seventh day of the chos nyid bar do the five Vidyadharas/ Awareness-holders—highly realized beings—will manifest, surrounded by yoginis, dakinis and spiritual heroes. If these are recognized by the deceased, he will be liberated.

Figure 11.8

The inscription (see Appendix, inscription 4) reads: ‘On the seventh day from the pure land of Khechara, the Vidyadhara deities come to meet [the deceased]. From the Vidyadharas appear the five consorts and around them, a numberless assembly of dakinis appears: those from the cemeteries, those of the four families, those of the three places, those of the twenty-four sacred places, along with male and female warriors, protectors and guardians.’ Each Vidyadhara is identified by an inscription: rNam par smin pa’i rig ’dzin, Sa la gnas pa’i rig ’dzin, Tshe la dbang ba’i rig ’dzin, Phyag rgya chen po’i rig ’dzin, Lhun gyis grub pa’i rig ’dzin. A five-coloured light of co-emergent pristine cognition (lhan cig skyes pa’i ye shes) emanates from the heart of each Vidyadhara and reaches the heart of the deceased, together with the green dull light of the animal realm (dud ’gro’i ’od ljang khu). Photo: Kaya Dorjay Angdus.

After the visions of these peaceful deities, there follows a series of more terrifying experiences of wrathful deities. On the mural we now have to go to the second row of paintings, starting from the viewer’s left.

Lamayuru (Ladakh)—Chenrezik Lhakang

Figure 11.9

285

Main wrathful deities: Mahottara Heruka, Buddha Heruka, Vajra Heruka, Ratna Heruka, Padma Heruka and Karma Heruka, each with his consort. Photo: Kaya Dorjay Angdus.

One after the other, we see the wrathful blood-drinking deities, each united with his mystical consort, and identified by an inscription, with the image of the deceased below each. Surprisingly, by contrast with the text in the Bar do thos grol, on this mural there are six deities: the first one representing Mahottara/Chemchok Heruka (Che mchog He –ru ka), the wrathful aspect of Küntuzangpo, is not mentioned in the text.15 In the Bar do thos grol text the visions experienced by the deceased are described as dreadful: very bright lights, very intense rays of light, and violent sounds. But the person guiding him will explain all details of these visions and will encourage the deceased not to be afraid and to become one with the deities. If he recognizes all those appearances as manifestations of his own mind, he will be liberated.

Figure 11.10

15

The assembly of the fifty-eight wrathful deities with their entourage. Photo: Kaya Dorjay Angdus.

Che mChog Heruka/Mahashri Heruka/Mahottara Heruka is mentioned among the main wrathful deities in one text included in the Zhi khro dgongs pa rang grol cycle and called ‘Zhi khro dgongs pa rang grol gyi cho ga sdig sgrib rnam par sbyong-ba’. This deity is also mentioned by Tucci (1988:136) as the central deity in a thangka representing a mandala of wrathful deities; Tucci considers these deities to be the centre of the mandala of the Narakas (infernal deities). The fact that Che mChog Heruka is not included in the

286

Blancke

After five more days the entire mandala of fifty-eight wrathful deities will manifest. Besides the Herukas, the female Gauris/Keurimas and Pishachis/ Tramenmas (Phra men ma, with animal heads) now also appear, as well as the four female gatekeepers, and the twenty eight Ishvari goddesses, yoginis of the different directions with animal heads, and female and male guardians of the directions, each deity being identified by an inscription with her/his name (Appendix, inscription 5).

Figure 11.11

Detail of 2 Keurima deities: Tseuri (Tse’u ri) and Tramo Marmo (Pra mo dmar mo).

Figure 11.12

Two Ishvari deities, yoginis of the southern direction: Gawa Marser (dGa’ ba dmar gser) and Zhiwa Marmo (Zhi ba dmar mo). Photo: Chiara Bellini.



Thos grol chen mo text might mean that either it ‘disappeared’ from the text in the course of time or that the Cho ga sdig sgrib did not belong to the nucleus of the text, but was added later on. Anyway, the fact that Che mchog Heruka does appear in the Lamayuru wall painting might mean either that the painter used another text in which it was still included or that the painting of the wrathful deities actually refers to a cycle popular in Ladakh, as demonstrated by the paintings in Chang described by Tucci (1988: 123–140).

Lamayuru (Ladakh)—Chenrezik Lhakang

287

3 The Intermediate State of Becoming (srid pa bar do) However, if one is swept away by one’s own fearful visions and does not recognize them as aspects of one’s own mind, one will have to proceed to the third phase, the intermediate state of becoming. One will have experiences such as being persecuted by the demons of one’s own karma, being pushed into a snowstorm or into pitch dark, and one’s sole desire will be to flee.

Figure 11.13

First image of srid pa bar do. The inscriptions read: ‘Many particular appearances arise: fierce winds, karmic flesh-eaters and cannibals brandishing many weapons, being followed by wild beast from the back, a very thick darkness in the front, hail storms, mountains collapsing, floods, blizzards, fire spreading, great winds, falling from a white, a black and a red mountain (Appendix, inscription 6). Photo: Chiara Bellini.

Then one’s own visions will assume the appearance of the Dharmaraja Yama, the Lord of Death,16 flanked by a white god and a black demon, each exposing a heap of small pebbles, white or black according to the good or bad deeds committed by the deceased.

16

For a description of the “judgment of Yama” scene, in the Zab chos zhi khro dgongs pa rang grol one chapter describes an allegorical play in which the archetypal evil-doer Laksanaraka and the archetypal virtuous householder Srijata meet with Yama in the srid pa bar do. The title of the chapter in the Tibetan text is Srid pa bar do’i dge sdig rang gzugs bstan pa’i gdams pa srid pa bar do rang grol. In Gyurme Dorje (2005:317–341) this is the 13th Chapter called “A Masked Drama of Rebirth”. Lo Bue (2005) has commented on this text on the basis of two Tibetan editions of the text compared with the translation of the chapter by Tucci (1972). The scene depicted on the mural does not completely correspond to the description in this chapter of the book: the characters are slightly different.

288

Figure 11.14

Blancke

Inscriptions on the slate of the monkey-headed assistant: ‘letter of white and black deeds’. Below: Yama,: ‘the Dharmaraja, Lord of Death, sees the white and black deeds in his mirror’. Below: ‘the inborn god and demon assemble white and black pebbles’ (Appendix, inscription 7). Photo: Kaya Dorjay Angdus.

Looking in his mirror of karma, the Lord of Death will decide where the deceased is to be reborn, and the latter will be dragged there without being able to resist. After all those terrifying appearances, the deceased will now want to find a place to take rebirth.

Figure 11.15

Signs for places where the deceased is to take rebirth: lights of the six realms, and feeling of going up, sliding down or walking flat. Photo: Kaya Dorjay Angdus.

Lamayuru (Ladakh)—Chenrezik Lhakang

289

Lights from the six samsaric realms will manifest, the brightest one indicating in which realm the karma of the deceased will lead him to reincarnate, and he will have the impression that he has to ascend a mountain (in which case he will take birth in the higher realms) or to descend (in case of rebirth in the lower realms) or else to walk horizontally (in case of rebirth in the human realm). But he is constantly haunted by fearful visions from which he wants to escape. So when he sees a couple making love he will be tempted to enter into the womb of the woman, thus taking rebirth. At that point several instructions from the Bar do thos grol are read to the deceased to allow him to avoid entering a womb and either to enable him to be liberated or, if that is not possible, to choose a kind of life that will allow him to progress on the path towards liberation.

Figure 11.16

Appearances arise such as whirlwinds, blizzards, hailstorms, darkness (Appendix, inscription 8). In the right corner below a couple is making love and that is where the deceased is tempted to flee (I have not been able to decipher the inscription below the couple). Photo: Kaya Dorjay Angdus.

As the deceased will have now to choose a place for rebirth, it is very important that he receives advice on which to choose. So he will learn to recognize the signs of the four continents: the eastern continent has a lake with swans; the southern continent, Jambudvipa, where he should take rebirth, has nice mansions; the western continent has a lake with horses; and the northern continent has a lake with cattle and nice forests.

290

Blancke

Figure 11.17

Image of the four continents. This portion of the paintings is badly damaged. Photo: Kaya Dorjay Angdus.

Then there will be appearances of the five realms he should avoid: nice houses and temples for the gods’ realm; a pleasant forest and wheels of light for the demi-gods’ realm; rocky caverns, crevices and straw sheds wrapped in dense fog for the animals’ realm; tree stumps, black protruding silhouettes, blind desolate gorges or total darkness for the hungry ghosts’ realm; and a land of darkness with red and black cubes, black earth pits and black roads for the hell dwellers’ realms. The text in the inscriptions advises him not to enter there.

Figure 11.18

Image of five unfavourable realms where the deceased should not take rebirth: gods’ realm, demi-gods’ realm, animals’ realm, hungry ghosts’ realm (Appendix, inscription 9) and hell realms (Appendix, inscription 10).

For example, the inscription for the hungry ghosts’ realm reads: As for the signs of taking birth as a hungry ghost, [one finds oneself] in empty plains and blind desolate gorges, or sees jungles and black worlds. One should not enter there.

Lamayuru (Ladakh)—Chenrezik Lhakang

291

And for the hell realms one reads (the inscription is not complete): If one is to take birth in a hell realm, hearing the song of those with bad karma one wants to enter [there]; appearances arise of arriving in a dark country, of red and black houses, and black holes in the earth and a black road. Pursued from the back and drawn from the front by executioners, in darkness and tumult, snow and rain and strong hailstorms and turbulent blizzards, one flees there; fleeing there one seeks refuge and liberation inside a mansion, a shelter, a cave, a dark forest, a lotus flower and so forth, thinking: ‘Now I should not get out of here’ and through attachment [to the chosen place] one will assume an inferior body and feel suffering . . . Having thus avoided an unfavourable rebirth, in the end the moment will come in which one can no longer delay rebirth. The inscription on the mural gives a few instructions: avoid attachment and aversion, and pray to be able to take a new birth that will be beneficial to others. If that is not possible, then

Figure 11.19

The inscription for the last image is as follows (Appendix, inscription 11): ‘If one is to take rebirth in an impure place, one will perceive a good smell from impure [objects] and, being attached to that, one will take birth. Therefore one should abandon attraction and aversion towards whichever appearance arises, and one should enter [a womb/a place] thinking: ‘May I assume the body of a fortunate person capable of benefiting all sentient beings.’ Ordinary beings who are not able to remain free from attachment and aversion in this way should call the Three Rare and Precious ones by name and take refuge in them and pray to the Great Compassionate One. Then they should enter the blue light of the human [realm], the white light of the gods’ [realm], nice houses and the like.’ Photo: Kaya Dorjay Angdus.

292

Blancke

pray to the Three Jewels or Chenrezik and, taking refuge in them, be born as a human or god.

Comparing the Bar do thos grol Text and the bar do Mural Painting in Lamayuru From the above detailed analysis of the different features in the mural painting, it can clearly be seen that all the aspects described in the ‘Tibetan Book of the Dead’ have been scrupulously represented here, with only minor differences such as the addition of Chechok Heruka and slight variations in the judgment scene. However, there are a few differences as far as the presentation is concerned. First of all, the Bar do thos grol, a manual to be read by a person guiding a deceased, is styled in the vocative form, addressing the deceased in the second person.17 This obviously does not hold for the mural. The function of the mural is to make people become familiar with the appearances that will manifest at the time of death. Therefore, rather than being presented in the second person, the inscriptions are put in an impersonal form and remain merely descriptive. Lama Könchok Gyatso suggested to me that the mural should be considered as a bar do mthong grol (‘liberation in the intermediate state through seeing’) rather than a bar do thos grol (‘liberation in the intermediate state through hearing’). In this sense the mural could be considered more similar to the gos chen thang ka tradition, in which a large brocade image of a deity, believed to bring liberation through seeing, is displayed to devotees on a particularly auspicious day. Secondly, after studying the inscriptions in the mural it also becomes clear that these are much simpler than the instructions given in the Bar do thos grol. Of course it would be impossible to write the full descriptions from the text on a wall, but that is not the only reason: the few instructions given, such as those reported in Fig. 18, are themselves much simpler than the ones proposed in the thos grol text, where plenty of references are made to quite complicated meditation instructions belonging to the Dzogchen and the Mahamudra lineages, which require a thorough training in those meditation systems before they can be applied during the bar do. It seems that the mural painting is addressed to a different kind of public: the paintings with their more sober descriptions can be readily understood by the general public viewing them, even without the preparation of years of meditation. To me that seems to indicate that the mural falls totally in line with the general purpose behind the paintings in 17

In the Appendix inscription 10 is compared with Gyurme Dorje (2005:295–296) and Karma gling pa (1992:245–247).

Lamayuru (Ladakh)—Chenrezik Lhakang

293

the Chenrezik temple as outlined in the introduction of this paper: Bakula Rangdröl Nyima wanted to instruct ordinary devotees so that at the moment of death they could take advantage of what they had seen by visiting this unique temple.

Appendix: On Some Inscriptions in the Mural Painting of the bar do So far I have not had the opportunity to note down all the inscriptions on the wall. Below are some of those I was able to make out from Chiara Bellini’s pictures. 1) Vairochana (Figs 11.5–11.6), bottom-left of the image: brgyal sangs pa dang zhag dang po la dbus phyogs thig le gdal ba’i zhing khams nas bcom ldan bdas rnam par snang mdzad yab yum ’char. The inscription for the wisdom light reads:chos kyi dbyings kyi ye shes ’od; the light from the six realms is the lha’i ’od dkar po 2) Amitabha, bottom-left of the image, above the deceased and the rays of light: yi dvags kyi ’od Below the deceased: sor rtogs ye shes kyi ’od zhag bzhi pa la nub phyogs bde ba can gyi zhing khams nas bcom lden bdas [written in contracted form] snang ba mtha’ yas ’khor bcas ’char: On the fourth day from the western pure land of Dewachen the Tathagata Amitabha appears with his retinue. 3) Amoghasiddhi, above the rays: lha min gyi ’od Below the rays: bya sgrub ye shes kyi ’od. 4) Five Vidyadharas (Figure 11.8). Each Vidyadhara is identified by an inscription, mentioning also the light from the animal realm:

294

Blancke

rnam par smin pa’i rig ’dzin| tshe la dbang ba’i rig ’dzin| phyag rgya chen po’i rig ’dzin|sa la nas pa’i rig ’dzin| lhun gyis grup pa’i rig ’dzin| dud ’gro’i ’od. On the right of the painting we find the following inscription: zhag bdun pa la dag pa mkha’ spyod kyi zhing khams nas rig ’dzin lha tshogs rnams kyis bsu ba la ong ste| rig ’dzin las yum lnga dang de’i phyi rim na mkha’ ’gro ma’i tshogs dpag tu med pa | dur khrod kyi mkhro| rigs bzhi mkhro| gnas sum gyi mkhro| yul nyi shu rtsa bzhi mkhro rnams dang|dpa’ bo dpa’ mo chos skyong srung ma dang bcas pa ’char| 5) Host of wrathful deities (Figure 11.10) From the available pictures I was able to identify some of the deities, almost halfway through the two lines of the inscription; the order is the same as in the text, with only minor differences in spelling. First line: ke’u ri| tse’u ri| pra mo dmar mo| pe ta li| pu ka si| kasma ri| tsan da li| sma sha li| sing ha| bya kri mu kha| sri la mu kha| msho na mu kha| tri ta mu kha| . . . tshangs ma sbrul mgo ma| lha chen| rtog ‘dod ma Second line: gzhon nu dmar mo| brgya byin dkar mo| dga’ ba dmar ser| zhi ba dmar mo| bdud rtsi dmar mo| zla ba dkar mo| be con ljang nag| srin mo ser nag| zla ba ljang nag bya rgod mgo| dbang ldan dmar nag| stobs can dkar mo|{srin} mo ser mo|’dod pa dmar mo| . . . . . rdo rje ser mo| rdo rje dmar mo| rdo rje ljang nag| 6) Karmic appearances in srid pa bar do (Figure 11.13) rlung dmar| las kyi sha za srin po mtshon cha thogs pa mang po|gcan zan gyis rgyab nas ‘ded pa| mdun nas shin tu ’thibs pa’i mun nag| ser ba babs pa| ri nyil ba| chu lud pa| bu yug| me mched pa| rlung chen po| ri dkar po| ri nag po| ri dmar po gsum las lhung la ’khad pa’i snang ba ’byung ngo|

Lamayuru (Ladakh)—Chenrezik Lhakang

295

7) Judgment scene (Figure 11.14), on the slate held by monkey-headed: las dkar nag gi yi ge| below: lhan cig skyes pa’i lha dang ’dres rde’u dkar nag ’du ba| below Dharmaraja: shin rje chos kyi rgyal pos las dkar nag me long la gzigs pa| 8) Scene illustrating the need to block entrance into womb (Figure 11.16) rlung ’tshub dang bu yug dang ser ba dang mun nag sogs kyi snang ba byung (I was unable to decipher the inscription below the couple making love). 9) Signs of rebirth in the hungry ghosts’ realm (Figure 11.18): yi dvags su skye ltas ni thang stong dang grog po phug rdugs dang gnang nags sing [sic for nag seng?] dang nag gling du mthong| der mi ’jug pa dgos| [in the Bar do thos grol text: yi dvags su skye na ni: sdong dum dang: nag breng nge ba dang: grog po phugs sdugs sam: nag ling bar mthong ngo| If you are to take birth as an anguished spirit, you will see tree stumps, black protruding silhouettes, blind desolate gorges, or total darkness (translation by Gyurme Dorje 2005:295). 10) Signs of rebirth in the hell realms (Figure 11.18) dmyal bar skye na ni las ngan pa’i glu dbyangs su thos pa’am ‘jug dgos pa’am|mun nag gling dang|khang pa dmar po dang nag po dang sa dong nag po dang| lam nag po la sogs par phyin pa’i snang ba byung| der kyang gshed mas rgyab nas ’ded cing mdun nas ‘khrid pa dang| mun nag| ku sgra| kha char| ser drag| bu yug ’tshub cing| der bros pa’i snang ba byung| der bros te skyabs ’tshol du phyin pas| khang bzang ngam| brag skyibs| sa phug| nags seb| me tok padma la sogs pa’i nang du thar pas| da ‘di nas phyir thon na mi rung ngo snyams ste chags pas gang men{sic for man} pa’i lus blang nas sdug bsngal myong ba . . . Comparison with the Bar do thos grol text (Dharamsala edition, pp. 245–247); the lines that are not reported on the mural painting are in italics.

296

Blancke

dmyal bar skye na ni las ngan pa’i glu dbyangs su thos pa’am| dbang med du ‘jug dgos pa byung na| mun nak gi gling dang| khang pa nak po dang| khang pa dmar po dang| sa dong nak po dang| lam nak po la phyin pa la sogs pa’i snang ba ong| der phyin na ni dmyal bar tshud dé| tsha grang gi sdug bsngal bzod glags med pa smyong ste| thon pa’i dus mi ong bas| de’i gseb tu mi ’gro bar gzab par bya’o |ci nas kyang ma tshüd par bya’o|mngal sgo bkag nas ru log dran par bya’o| zhes gsungs pa de dus da res dgos pa yin no| kye rigs kyi bu| khyod mi ’gro bar ‘dod kyang rang la dbang med par ’gro dgos pa dang| mdun nas gshed ma dang srog gcod kyi khrid pa dang| mun nag dang| rlung dmar ’tshub ma chen po dang| ku dgra dang| kha char dang| ser ba dang| bu yug ’tshub cing bros pa’i snang ba khyod la ’ong ngo| der bred cing skyabs tshol ba la phyin pas|gong du bshad pa’i khang bzang sam|brag skyibs dang| sa phug dang| nags gseb dang| me tog padma la sogs kha zum pa’i nang du thar nas| de na gab ste phyir thon gyi dogs nas| nga ’di nas phyir thon na ni mi rung ngo snam nas| de dang bral gyis dogs pas khyod gnas de la shin tu chags par ’gyur ro| de nas phyi rol tu thon na ni bar do’i ’jigs skrag de rnams dang phrad kyis dogs pas| de dag la ’jigs shing dngangs pas nang du gab pas| gang nas man pa’i lus ngan pa cig blang nas| sdug bsngal rna tshogs nyams su myong tu ’ong bas. . . . Translation by Gyurme Dorje (2005:295–296): If you are to take birth as a hell being, you will hear songs of those of negative past actions. Or, quite simply, you will feel powerless and compelled to enter. Whereupon, the perception will arise that you are moving into a land of darkness, where there are black and reddened houses, black earth-pits and black roads. Were you to be drawn to this place, you would enter the hells, and experience the [searing] unbearable sufferings of heat and cold. Be careful! Do not enter into the midst of this, for there will be no opportunity to turn back. Do not enter there, under any circumstances! As it is said [in the root verses]: ‘You must obstruct the womb entrances and call to mind the methods of reversal.’ These are [wholly] necessary now! O, Child of Buddha Nature, although you do not wish to move forward, you are powerless not to do so. The avenging forces, who are the executors of the unfailing laws of cause and effect, will be pursuing you. You will have no choice but to move forward. Before you, the avengers and executors will be leading the way. The experience will arise of trying to flee from these forces, of trying to flee from the darkness, from the most violent windstorms, from the [thunderous] tumult, the snow, the rain, the

Lamayuru (Ladakh)—Chenrezik Lhakang

297

hail and the turbulent blizzards, which swirl around you. [Frightened], you will set off to seek a refuge and you will find protection inside an enclosed space, such as within the mansions just described, or in rockshelters, or holes in the ground, or amongst trees, or within the bud of a lotus flower. Hiding there, you will be very hesitant to come out, and you will think: ‘I should not leave here now’. You will be very reluctant to be separated from this protected place and you will become utterly attached to it. Then, because you are so very reluctant to go outside, where you would be confronted by the fears and terrors of the intermediate state, you will, because of this fear and awe, continue to hide away. Thus, you will assume a body, however utterly bad that may be, and you will, [in time], come to experience all manner of sufferings . . . 11) Last image (Figure 11.19). Mi gtsang ba’i khrod du skye ba zhig yin na ni| mi gtsang ba’i dri zhim pa’i ’du shes skyes nas der chags te skye bar ’gyur bas| de rnams kyi snang ba gang byung yang chags sdang spangs te| sems can thams chad kyi don byed nus pa’i bsod nams can gyi lus shig blang bar bya’o snyams nas ’jug pa las| de ltar chags sdang dang bral ma thub pa’i dbang po tha ma rnams kyis dkon mchog gsum gyi mtshan nas brjod cing skyabs su ’gro ba dang thugs rje chen po la gsol ba gdebs cing mi’i ’od sngon po dang lha’i ’od dkar po khang bzang la sogs la ’jugs dgos pas dge’o| Acknowledgements My greatest thanks are to Prof. Erberto Lo Bue, who sent me the pictures of Chiara Bellini and Laura Jokisaari and encouraged me to work on this project. I thank Chiara and Laura for allowing me to use their pictures, and Hilde Vets for sending me her pictures and article. I thank Bakula Rangdröl Nyima Rinpoche who kindly allowed me to copy the chronicle of his predecessor and gave me some practical advice on how to proceed with my research. Further thanks to Lama Könchok Gyatso in Lamayuru who took the time to describe enthusiastically the personality and the endeavors of Bakula Rangdröl Nyima Rinpoche and who pointed out the image of Mount Kailash in the Chenrezik Lhakhang. Many thanks also to Lama Könchok Rigzin of the Central Institute of Buddhist Studies (CIBS) in Choglamsar who prepared a short text in which he gathered useful informations about the Chenrezik Lhakhang. I thank Kaya Dorjay Angdus who provided me with a new set of images of the mural painting. Final thanks to John Bray for editing my article.

The Lost Paintings of Kesar John Bray One of the themes of this book is the need to document and if possible conserve fragile buildings and art works before they are lost. The purpose of this article is to bring together the sources and surviving illustrations of a set of mural paintings that were already lost almost a century ago. The paintings existed at the house of the Kalon (bKa’ blon—minister) in Changspa (Changs pa), on the edge of Leh, and are said to depict the epic hero Kesar (Ke sar/Ge sar). They are unusual both because of their subject matter and because they are among the few recorded Ladakhi examples of detailed narrative paintings—as distinct from decorative patterns—to be found on the walls of a private house rather than a monastery.

Kesar in Ladakh and Tibet

For most Ladakhis, the story of Kesar is an intrinsic part—albeit somewhat vague—of the history of their own homeland. According to the La dvags rgyal rabs, Ladakh’s royal chronicle, the descendants of Kesar ruled Upper Ladakh in the 10th century CE (Francke 1926:93). Local landmarks include his footprints in Zangskar, and the stone on which he sat in Lower Ladakh (Tsering Mutup 1983:10). At the same time, the epic is well-known for its numerous variants across Tibet and as far as Mongolia. The common features of the epic are that Kesar is a divine being, the son of the lord of heaven, who is born on earth into the kingdom of gLing. As a hero with miraculous powers, he fights a series of battles, notably against the Hor, who in Ladakh are identified either with the Mongols or the peoples of Eastern Turkestan/Xinjiang. The variants from the eastern Tibetan region of Kham (Khams) include a further campaign against lJang, which is understood to have been located in what is now Yunnan. Through most of its history, and still today, Kesar has been primarily an oral epic, recited and sung by local bards, and passed on from generation to gener­ ation by word of mouth. However, in the last three to four centuries there have also been a number of written versions originating in Tibet and Mongolia.1 1 The first text to be published in a Western language was Die Thaten Bogda Gesser Chan’s, Isaak Jakob Schmidt’s 1839 German translation of a Mongolian text printed in Beijing in 1716. On the wider history of the Kesar saga see in particular Stein (1956) and Samuel (1992). On

© koninklijke brill nv, leiden, ���4 | doi ��.��63/9789004271807_014

the lost paintings of kesar

299

No such texts are known to have originated in Ladakh, but Tibetan versions seem to have circulated there. In 1834 the Hungarian scholar Alexander Csoma de Kőrös, who had conducted his initial research in the Zangskar region of Ladakh, cited the epic among the literary works to be found in Tibet.2 By the mid-19th century one of these Tibetan texts had found its way to the house of the Kalon in Changspa.

Kesar and the Kalons

As their name suggests, the Kalon family traces its descent from the leading minister of the historical kingdom of Ladakh. In 1834 the Dogra general Zorawar Singh invaded Ladakh on behalf of Raja Gulab Singh of Jammu. After a series of attacks and counter-attacks, the kingdom finally lost its independence in 1842, and in 1846 it was incorporated into Jammu & Kashmir, a ‘native state’ within Britain’s Indian empire. Despite these changes, successive heads of the family retained the title of Kalon even though they now exercised no formal political power. The first reference to the Kalon in connection with Kesar comes from 1865 in an exchange of letters between Professor Anton Schiefner (1817–1879) of St Petersburg and the Moravian missionary Heinrich August Jäschke (1817– 1883). In October 1865, Jäschke wrote that the Kalon possessed a written copy of the Kesar epic, but was unwilling to let any outsider read the text, let alone copy it (Schiefner 1869:1). Two years later, he was able to report that the situation had changed. The Government of India had deputed a British officer to Leh with a brief—according to Jäschke—of protecting Indian merchants from the extortion of local officials.3 This British officer exercised his influence on Jäschke’s behalf, and the Kalon was persuaded to release his treasure for transcribing by a team of local lamas in the u med Tibetan script. Jäschke in turn arranged for the transcript to be forwarded to St Petersburg. He had no time for the Ladakhi variants see in particular Francke (1905–1945), Tsering Mutup (1983), Jettmar (1990), Herrmann (1991) and A. Khan (2009). 2 Csoma de Kőrös (1834:180) described the epic as a “fabulous history (in Tibetan) of a celebrated warlike king (called Qésar) in the high, central or northern part of Asia; but the time in which he lived, the Tibetans cannot determine”. 3 This is evidently a reference to Dr Henry Cayley who was deputed to Leh in 1867 as an Officer on Special Duty with orders to monitor trade and gather intelligence. In 1870 the Government of India signed a treaty with Kashmir providing for the appointment of a British Joint Commissioner in Leh on a regular basis.

300

bray

more than a cursory look at the manuscript before despatching it, but the title of the text indicates that it refers to Kesar’s war against lJang.4 It may therefore have come from Kham. The next reference comes from Dr Karl Marx (1857–1891), a second Moravian missionary who—alongside his medical activities—conducted research into Ladakhi history. His main contribution was to begin a new translation of the La dvags rgyal rabs, the Ladakhi royal chronicle, and this was published in three parts in the Journal of the Asiatic Society of Bengal (1891, 1894, 1902). In a footnote to the first article, Marx (1891:116) writes: In a house at Leh, belonging to one of the old Ka-lon (State-minister) families, illustrations of the story of Ge-sar may still be seen painted on the wall all round one of the rooms. He adds that his own small collection of Tibetan manuscripts includes part of the ‘Iang’ (presumably lJang) and Hor wars. The former—which may have been the same as the Kalon’s text—was written in the dialect of Kham, and was therefore difficult for people based in Ladakh to read. The text on the Hor war was much easier. Marx died of typhus in 1891, and his immediate successor as a historian of Ladakh was a third Moravian missionary, August Hermann Francke (1870– 1930), who arrived in Leh in 1896. Francke had a personal link to Marx in that his wife Anna Theodora was the sister of Marx’s widow. He subsequently became well-known for his researches into the language, oral literature and history of Ladakh and, together with his wife, was responsible for supervising the publication of the second and third parts of Marx’s work on Ladakhi history.5 In a letter written in 1900 to Professor Albert Grünwedel (1856–1935) at the Museum für Völkerkunde in Berlin, Francke explained that he had followed up Marx’s footnote. In a sorrowful tone that anticipates the laments of 21st century conservationists, he lamented that the custodians of the paintings were neglecting their heritage:

4 The St Petersburg manuscript is listed in Stein (1959:84–85) who records the title as ’Jam gling rgyal mchog Ge sar dgra ’dul gyis lJang yul bdud sde ’dul ba’i g-yul ’gyed kyi le’u dpa’ bo bldan rnams kyi yid ’ong. I am grateful to George Fitzherbert for pointing me to this reference. 5 On the missionary background to Francke’s researches see Bray (1999/2008). Walravens & Taube (1992) present a detailed bibliography of Francke’s published works and surviving manuscripts.

the lost paintings of kesar

301

. . . I have been to the relevant house in Leh and found almost everything destroyed because the owner in incomprehensible carelessness had begun to whitewash the room. Apart from a few pictures referring to the Dogra war, I found a single but almost undamaged picture of the Kesar saga, as attached, and at once had it copied (Grünwedel 1901:281).6 Grünwedel in turn wrote to Richard Andree (1835–1912), the editor of the geographical journal Globus: Illustrierte Zeitschrift für Länder- und Völkerkunde, and offered to prepare a line drawing of Francke’s sketch for publication.7 Andree accepted, and the drawing was duly published in Globus (Grünwedel 1901:281) as below:

Figure 12.1

Copy of the Kesar wall painting in Globus 79 (1901:281).

6 My translation. An extract from what must be a subsequent letter, dated 1 February 1901, is included in Walravens & Taube (1992:342). 7 Grünwedel to Andree, 31 December 1900. Reproduced in Walravens (2001:145–146). I am grateful to Dr Walravens for the following additional comment (in an e-mail communication on 2 February 2010): “It would have been impossible to print a coloured painting in Globus and that is why Grünwedel turned it into a line drawing. He was a good draughtsman as his father had been a painter. Nowadays there is some criticism that he did not copy carefully enough and let himself be influenced by his European background. This may be true but it is very easy to criticize from today’s point of view.”

302

bray

According to Francke, the figure in the top centre riding a ‘kyang-like’ horse is dBang po rgyab zhin, the lord of heaven. His wife bKur dman mo is riding a blue-maned snow lion to the right. To his far left is his son Don grub, who would be reborn in gLing as Kesar.8 His other sons are to his immediate right and left in the same ‘cloud’ as himself. In the scene below, Kesar is sitting in council with the Agus,9 maidens and noblemen of the land. Grünwedel’s own comments in the Globus article highlight Kesar’s distinctive white headgear both in his heavenly pre-existence and on earth. Francke suggested that the paintings were relatively recent in date: The pictures are not as old as many suppose. The owner assured me that they had been painted after the Dogra war. Even so, the models (Typen) in the pictures obviously go back to older templates (Vorbilder) because in this country the models of the Agus and dBang po rGyab bzhin are standard, and are also painted in the same way by painters who have never come to Leh (cited in Grünwedel 1901:282).10 It was possible to commission such paintings quite easily, and he had also sent Grünwedel a photograph of two watercolours depicting two scenes from the epic.11 The Kesar saga subsequently became one of Francke’s particular research interests, in part because he believed that it gave an insight into the preBuddhist religion of Ladakh.12 Like his Ladakhi informants, Francke came to

8 9

10 11

12

Stein (1959:96) suggests that this figure might have been a holy lama, not Don grub. In the Ladakhi version of the epic, the 18 Agus (loosely translated as ‘heroes’) are the offspring of an ancestral figure, Dong gsum Mi la sngon mo. They themselves had miraculous births, and play a supporting role in the epic. See Francke (1901, 1902, 1905–1941:1–29). Again, my translation. The scenes are from the prologue to the epic and show Agu Palle shooting a black, demonic bird, and a scene from slightly later in the story showing dBang po rgyab zhin. The paintings are reproduced in line drawings in the Globus article. The same or similar paintings are now in the Völkerkunde Museum in Herrnhut, Germany. See also IckeSchwalbe (1990:48). Contemporary scholarship takes a more nuanced view. It is possible that the various versions of the epic contain elements that are indeed pre-Buddhist, or related to nonBuddhist cultures such as those of Iran. However, scholars now question whether it will ever be possible to identify an ‘original’ Kesar. The various oral versions, and the texts that are based on them, show evidence of repeated reformulation long after the arrival of Buddhism in Tibet.

the lost paintings of kesar

303

believe that the epic had originated in Ladakh.13 With the help of local assistants he arranged for oral versions of the Ladakhi Kesar saga to be transcribed and published, first in the Mémoires de la Société Finno-Ougrienne (1900, 1902) and then the Indian Antiquary (1901) and the Journal of the Asiatic Society of Bengal (1906, 1907a). A consolidated version of the papers containing his various researches into the saga was published by the Asiatic Society of Bengal in instalments from 1905 onwards. It finally reached completion in 1941, eleven years after Francke’s death, and represents one of his prime scholarly achievements. Having started his missionary service in Leh, Francke went to Khalatse in Lower Ladakh from 1899 to 1906, and then for a further two years in Kyelang (Lahul) from 1906 until 1908. He then had to retire from formal missionary service because his wife’s ill-health obliged him to return to Germany. However, in 1909 he returned to India for an assignment with the Archaeological Survey of India (ASI) where he worked under the supervision of the Dutch scholar Jean Philippe Vogel (1871–1958).14 His task was to conduct an initial survey of archaeological monuments in the western Himalayan border regions, and in June 1909 he set out on a journey which took him up the Sutlej valley as far as Shipke, just inside the Tibetan border, and then via Spiti to Ladakh and Kashmir. On this journey he was accompanied by a skilled ASI photographer, Babu Pindi Lal. Francke’s account of his researches was published in the first volume of his Antiquities of Indian Tibet (1914), and illustrated with Pindi Lal’s photographs. When Francke reached Leh in late August 1909, he took the opportunity to return to the paintings at the Kalon’s house in Changspa. He describes his visit as follows: In a garden-house (rab-gsal) are the remains of frescoes illustrating the Kesar-Saga. Several years ago, I ordered one of them to be copied by a local painter. This time, I had photos taken of three of them, in addition to a copy in colours executed by a local artist. The frescoes in the garden house will soon be gone altogether, and as pictures relating to the KesarSaga are very rare, I was resolved to save for science what could be saved (Francke 1914:79). 13

14

Stein (1956:164–168) is very dismissive of Francke’s thesis, arguing that the Lower Ladakhi oral version of the epic nevertheless shows traces of literary influence, as well as geographical references to Eastern Tibet. For a more recent discussion of Francke’s hypothesis, see Jettmar (1990): he suggests that the epic might after all have had historical roots in the wider Ladakhi region. On Vogel see Theuns-de Boer (2008).

304

bray

Figure 12.2

The first of the Changspa Kesar paintings. Courtesy Kern Institute, Leiden University, ms nr. XI. fol. 31.

Babu Pindi Lal’s photographs from Changspa were not included in Antiquities of Indian Tibet. However, Francke subsequently published two of the photographs in his book on Ladakhi marriage songs (Francke 1923: plates 15 and 16), and original prints survive in the British Library in London and in Vogel’s papers at the Kern Institute at the University of Leiden.15 The images below come from the Kern Institute. The right-hand part of the first photograph clearly represents the scene which was published by Grünwedel in Globus, but with some additional details that were not included in the earlier sketch. Below the ‘council scene’, we can see three dancing figures. To the left there is what appears to be another scene altogether where a high-status figure is seated in a tent while servants or followers crouch deferentially in front of him. In his 1923 caption, Francke identifies it as a scene from the Dogra war, but without giving any further 15

The images in the British Library are original prints and are included in: Archaeological Survey of India (1910), catalogue reference: Photo 1006/1 (539–541). I am grateful to Matthew Akester for drawing these images to my attention. The ASI’s reports were distributed to major libraries in India and Europe, and should therefore be more widely available.

the lost paintings of kesar

305

elucidation. The scene takes place in a Tibetan-style tent, but could conceivably represent the Dogra general Zorawar Singh. However, this attribution is purely speculative. Francke’s observations in Antiquities of Indian Tibet focus on the right-hand side of the picture, and are similar to his comments recorded in Globus: The following notes on the dress of the pre-Buddhist divinities represented in the frescoes may be of iconographical interest: dBang-po-rgyabbzhin has a red coat, and white cloak with blue seam; Gog-bzang-lha-mo16 is dressed in white but her trousers are red and she has a green shawl; one of her sons (Don-yod?) has a red jacket with green seams; another of her sons (Don-ldan?) has a white and green jacket and red trousers; Kesar has a red coat and a white cloak with green seams; he sits on a red carpet, and the background behind him is blue. In his analysis of the paintings (Francke 1914:79), he draws a distinction between the oral version of the epic and the various literary versions. All the frescoes in this hall refer to a chapter of the Kesar-Saga, entitled Ljang-dmag, “the war against the country of Ljang”, as we find it in the literary version of the Kesar-Saga. Let me remark that the famous epic of Tibet, the Kesar or Gesar Saga, is preserved in two versions which are very different from one another.17 One of them, the oral version, exists only in the mouths of the people, whilst the literary version is found in several manuscripts in Ladakh, and possibly even in woodprints. All my publications with regard to the Kesar-saga deal only with its oral version. A manuscript of the literary version of the saga is in the hands of the present ex-minister of Changspa.18

16

17 18

In the Ladakhi versions of the epic recorded by Francke, Gog-bzang-lha-mo is Kesar’s earthly mother. See: Francke (1905–1941:30–72; 380–381). Here he seems to be using the name to refer to dBang po rgyab zhin’s wife in heaven, bKur dman mo. This is odd since his own text implies that they are distinct personalities (Francke 1905–1941:70). The suggestion that there are only two versions is of course a gross underestimate. From the evidence that Francke presents here, it is not clear why he thinks that the frescoes refer specifically to the war against Ljang since all the characters mentioned here appear in the Ladakhi oral versions of the epic that he had himself recorded. The answer may lie in his interpretation of the other paintings in the building, or in information acquired from his Ladakhi hosts. Unfortunately, he does not elaborate.

306

bray

Figure 12.3

The two warriors. Courtesy Kern Institute, Leiden University, ms nr.XI. fol. 31.

The last sentence obviously refers to the same text that was transcribed for Jäschke. Francke evidently had further copies made for the ASI and for his own use.19 The second photograph (Figure 3) illustrates the scale of the damage to the paintings: the plaster is cracked and has suffered from rain leaking in through the roof. It shows two figures on horseback, fighting each other with spears. It presumably represents a battle scene from the epic but it is not clear precisely which episode. To the left of the painting there is a building with a written text below. As far as can be deciphered, this says gral tshog gyi gon pa, which appears to mean

19

See Francke (1914:118). This is a list of the manuscripts collected during his expedition including: “Tang-dmag [sic]—the war against Ltang [sic], copied from MS containing the literary version of the Kesar-saga, in the possession of the bKalon at Changspa by Joseph Thse-brtan of Leh.” For the reference to Francke’s own copy see Walravens & Taube (1992:513, 525). Stein (1956:82,83) cites manuscripts of the Hor and lJang episodes of the saga which were collected in the Rupshu region of Ladakh by George Roerich, but which bear the imprint of the Kham dialect.

the lost paintings of kesar

Figure 12.4

307

Close-up of one of the mounted warriors.

“Monastery of Drel Tshog or Drel Tsag”.20 Below the lower of the two mounted warriors, it is tentatively possible to decipher the words pa’i nye bum meaning “of nyi bum”. This could be a misspelling of nyis ’bum, meaning “of two hundred thousand suns”, or nyi bum could refer to a place or tribe. It appears that in eastern Tibet gSer ba’i nyi ’bum dar yag is known as one of Kesar’s 30 warriors. 21

20

21

I am grateful to Tsering D. Gonkatsang for his deciphering of this text. In a personal e-mail communication to the author (12 January 2009), he adds: “Gral means a ‘row’ or ‘column’, as when monks sit in a row. Tshog could be misspelling of tshogs which means ‘gathering’, ‘congregation’, ‘assembly’. However, I presume there is a vowel ‘o’ over the letter tsha although I cannot see it clearly. It could of course be tshag without the ‘o’. Gyi is the genitive ‘of’. If the preceding word is tshog the correct genitive should be gi and if the word is tshogs, the genitive should be kyi. Gon pa is presumably a misspelling of dgon pa meaning ‘monastery’”. I am grateful to George FitzHerbert for the following information: “In the Gesar temple at Khra gling monastery (near Darlag/Gyurme/Daru xian in mGo log), sGer ba’i nyi ’bum dar yag is represented as one of the 30 Warriors, and identified as an emanation of the Mahasiddha Ku ku ri pa. Mkhan po ’Jigs me phun tshogs, the modern charismatic lama and founder of Lha rung sgar in gSer rta (Kandze Prefecture, Sichuan), recognized bsTan pa’i nyi ma, the head lama of Khra ling monastery, as an emanation of this warrior. The same warrior, gSer ba’i nyi ‘bum dar gag, is listed in gCod pa don grub & Bsod los (1996:39), a key reference work on Gesar.” Personal e-mail communication, 18 May 2010.

308

bray

Figure 12.5

The palace. Courtesy Kern Institute, Leiden University, ms nr.XI. fol. 31.

The third photograph shows a palace or castle. Again, the painting is badly damaged, with marks showing where water has run from the ceiling. Part of the plaster appears to have fallen away at the right of the picture. According to Francke, this image depicts Kesar’s palace in gLing. In the middle of the first floor there is a high-status figure surrounded by attendants and, as will be seen, this is likely to be Kesar himself. In the sky to the centre left there is a cloud, and one can just make out the halo of what must be a deity. In Ladakh the roofs of both monasteries and houses tend to be flat. By contrast, the roofs of the building in the painting are sloping, and curved at the end in an architectural style that is more typical of central and eastern Tibet, for example the Jokhang temple in Lhasa.22 As noted above, Francke mentioned that he had had the murals copied by a local artist. It turns out that the Völkerkundemuseum at Herrnhut (Germany) 22

For the Jokhang roofs see, for example, the photograph taken for Sir Charles Bell in 1921: The Tibet Album. “Spectators in Jokhang at Monlam Chenmo”, Pitt Rivers Museum. Accessed 21 Feb. 2010 http://tibet.prm.ox.ac.uk/photo_1998.285.213.html. For another Lhasa comparison, see the photograph of the Eastern Courtyard in the Potala, taken by Frederick Spencer Chapman in 1936: The Tibet Album. “Deyang Shar, Potala”. The British Museum. Accessed 20 Feb. 2010, http://tibet.prm.ox.ac.uk/photo_BMR.86.1.12.1.html.

the lost paintings of kesar

Figure 12.6

309

Painting from the Völkerkundemuseum Herrnhut. Catalogue. No. 67559.

possesses a watercolour presented by Francke which is described in the catalogue as “Ansicht des Königspalastes in Lhasa”: ‘view of the royal palace in Lhasa’.23 However, on close inspection, it is seems very likely that this is an eye-copy of the Changspa mural—not perfect in every detail but with enough in common for the resemblance to be unmistakable. The watercolour is not a precise reproduction of the Changspa painting: for example, one window is missing on the lower left. However, it is close 23

The catalogue states that Francke lent the painting to the museum in 1905, and sold it in 1909. If the entry is correct, it must be an earlier copy of the Changspa painting, not the one made in the summer of 1909. A black-and-white image of the painting was included in an article on the Herrnhut museum by Lydia Icke-Schwalbe (1990:43), with a caption describing it as the “king’s castle in Leh”. However, as noted above, this ascription is unlikely.

310

bray

Figure 12.7

Detail of painting in Völkerkundemuseum Herrnhut.

enough in other respects for the resemblance to be clear. The copy is helpful not only in showing the colours, but also in giving a clearer view of the figures in the centre of the picture. The central figure with his left hand raised is wearing a white, three-cornered hat which appears to be characteristic of Kesar (Stein 1956:342–345).

Discussion and Conclusion

The surviving information on the Changspa paintings is meagre enough, but nevertheless enriches our understanding of the development of Ladakhi art in at least three important respects.

the lost paintings of kesar

311

The first is that it provides what is now a rare 19th century example of the iconography of Kesar.24 Such examples may once have been more common. As noted above, Francke himself comments that the Changspa murals’ depiction of Kesar is based on established templates that were known ever to artists who had not been to Leh, and indeed that it was possible to commission such paintings relatively easily: the collection of the Völkerkunde Museum in Herrnhut contains at least one watercolour example (Icke-Schwalbe 1990: 39, 48). To my knowledge, no other 19th or early 20th century examples from Ladakh survive. Secondly, on a related point, the murals points to the existence of a secular— or at least non-monastic—style of narrative wall-painting. Such paintings would certainly have been a mark of high social status, but may once have been more widespread in Ladakh, at least in the houses of people prosperous enough to afford them. One other recorded example comes from Roger Lloyd Kennion, who served as British Joint Commissioner in Ladakh at the turn of the 19th and 20th centuries, and reports that his official residence in Leh had been richly decorated in the 1870s on the orders of one of his predecessors: The walls were frescoed by lama artists in the brilliant pigments used in Tibetan monasteries. Those in the hall and climbing up the stairs represented the Dogra invasion of Ladak [sic], with elephants, cavalry and foot-soldiers climbing over the diminutive green and white pyramids which conventionally represent the snow-crested Himalaya; while upstairs the walls of the passages and of the dining room were done in rectangles, each one depicting a Tibetan fairy tale (Kennion 1910:176).25 I am not aware of any comparable murals that survive to the present day in Ladakh, but this is perhaps scarcely surprising. In Tibetan and Ladakhi monasteries it has long been customary to paint over sacred paintings when they become too worn. Lay people would have had even fewer inhibitions to prevent them doing the same to decorative paintings in their own households.

24 25

Stein (1958) and Macdonald & Pema Tsering (1979) discuss eastern Tibetan thangkas of Kesar, but the style is entirely different. The British Joint Commissioner’s residence was in the Karzu district of Leh, and is now the army’s Alpha Mess. The Völkerkunde Museum collection in Herrnhut also contains another Ladakhi image of the Dogra war which is reproduced in Icke-Schwalbe (1990:41).

312

bray

Contemporary members of the Kalon family have no inherited memory of the Kesar paintings. Indeed, they express surprise that they could ever have existed.26 It seems that Gyalpo Skunma, the protective deity of the household, is a Hor and therefore one of Kesar’s enemies. They had therefore had nothing associated with Kesar in the household, and one family member even recalled an aunt hurriedly switching off the radio when the epic was recited on the local radio station. Kesar’s incompatibility with Gyalpo Skunma may have provided an extra reason for neglecting and ultimately destroying the paintings. However, it also raises the question why they were created in the first place. Very speculatively, one wonders whether the murals might have been considered more acceptable because they depicted Kesar’s war against lJang, rather than against the Hor. It is unlikely that there will now be a definitive answer. Finally, the reference to lJang and therefore to Kham, which at that time was several months’ journey away on the opposite side of the Tibetan plateau, is itself of interest. It is possible—although Francke does not explicitly draw this connection—that a former Kalon had commissioned the murals because he had been inspired by the written text. We can only speculate how the link with Kham came about, but there are a number of possibilities. Members of the Kalon’s family almost certainly travelled to Lhasa from time to time, and might have encountered Khampas there. For example a Kalon Tsepal led the lo phyag mission from Ladakh to Lhasa in 1870/1871, and it is likely that he was the then head of the family (Bray & Gonkatsang 2009:101–103). Similarly, it was not unusual for travellers from Kham to come to Ladakh in their turn, perhaps as part of a wider pilgrimage. A representative example is Paul Sherap (Dorje Zodba) from the eastern Tibetan border town of Tachienlu (also known as Dartsendo and now as Kanding), who visited Ladakh in the late 19th or early 20th century (Combe 1926:174).27 Even the Rong btsan, the deities who take possession of the two oracle monks of Matho monastery during its annual festival, are thought to have come from Kham (Dargyay & Gruber 1985). Kesar is very much a local hero in Ladakh, and the many Ladakhi variations of the epic are a subject for study in their own right. At the same time, the Changspa paintings serve as reminder of the richness and complexity of the constant cultural interplay between Ladakh and the wider Tibetan world.

26 27

These comments are based on informal interviews in July 2003 and July 2009. Paul Sherap later served as one of R.A. Stein’s main informants in the latter’s researches on Kesar in eastern Tibet.

the lost paintings of kesar

313

Acknowledgements I am very grateful to the following for comments and suggestions on earlier drafts of this paper: the late André Alexander, Aaron Dana, George FitzHerbert, Tsering D. Gongkatsang, John Harrison, Katia Buffetrille, Erberto Lo Bue, Guido Vogliotti, and Hartmut Walravens. I gratefully acknowledge permission from the Völkerkunde Museum in Herrnhut and the Kern Institute in Leiden to reproduce images in their custodianship. All errors remain my own responsibility.

Tshogs zhing: A Wall Painting in the New ’Du khang of Spituk (dPe thub) Filippo Lunardo Introduction Tshogs zhing is a compound consisting of the terms tshogs: ‘assembly’, ‘mass’, ‘group’, or verbs such as ‘to gather’, ‘to collect’, etc.; and zhing, ‘field’ in the sense, for example, of a farmer’s field (Yablonsky 2000:49–50). Within specific instruction and liturgy literatures, both tantric and exoteric, this compound refers to the visualization of gurus conceived as a complex spiritual field. By means of such visualization, the meditator receives from the assembly of masters and deities the blessings, inspiration and transformative energy necessary for the accumulation of spiritual merit and for the destruction of negativity and obstacles, aimed at obtaining the goal set by the instruction that is put into practice. The Tshogs zhing, basically defined as ‘field of the accumulation of merit’, envisages the visualization of a root guru, the one from whom the lineage or religious order has sprung, as its main figure, considered as fully divine. All around this figure there are the lineage holding masters, and masters associated with specific experiential practices which are in any case related to the lineage. Besides these groups of figures are found different classes of deities, ranging from the main yi dam of the anuttarayogatantra cycle down to the divinities of lesser spiritual level considered to be the protectors of the Buddhist religion in general and in particular of the path of the spiritual practitioner. All the figures of the tshogs zhing, always considered as emanations of the root guru, are visualized as positioned on a wish-fulfilling tree that functions as an axis mundi whose various parts symbolize different experiences of the Buddhist Dharma.1

1 In the dGe lugs pa order of Tibetan Buddhism, the literature that describes the visualization of tshogs zhing in relation to tantric instructions is that of the bla ma mchod pa, which is essentially linked to the liturgy that revolves around devotion to the guru. On the other hand the tshogs zhing linked to instructions unrelated to initiations is that which is prescribed especially in lam rim literature, the gradual path to awakening.

© koninklijke brill nv, leiden, ���4 | doi ��.��63/9789004271807_015

tshogs zhing

315

The tshogs zhing Tradition within the dGe lugs pa Order Within the dGe lugs pa order there are three traditions relating to the visualization of the field of the accumulation of merit. According to the first, the root guru is visualized in the likeness of the Buddha Śākyamuni. In this tradition he represents every type of master and deity, and thus comes to be visualized alone and not surrounded by any other figure. In the tradition stemming from the lam rim tradition the visualization of the field of accumulation of merit is arranged around the figure of the guru visualized as the Buddha Śākyamuni, though surrounded by lamas and deities. This type of tshogs zhing is called khrom tshogs. In the tradition stemming from tantric and exoteric liturgy and practices connected to the guru devotion, that is to say the bla ma mchod pa tradition, the tshogs zhing has the founder of the dGe lugs pa order itself, Tsong kha pa, as its principal figure. Tsong kha pa appears as an emanation called Bla ma Blo bzang rdo rje ’chang, surrounded by masters and deities. Such type of tshogs zhing goes by the name of mtho brtsegs. In the lam rim and bla ma mchod pa literature, the visualization of the tshogs zhing belongs to the so-called preliminary practices, sngon ’gro, despite the fact that the meditator continues with them throughout the whole course of his or her practice. In the lam rim tradition the visualization of the field of accumulation of merit belongs to the fourth of six preliminary practices known as sbyor chos2 (Dalai Lama 1991:22, 28–39.), whereas, in the case of the bla ma mchod pa, the visualization of the tshogs zhing (Panchen Lama 2003:5–9; Dalai Lama 1996:76– 111; Pabongka Rinpoche 1997:157–161, 768/769) follows the literature of the taking of refuge, the tantric practice of the practitioner’s self-generation in the form of a yi dam, usually two-armed Vajrabhairava, and the blessing of offerings that are considered to be derived from the union of bliss and emptiness.

The Artistic Tradition

Both in relation to lam rim instructions and to those of the bla ma mchod pa, the iconographic codification of the visualization of the tshogs zhing is structured around the wish-fulfilling tree. This functions as the actual vertical axis of the entire image. According to the instructions found in the relevant literature, both the Buddha and Tsong kha pa should be depicted seated at the 2 sbyor ba’i chos drug.

316

lunardo

centre of the tree, but in fact this only occurs in the oldest images, or at least in those which refer back to the earlier textual sources3 (Figure 13.1). In effect, by the beginning of the nineteenth century the two figures come mostly to be represented above the tree (Figure 13.2). To the right of the principal figure, there is a representation of the Bodhisattva Maitreya along with the masters and lamas linked to instructions related to the bodhisattva ethics as this has been handed down from the Indian Yogācāra philosophical school. To the left, by contrast, we find a representation of Mañjuśrī along with the masters and lamas connected with the philosophical vision of the Madhyamaka school. Directly above the head of the root guru we find Vajradhāra along with the masters holders of the tantric practices. These three groups of figures are not represented within the bounds of the tree, but in the area of the sky around. Below the head guru, subdivided into eleven iconographic groups arranged on eleven strata of an enormous lotus (which as a matter of fact masks the whole figure of the tree) there are various classes of deities: the yi dam of the four classes of the tantric cycles,4 Buddhas, bodhisattvas, pratyekabuddhas,

3 In relation to the images relative to the bla ma mchod pa tradition, these have undergone iconographic and compositional evolution in the course of time, such that they reflect the same evolution undergone in the bla ma mchod pa literature itself. The practices of bla ma mchod pa derive from the oral instructions defined as essential, man ngag, and later codified in a specific literature specifically linked to a literary field better known as guruyoga. Bla ma mchod pa literature has undergone important change and evolution over the centuries, above all in regard to the explanations and methods of practice, but leaving the philosophic and spiritual vision unchanged. This sheds light on the fact that a living tradition maintains itself thanks to its own internal change, linked to the personal experience of the masters and yogins within the lineage who practice or realize the instructions to then be transmitted directly on to their disciples. In the course of our studies we have been able to identify at least three chronological phases into which to classify three different phases of tshogs zhing. Just as in the texts the images show changes in the form of various elements of greater or lesser importance, it has been noted that the first representations appear some centuries after the text to which they refer. The most ancient tshogs zhing related to the bla ma mchod pa can be dated to the end of the 18th century or beginning of the 19th century. The figure of Tsong kha pa comes to be represented precisely at the centre of the composition, thus within the bounds of the wish fulfilling tree. However, we do not see any structural order for the representation of the various masters and surrounding the root guru. Such an order codified at the beginning of the second iconographic phase relative to the 19th century, tends to represent and identify a distribution of the various figures according to a spiritual hierarchy. However, this will be better explained in a forthcoming work. 4 Anuttara, yoga, caryā and kriyātantra.

tshogs zhing

Figure 13.1

The tshogs zhing’s oldest iconography, late 18th century/early 19th century.

317

318

Figure 13.2

lunardo

19th century tshogs zhing, Courtesy of Renzo Freschi Oriental Art.

tshogs zhing

Figure 13.3

19th century tshogs zhing, Courtesy of Renzo Freschi Oriental Art.

319

320

lunardo

arhats, ḍākas and ḍākinīs, the dharmapālas and, usually, but not always, outside the tree, the guardians of the four directions (Jackson 2005:10–11). Underneath the tree, which is represented as springing from an ocean of special waters (Dalai Lama 1996:80),5 there are the symbols of the treasure of the universal monarch, of the universe with the cosmic Mount Sumeru at the centre of the four continents and eight subcontinents—based on the cosmological conception of Vasubandhu’s Abhidharmakośa (Brauen 1997:18–21)— and of a haloed monk (Jackson 1996:236–241; Linrothe 2001:30),6 alone or with other monks and lay devotees, who is represented in the act of the maṇḍala offering (Figure 13.3).

A Wall Picture in the New spyi khang of Spituk Monastery (dPe thub) in Ladakh

During some of our research in Ladakh, carried out since 2006, we have been able to document an important evolution of the tshogs zhing iconographic themes connected to the bla ma mchod pa liturgy in the new spyi khang of Spituk monastery (Thub bstan dpal ldan 1990:112–116; Snellgrove and Skorupski 1977:109). First of all, the field of accumulation of merit is not depicted on a thang ka but on a wall, which is an unusual example according to our documentation for these images. The picture is painted on the main wall opposite the entrance to the hall. The image is not covered, in contrast to the thang kas which, if they are not on show in the small museums of monasteries, are usually kept covered away from the eyes of the uninitiated. However, the great innovation in the Spituk mural is the fact that it does not show at all the wish-fulfilling tree, which is the vertical axis of the image and the space where the deities would be positioned. In the spyi khang the 5 The 14th Dalai Lama explains that it represents an ocean of milk, which symbolism is tied to the three principal aspects of the Buddhist path. In any case, the water element connects with the theme of cosmic waters, the container of all that is necessary for life, much beloved of Indian cosmology. 6 The representation of this figure immediately recalls a specific iconographic form, related to a series of representations belonging to the lineage of the Paṇ chen lamas executed at sNar thang by Tsong kha pa’s disciple mKhas grub dGe legs dpal bzang po. In the course of the time, within the dGe lugs pa order, such iconography, though always linked to the presence of an image of Tsong kha pa, has also been employed to portray in a generic from the category of practitioners, perhaps even patrons of the tshogs zhing images themselves, shown while looking at the field of merit in a devoted way while carrying offerings.

tshogs zhing

Figure 13.4

321

Central portion of the Spituk/dPe thub New ᾿Du khang tshogs zhing.

vertical structure of the tshogs zhing is exchanged for a structure having three horizontal registers superimposed on one another where the groups of masters and classes of deities have found their place, arranged surrounding the chief figure of Tsong kha pa, which is in turn immediately surmounted by that of Vajradhāra (Figure 13.4). Thus here the representation of the tree is wholly missing. As for the commissioning of the image, both the artist who executed the mural painting, Tshe ring dbang ’dus (Lo Bue 2005:365–373),7 as well as several monks of the monastery, confirmed that it was the former Bakula rin po che and his root master who wanted the work to be executed. This was carried out at the time of the building of the spyi khang at the beginning of the 1970s. In any case, neither the monks nor the artist were able to give an explanation as to the variation in the iconography that is to be seen in the image. Nor were they able to provide any information as to the existence of written records relative to the wishes of the Bakula rin po che in relation to the mural painting. It therefore appears that further investigation is necessary. Nevertheless, both the artist and the monks have indicated that the image shows the visualization of the field of merit in line with the spyor chos related to the lam rim tradition. In this way it constitutes an interesting evolution within 7 We would like to use the occasion to thank Tshe ring dbang ’dus for the helping us during our research, and in particular for the interview given in September 2008.

322

lunardo

the iconography of such images. In the first place, the chief figure of the tshogs zhing of the spyi khang is that of Tsong kha pa and not that of Śākyamuni, as would be expected in the lam rim tradition. Moreover, the iconography of Tshong kha pa here suggests that it follows the bla ma mchod pa tradition, where the root guru takes the name of Bla ma Blo bzang thub dbang rdo rje ’chang (Dalai Lama 1996:87/88; Pabongka Rinpoche 1997:188–192),8 even though in the case of Spituk according to the monks the iconography would not exactly fit that of the bla ma mchod pa, since in the heart of the small Buddha who this time is positioned within Tsong kha pa’s heart, the image of Vajradhāra is missing and replaced by the seed syllable hūṃ. According to the monks there, this would be due to the fact that the presence of Śākyamuni instead of the Buddha Vajradhāra within Tsong kha pa’s heart is a clear indication of specific lam rim practices (Figure 13.5). However, we need to note here that, linked to the tradition of Pha bong kha bDe chen snying po (Pabongka Rinpoche 1997), the modern texts of the bla ma chod pa and the tshogs zhing visualization related to these practices always require the presence of Śākyamuni in the heart of the central guru, whereas Vajradhāra is to be visualized in the heart of

8 Bla ma indicates that the root guru has to be identified with the practitioner’s own guru; Blo bzang reminds us that in the dGe lugs pa’s tradition guru has to be identified with Tsong kha pa himself; Thub dbang indicates that the practitioner’s guru, as Tsong kha pa himself, has the same nature of the Buddha Śākyamuni when he granted to his disciple instructions based on sūtras; rDo rje ’chang indicates that the guru, as Tsong kha pa himself, has the nature of Vajradhāra in giving instructions based on tantras. All those are different emanations of the same essence: an enlightened mind that appears as Śākyamuni/Nirmāṇakāya in the exposition of the sūtras and as Vajradhāra/Sambhogakāya in the exposition of the tantras. In the exposition of the entire Dharma, it appears having the aspects of great masters such as Tsong kha pa and as the practitioner’s own guru that represents in essence all the Buddhas as expressions of the wisdom of the Dharmakāya. This complex experience is usually represented in the tshogs zhing iconography by the image of Tsong kha pa seated in the varja posture, making the gesture the vitarkamudrā with his right hand and the dhyānamudrā with the left one. The hands hold two lotuses holding up a sword and a book, emblems of Mañjuśrī, and a bowl in the left hand. In the heart of Tsong kha pa, a small Śākyamuni is represented and, in the heart of him, Vajradhāra with the consort. This visualisation is very important For the bla ma mchod pa instruction, because, according to Pha bong kha bDe chen snying po, Tsong kha pa and Śākyamuni at his heart have to be experienced as the samayasattva, together they are the commitment-being, Vajradhāra in the union with his consort, representing the great bliss wisdom that experiences emptiness, and which has to be experienced as the jñānasattva, the gnosis being, and the hūṃ at the heart of this has to be meditated on as the samādhisattva, the concentration-being.

tshogs zhing

Figure 13.5

323

Detail of the main figure of the tshogs zhing, Tsong kha pa.

the historical Buddha.9 Pha bong kha describes various traditions in relation to the visualization of this field of merit, also indicating that it was actually his root master, the dwags po Blo bzang ’jam dpal lhun grub (1845–1919), who unified them into one unique transmission which was handed down by Pha bong kha in his teachings. From these the most widespread tshogs zhing of today originated.10 The field of merit of the spyi khang of Spituk, by way of its 9

10

In the bla ma mchod pa’s literature tradition previous to Pha bong kha bDe chen snying po, the term thub dbang was meant to be an adjective understanding Vajradhāra: thub dbang translates the Sanskrit munīndra, the lord of the sages, a term which can be used for every Buddha of any time and dimension. The 14th Dalai Lama, expounding on a commentary on the bla ma mchod pa, understands the term thub dbang in the same way as Pha bong kha, i.e. indicating the figure of the Buddha Śākyamuni. In a private communication in Rome, in the September 2008, the dGa’ ldan byang rtse’s dge bshes Thub bstan dar rgyas, explained to us that the meaning of the visualization of Vajradhāra in the heart of Śākyamuni is that Vajradhāra, here thought of as a Sambhogakāya experience of the tenth bhūmi, cannot be directly perceived by conventional minds: those minds are able to directly perceive only an experience generally considered easier to approach and understand. The Nirmāṇakāya/Śākyamuni is the only form of the Sambhogakāya experience a conventional mind can approach, and, once again, the guru, following Śākyamuni, is the best embodiment of the Nirmāṇakāya a practitioner can relate to. The root guru sits in the centre of the entire image: below there are the eleven categories of deities, placed on the tree; at the sides there are the Madhyamaka and Yogācāra gurus; below we find three sub-lineages of the bKa’ gdams pas and, below those, a group of dGe lugs pa masters, starting from Tsong kha pa. Above the root guru there are five vertical

324

lunardo

unusual iconography, may thus represent yet another evolution of this theme, unrecorded in any text, though it seems that the specific image was painted more in order to indicate the lam rim practices than those of the bla ma chod pa. Further studies are necessary in order to come to a better understanding of this problem.

Analysis of the tshogs zhing Iconography of Spituk’s New spyi khang

The key figure of the whole imagery, placed at the centre of the mural, depicted with a greater size in relation to all the other figures,11 is that of Tsong kha pa. All around him, on the central register, feature the four yi dam connected with the most important anuttarayogatantra cycles of the dGe lugs pa order. Directly to the guru’s right, we find Guhyasamāja12 and Vajrabhairava, linked to the so-called father tantra classes, while on the left we find representations of Cakrasaṃvara and Hevajra, linked to the classes of the so-called mother tantra. Above Tshong kha pa’s halo, on the upper register, Vajradhāra is represented with his consort in the yab yum form. To his left is the siddha Tilopa, while to his right is the siddha Nāropa. These three figures together represent the entire tantric lineage of practices and blessings. Apart from Vajradhāra, the Buddha who confers the tantric teachings and who is the one who symbolizes the ultimate experience shared by the tantric systems of the gsar ma traditions, it is necessary to note that just as in the bKa’ brgyud pa orders, Tilopa and Nāropa are venerated as the principal tantric masters who connect the dGe lugs pa order to the Indian tantric tradition. Beneath the root guru there are three figures, all seated. The central figure represents Tsong kha pa himself, however the iconography represented here shows the attributes of the root guru, i.e. the sword and the book, both held by the lotus whose stem originates from the right hand of the master, and the flask containing the amṛta, placed in Tsong kha pa’s left hand which is resting in his lap. Such iconography represents a special form of the master used for a particular long-life pūjā. To the sides of Tsong kha pa we find his direct disciples: mKhas

11

12

rows of different lineages of tantric transmission: the Guhyasamāja, Vajrabhairava, dGa’ ldan snyan rgyud, the Sixteen Drops of the bKa’ gdams pas, and Cakrasaṃvara. The rules on proportional symbolism apply here. Furthermore, we need to know that according to the bla ma mchod pa’s instructions: all the figures in the field of merit, representing masters or deities, ought to be experienced as emanations of the root guru. This deity is considered as “king of the tantras” by the dGe lugs pas.

tshogs zhing

Figure 13.6

325

The portion of the tshogs zhing to the proper left of the main, central figure, Tsong kha pa.

grub dGe legs dpal bzang po, on his right, and rGyal tshab Dharma rin chen on the left. In the dGe lugs pa iconographic tradition such a triad as this tends to represent as well the main deities of the Mahāvairocanābhisaṃbodhitantra, a tantra of the caryātantra class. Tshong ka pa represents the Buddha Mahāvairocana, while his disciples embody the Bodhisattvas Avalokiteśvara and Vajrapāṇi, shown immediately to the sides of the triad. Preceding the Bodhisattva Avalokiteśvara there is an image of a dGe lugs pa master depicted as larger in size to the triad, but nevertheless there are no inscriptions that permit his identification.13 On the upper register, to the left of the siddha Nāropa (Figure 13.6) there is the Bodhisattva Mañjughoṣa, and to his left are depicted the images of nine masters linked to Nāgārjuna’s Madhyamaka tradition. The last figure on this register is that of the Bengali master Atiśa, and below this in the identical position on the lower register is his direct disciple ’Brom ston pa. On the central register, to the left of the yi dam Hevajra, are found arranged the figures of Sarvavid, Mahāvairocana, Green Tārā, Sarasvatī and Śākyamuni, while on the lower register, to the left of the Bodhisattva Vajrapāṇi, are found the images of the śrāvaka Nanda and that of the arhat Bakula, who is recognizable because of the image of the mongoose. This is followed by a white ḍāka who holds a ḍamaru, a khaṭvāṅga and a kapāla. This is probably Vajradharma (Chandra 1991:497), who is then followed by Dharmarāja without his consort, and finally the guardians of the directions Vaiśravaṇa and Virūpākṣa. 13

Thus far the monks have not been able to help us in the identification of this figure.

326

Figure 13.7

lunardo

The portion of the tshogs zhing to the proper right of the main, central figure, Tsong kha pa.

On the other portion of the image, beginning from the upper register to the right of Tilopa (Figure 13.7), there are nine masters of the Yogācāra tradition, headed by Asaṅga and Vasubandhu. Just as in the case of the figure of ’Brom ston pa, also the last master of the Yogācāra tradition is depicted as the last figure on the lower register. On the central register, beginning from the right of the yi dam Vajrabhairava, we find depicted the deities Sitātapatrā, Vajravidhāraṇā, Amitāyus, Sitatārā and Bhaiṣajyaguru, whereas on the lower register, starting from the right of the Bodhisattva Avalokiteśvara, there are the representations of the arhats Panthaka and Aṅgaja, of a ḍāka similar in colour and position to the ḍāka Vajradharma, but with a kartarī in place of the ḍamaru. Then follow the representations of Ṣaḍbhuja Mahākāla, Yellow Jambhala, and the two guardians of the directions Dhṛtarāṣṭra and Virūḍhaka. Conclusions The main iconographic elements of the traditional tshogs zhing, based on the vertical element of the tree, in the depiction of the field of merit of the new spyi khang of Spituk are confirmed by the present investigation. In the tshogs zhing, the spiritual hierarchy of the diverse classes of figures is generally respected and each class is accordingly distributed in distinct positions within the representation. The chief figure is that of the root guru, positioned in what is truly the fundamental point of the whole image, i.e. the centre. After that of the root guru, the most important class of figures is that

tshogs zhing

327

of the direct masters and the lineage masters. To them is reserved, as is natural to their spiritual status, the background sky in the upper register of the mural. The diverse classes of deities, beginning with the yi dams of the anuttarayogatantra up to the guardians of the directions14 are depicted on the various petals (eleven, as a rule) of a great lotus placed on the wish fulfilling tree. In the tshogs zhing of the spyi khang of Spituk the vertical structure of the image created by means of the tree is completely absent, and the formulation of the hierarchical division of the various classes of figure seems to be a complete innovation. No longer vertical but horizontal, we find a simple structure on three registers one above the other, with the most important figure, that of the root guru, naturally in the central position. On the upper register, which is therefore assigned the same symbolic value as is given to the sky in the vertical tshogs zhing, the area which is the most important in the image, the fundamental tantric masters of the dGe lugs pa order, Tilopa and Nāropa, as well as the masters linked to the Madhyamaka and Yogācāra philosophical traditions, are depicted. In the central register, we find the deities connected with all the four different classes of tantra, as is the image of the historical Buddha who comes to represent the entire class of Buddhas or the 35 Buddhas of the confession. Finally, in the lower register, we find representations of classes of bodhisattvas, of the pratyekabuddhas and of the arhats, as well as those of the protectors and guardians. It is certainly true that in comparison to the figures of the vertical tshogs zhings, the representations mentioned above at Spituk are far fewer in number, but this comes as a result of a deliberate choice such as having Tilopa and Nāropa selected to represent the whole group of siddhas and tantric masters in general. The most beloved masters and the most important deities for the practices adopted by the dGe lugs pa order are here taken up to symbolize whole classes of masters and deities. In any case, we can take note of the fact that in the images in Spituk, compared to modern renderings of vertical tshogs zhing which come from the Pha bong kha transmission, the figure of Hevajra is chosen and placed in the location usually preferred for representing the tantric system of “the sixteen drops of the bKa’ gdams pa”. The only representation belonging to Atiśa’s tradition is the image of the Bengali master and of his direct disciple, while in modern vertical tshogs zhing the bKa’ gdams pa order is depicted through a number of 14

The classes of the deities that are represented are: the yi dams of the cycles of the four tantric classes, the class of the Buddhas, those of the bodhisattvas, ḍākas and ḍākinīs, of pratyekabuddhas and arhats, of the protectors of the Dharma and those of the guardians of the direction.

328

lunardo

important figures divided into four sub-lineages and placed in the sky to the side of the root guru. One of these sub-lineages represents the actual lineage of the sixteen drops, portrayed in place of the Hevajratantra transmission. The tshogs zhing of Spituk is important because it reflects a modern development of an iconography that has never really been firmly established. Such an image lends support in demonstrating that over the centuries, a tradition maintains its vitality when it reflects the experience of masters and practitioners. This in the end represents a direct agent in the evolution of literature and texts relative to the tradition itself.15 It is precisely for this reason that it would be interesting to know the reason why Bakula rin po che wanted such an image in such a position, and why he commissioned the depiction of particular figures to represent entire spiritual classes. Until similar images in other areas of Tibetan culture are either discovered or described, we can affirm that the tshogs zhing of the new spyi khang of Spituk represents a distinct form of the experience of Buddhist Ladakh, and underlines the freshness and brightness of a tradition that is constantly renewing itself. 15

This will be the topic of a work in preparation.

From Benaras to Leh—The Trade and Use of Silk-brocade1 Monisha Ahmed A weaver in Benaras (now Varanasi) sits at his pit loom meticulously creating a textile piece of Mahakala, an important god of protection for Buddhists (Figure 14.1). A monk at the annual festival at Hemis monastery performs the religious dance (’cham), the Mahakala image on his apron (pang kheb) gazing out at the devotees as he pirouettes around the courtyard (Figure 14.2). These two scenes demonstrate the beginning and end of the journey of silk-brocades from Benaras to Leh.

Figure 14.1

Weaving Mahakala. A weaver in Benaras shows the image of Mahakala, the God of Protection, that he is weaving on his pit loom. Photo: Monisha Ahmed.

1 Part of this paper was published in Ahmed 2005b.

© koninklijke brill nv, leiden, ���4 | doi ��.��63/9789004271807_016

330

Figure 14.2

ahmed

The apron of Mahakala is wrapped around a statue of the same god at Hemis monastery, Ladakh. Photo: Monisha Ahmed.

from benaras to leh

331

This paper looks at the historical context of the trade in silk-brocades from Benaras to Leh, and discusses how this trade first started. It presents how these fabrics are made in Benaras and discusses their various uses in Ladakh. Finally, it examines the contemporary status of the trade and the continued importance of silk-brocades in the lives of Buddhist Ladakhis.

Silk-brocade in Ladakh

Lying at the crossroads of high Asia, Ladakh was situated squarely between some of the great mercantile towns of south and central Asia (Rizvi 1983:75). Trade flourished there, from the time the Ladakhi kingdom was established in the 10th century CE to modern times. Silk-brocades were among the most prestigious of textiles that entered Ladakh through trade.2 The high cost of these fabrics made them luxury textiles. As symbols of status their use was restricted to the royal family, nobility and the clergy. In the Ladakhi language the word for silk-brocade is gos chen (literally ‘the great garment’). However, Tibetans call it rgya ser, a term also used by the weavers and dealers of the fabric in Benaras. I use the word brocade here as a trade name and in a generalized sense use it to refer to richly patterned silk fabric characterized by the use of gold and silver thread. In a more specific context I use it to refer to a silk fabric that is patterned with Buddhist symbols such as the eight auspicious signs, the thunderbolt and the bell, and lotus flowers, as well as dragons and clouds. The most important use of silk-brocades was in the religious world of the monastery. Here, these fabrics were used for altar and seat coverings, canopies (‘u lep’), door hangings (‘cheb le’), and pillar covers (‘ka phen’). They were also used as mountings for thang kas (scroll paintings), and for making the patchwork pieces in an appliqué thang ka (Figure 14.3). The fabrics were used to make robes worn by the monks during religious dances (’cham), at monastic festivals as well as edging for their hats and boots (Figure 14.4).

2 Other high-prestige textiles were cotton and single-colored silk fabric, silk ikat, and wool-pile carpets.

332

Figure 14.3

ahmed

Appliqué and patchwork thang ka of Gyalse Rinpoche, Hemis monastery. Photo: Monisha Ahmed.

from benaras to leh

Figure 14.4

Monks, resplendent in their brocade robes, perform at the annual monastic festival, Hemis monastery. Photo: Monisha Ahmed.

Figure 14.5

A participant at a horse race is dressed in his brocade robe and hat, Rupshu, Changthang. Photo: Monisha Ahmed.

333

334

Figure 14.6

ahmed

A woman’s cape made from Chinese silk-brocade from the Shangara family. It is embellished with Chinese symbols of long life and prosperity. First part of the 20th century. Photo: Monisha Ahmed.

In the secular world brocade was used to make robes (the sul ma for women and the gos for men), capes (sbog), and hats, ti bi (Figure 14.5). While hats were worn on a daily basis, the robes and capes were generally worn on special occasions such as weddings or religious functions (Figure 14.6). The common people who could not afford to make entire robes out of brocade would use just a little bit to embellish their garments. It was the same for those who did not want to wear very heavy silk-brocade robes, especially during the day. They would stitch a strip of brocade on to the cuffs of their sleeves or the slits of their robes, or the edging around the collar of the male robe. In fact, during the time of the Namgyal dynasty the common people were not allowed to use or wear these brocaded textiles extensively.3 But gradually, after the powers vested in the royal family declined, also common people who could afford to buy these luxury textiles began wearing them.

3 It was the same in Tibet, where imported materials were luxury goods and used as marks of rank by both clergy and aristocracy (Reynolds 1981:7).

from benaras to leh



335

Early Evidence of Trade in Silk-brocades

There is no recorded evidence that indicates when exactly this trade first started in Ladakh, though it was almost certainly some time after the establishment of the Ladakhi kingdom in the 10th century. In the absence of written information, paintings on monastery walls and ceilings become a source of information regarding the historical development of textiles in Ladakh. Further, they can give an indication of the mode of dress, particularly from the style shown in representations of the donors who commissioned the work, but also, to a lesser extent, in that of the deities, as artists were free to follow the fashions and modes prevalent at that time. Some of the earliest examples can be drawn from the three-storey Sumtsek (gSum brtsegs) at Alchi, one of the oldest monastery complexes in Ladakh dating from the 11th to 13th centuries. On first impression, the ceiling of the Sumtsek looks as though it is covered by real textiles.4 The ceilings of the ground and first floor of the Sumtsek are divided into 48 panels. The panels reproduce textiles made with various techniques of manufacture, some of which were produced in Ladakh as well as others that came in through trade, such as brocade.5 The painted textiles seem very specific and realistic, and it is possible that they were used by the royalty and nobility of Ladakh at the time when Alchi was being built and decorated. It is also assumed that they are based on actual textiles available in Ladakh or at least seen by the artists in the 11th century and later, as the weaving and dyeing techniques have been realistically represented by the painters who decorated the wooden panels (Goepper 1996:225). The textile patterns which cover the ceilings are also shown on royal garments in the wall portraits in the Sumtsek, providing confirmation that these reproductions of textiles are not a figment of the painters’ imagination, but that they were in actual use. These monuments are also early evidence of the adaption of the secular garments of royalty and wealth into the spiritual realm of Buddhism (Reynolds 1997:122).

4 “The copying of textiles in paint on the ceilings may derive from the custom of fixing actual pieces of cloth under the ceilings of Ladakhi buildings, partly as embellishment but also for the practical reason of preventing dust or mud particles of the ceiling construction from falling into the rooms below” (Goepper 1996:225). This practice continues in some Ladakhi homes today. 5 Amongst the textiles that were made in Ladakh, a few panels clearly evoke the technique of tie-resist dyeing (thig ma). This comparatively simple technique is still used in parts of Ladakh.

336

ahmed

Apart from monastic murals, textiles are the basic materials for many sacred images throughout the Mahayana Buddhist world and many of these have survived (Ahmed 2005:70). For instance, some of the earliest fragments of thangkas come from Dunhuang which was an important Buddhist pilgrimage centre during the 8th and 9th centuries. Mimicking the scroll paintings there are also numerous fine 12th to 13th century pictorial banners woven in silk, known as kesi, that came to Tibet from China. While evidence of these is still not apparent in Ladakh it is probable that they exist. It is also difficult to date textiles in Ladakh because documentation on most collections (private and at monasteries) is not easily available. But textiles from the 16th and 17th centuries exist—the large appliqué and patchwork thangka of Gyalse Rinpoche, shown every year at Hemis Festival (VV.AA. Forthcoming), is said to date from the 17th century.

Trade in Ladakh: 16th to 20th Centuries

Leh, as well as Kargil, were important trading posts on a network of regional and international routes as they lay at a meeting point for trade routes from Northern India (mainly Kashmir and Punjab), Tibet, China, Russia, and Central Asia (largely Yarkand, Kashgar and Khotan). The traders that came to Ladakh brought with them carpets, felt, bales of cotton and silk fabric, brocade, silk thread, porcelain, turquoise and coral stones, gold, dyes and medicines, amongst a host of other commodities and luxury goods. Some of this merchandise they traded in Ladakh itself, and the rest they either carried with them on their further journeys or sold to traders who took it to other places to sell. As Rizvi (1999:19) has pointed out, “Though Ladakh was off the line of the classical Silk Route, the Leh-Yarkand trail represented one of the Silk Route feeders and silk was an important commodity of the trans-Karakoram trade in both directions”. Silk yarn would come from Yarkand to be woven in the Punjab, and silk from India (including brocade from Benaras) would be exported to Central Asia. Apart from the earlier evidence shown in monastic paintings of textiles that entered Ladakh through trade, we know that from at least the 16th century and definitely during the 17th century, when the first trade treaties were signed, the trade routes were firmly established. In the first part of the 17th century, King Senge Namgyal (Seng ge rnam rgyal) shifted the royal residence from Shey to Leh. He did this because he found Leh more advantageous than Shey mainly

from benaras to leh

337

because it lay at a vantage point for the trade routes that were passing through Ladakh. The Treaty of Tingmosgang (gTing mo sgang) was one of the first treaties signed between Ladakh and Tibet, after the conclusion of the Tibeto-LadakhiMughal war, in 1684.6 Under this treaty it was agreed that the western Tibetan authorities would supply their entire wool and pashmina to Ladakh, who in turn would supply this along with their indigenous produce to Kashmir (Petech 1977:75–77). Under the same treaty two missions were set up: the Lopchak (lo phyag) and the Chaba (cha pa). The Lopchak was a triennial mission that went from Leh to Lhasa, under which the king of Ladakh would send a variety of gifts, which were offerings to the Dalai Lama for his protective blessings and to Tibet for the salutation of the New Year (Ahmad 1968:354, Bray 1991:117). The Chaba was the annual trade caravan from Lhasa to Leh which carried with it two hundred animal-loads of tea (Bray 1990:78). It also specified that nowhere but via Ladakh should rectangular tea-bricks be sent from western Tibet (Marx 1894:97). This practice appears to have been followed throughout the 17th and 18th centuries. Much of the long-distance trade between Tibet and Ladakh was based on and around these two missions. Apart from the items specified by the missions, the traders who led them were also allowed to carry their own merchandise for sale in Tibet and in turn bring merchandise back from Tibet for sale in Ladakh. In addition, the leader of the mission would be accompanied by several private traders who were allowed to conduct their own business alongside his. One of the items of this trade between Ladakh and Tibet was silk-brocade that came from China. Lamas, who went to Lhasa to study, would also bring back pieces of brocade for their monastery when they returned to Ladakh. Most of the brocade came as unstitched fabric, but some of it also came as stitched robes or boots. For instance, traders would bring the chu pa—the sleeveless ankle-length robe worn by Tibetan women. Chu pas would be so large that Ladakhi women were able to redesign them into their sul ma, which is also an ankle-length robe but has long sleeves and gathers around the waist. 6 In the early 1680s relations between Ladakh and Lhasa were increasingly hostile and further acerbated when Ladakh intervened in a quarrel between Tibet and Bhutan, supporting the latter, resulting in the Fifth Dalai Lama declaring war on Ladakh (Petech 1977:70–71). The Ladakhi king, Deldan Namgyal (bDe ldan rnam rgyal), lacking the strength to fight the Tibetans, asked for and obtained the intervention of Ibrahim Khan, the Mughal Governor of Kashmir, who succeeded in stopping the Tibetan advance (ibid: 74). Though this action did not bring Ladakh directly under Mughal administration, it did bring Ladakh within the orbit of the Mughal Empire in that it was obliged to strike coins in the name of Aurangzeb.

338

ahmed

For the sleeves, they would attach sleeves by trying to find similar fabrics of the same colour and pattern. The first references to the trade in brocade are found in Moorcroft’s journals. William Moorcroft worked for the East India Company, and accompanied by George Trebeck went to Ladakh in 1820, where they spent two years, hoping to travel on to Yarkand. During their stay in Leh, Chinese brocade was among the merchandise they saw arriving on the caravans (Moorcroft and Trebeck 1841:322–6). Leh’s two wealthiest mercantile families were the Radhu and the Shangara, and at some point or the other members of these families had been the leaders of the king’s Lopchak mission.7 This practice continued right till the first part of the 20th century and ended after the last Lopchak mission went in 1946.

The Trade in Brocade from Benaras

At the same time as Chinese brocade was available in the market in Ladakh, brocade was also being imported from Benaras along the trade routes entering the region from Kashmir and the Punjab. However, the brocade from Benaras differed from its Chinese counterpart in that it was not patterned with Buddhist symbols. In addition it was not called by the Ladakhi word gos chen, but was referred to by the Urdu word kinkhwab. Kinkhwab, which literally means ‘little dreams’, is a variety of woven brocade which was popular in the Mughal courts (Kumar 1999:313). Like gos chen, kinkhwab was also a luxury textile and used primarily by the royal family and nobility of Ladakh. But, unlike gos chen, it was rarely used in the monasteries. Similar to the use of gos chen mentioned earlier, kinkhwab was also used to make women’s robes and capes, and to edge the collar of a man’s robe or the slits of a woman’s dress. As it was devoid of Buddhist symbols, the fabric was, quite predictably, more popular with the Muslim population of Ladakh. References to the trade in brocade from Benaras are given in Moorcroft and Trebeck (1841:325). Brocade from Benaras also figures in later trade reports of the 19th century (Government of Punjab Report 1862: ccix, Aitchison 1874:219). Some of the reports differentiate between the varieties of silk entering Ladakh (i.e. Benaras brocade or kinkhwab, silk); others do not.8

7 Examples of Chinese brocade traded from Tibet to Ladakh can still be seen in Ladakh among the collections of members of the Radhu and Shangara families, as well as other families. 8 Janet Rizvi, personal communication, October 2002.

from benaras to leh



339

The Introduction of gos chen in Benaras

In the first part of the 19th century gos chen was not made in Benaras. But some time around the middle of the 20th century weavers in Benaras began to copy the Chinese brocade with Buddhist motifs and patterns. They began to supply this fabric to Ladakh and other Buddhist countries such as Tibet and Bhutan, thereby breaking China’s monopoly. There are many explanations given on how this gos chen or brocade for the Buddhist market first began to be made in Benaras.9 The first account given is that Tibetan traders came to Benaras when, in the 19th century, it was already a well-established weaving centre for brocades. In fact it was India’s main brocade weaving centre. Benaras was already well known to Tibetans because of its proximity to Sarnath, the place where Buddha gave his first sermon after receiving enlightenment, and those that made the pilgrimage to Sarnath would often return with lengths of kinkhwab. Tibetans knew that kinkhwab was being woven in Benaras and thought that, if the weavers could weave that fabric, it should not be too difficult for them to weave rgya ser. The Tibetans carried some samples of the Chinese brocade with them and asked the weavers in Benaras if they could make similar fabric. The reason given for the Tibetans doing this is that they were finding Chinese brocade very expensive and wanted to look for alternative sources to meet their demand.10 Political relations between China and Tibet were strained and the Tibetan traders felt they could not always depend on the supply of gya ser from China. If they wanted their business to continue they had to look for alternative sources for the material. The second version is that traders from the Marwari community, who come from Rajasthan and are well-known in India for their business acumen, settled in Kalimpong, a large commercial market on the trade route between India and Tibet, and took the first samples of gos chen to Benaras. Some people in Benaras said it was not Marwaris but Nepalese traders, also living in Kalimpong, who brought the samples to Benaras. The actual version of what happened may never really be known but what we do know is that weavers in Benaras were amenable to weaving and copying 9 10

Most of this information is based on oral sources, mainly from families living in Benaras whose forefathers were involved in initiating this trade. In the Buddhist world Tibetans probably used the most silk-brocades. For instance, “ . . . brocades, which were always in demand by Tibetans for their formal dress, because in the administration every officer of rank, even if he were not very highly placed, had to wear this official dress” (Radhu 1997:193).

340

ahmed

the samples of Chinese brocade fabrics that were shown to them. If they had been reluctant or opposed to it in any way, gos chen might never have been woven in Benaras. The names of two brothers, Haji Kasim and Haji Mohammad Ishaque, are often mentioned as the original initiators of rgya ser weaving in Benaras.11 Members of the family, whose business now goes under the name of Kasim, recalled that, although there was opposition from the clergy at weaving another faith’s religious fabric, this did not stop them from weaving rgya ser. “It filled our stomachs then and still does,” said Haseen Ahmed, who today looks after most of Kasim’s rgya ser business (Figure 14.7).12 As it turned out, Ladakhis, Tibetans, and people from other Buddhist countries preferred the brocade that was made in Benaras. The main reason was that it was cheaper. One of the explanations given for the lower price was that labour was cheaper in Benaras as compared to China. Another reason for their partiality towards gos chen from Benaras was that the quality of its gold was better. Yet another explanation given was that the fabric was thicker and of better quality. Whatever the reasons, gos chen made in Benaras gradually gained popularity in the Buddhist world of Ladakh. Traders started to bring gos chen to Ladakh from Benaras on the old established trade routes: coming from Kashmir or the Punjab. It also came from Lhasa via Kalimpong. What is interesting is that many shop-owners and consumers in Ladakh continued to believe that this fabric was coming from China; few knew that it was being made in Benaras.13

11

12 13

The family was already well-known for their kinkhwab. In 1885 they received a first class medal at the Lucknow exhibition from the British government, and in 1886 at the Colonial and Indian exhibition in London they received a medal for their fine kinkhwab from Prince Edward. Interview with Haseen Ahmed, 13th August 2002, Benaras. Tashi Challi, who first opened his shop in Leh in the early 1950s, said that he always thought that the gos chen was coming from China as that was what the Tibetan traders who brought the fabric to Leh told them. It was only when he visited Kalimpong in 1960 that he found out that the fabric was being made in Benaras. He then started ordering the brocade directly from Benaras (Interview with Tashi Challi, 3rd July 2002, Leh).

from benaras to leh

Figure 14.7

Haseen Ahmed Kasim hold up a piece of silk-brocade in his office in Benaras. Photo: Monisha Ahmed.

341

342

ahmed

The Trade in Brocade from Benaras—after 1947

After India’s independence most of the international trade routes into Ladakh closed. The last Lopchak Mission from Leh to Lhasa went in 1946.14 Members of the Radhu and Shangara family, who were once very avidly involved in trade, turned to other occupations or moved out of Ladakh. Traders from Kashmir and Punjab, whose main purpose had been to trade with Central Asia, also stopped making the journey to Ladakh. While some brocade still came via Tibet, this trade also stopped by the early 1960s. But the demand for brocade in Ladakh did not cease. Without the support of the Kashmiri and Punjabi traders or the large trading families of Ladakh, the nobility and clergy of Ladakh turned to alternative sources for their brocade. They, as well as the shopkeepers in Leh bazaar, turned directly to suppliers in Benaras. It was the same for the dealers in Benaras who, without the Tibetan, Kashmiri, or Punjabi traders to carry their brocade to Ladakh, had to turn directly to the monasteries and shopkeepers in Ladakh. Brocade dealers from Benaras, or their agents, now visit Ladakh in the summer carrying samples of their brocade with them. They do the rounds of the main cloth-shops in the market in Leh, as well as of the surrounding monasteries, and show them samples of their brocade (Figure 14.8). They give them the rates and take orders. They may also go to Kargil if they think it is necessary but, as less brocade is used in Kargil as compared to Leh, most of them do not make the trip to Kargil. They also collect any outstanding payments that might be left over from the previous year’s orders. After that they return to Benaras, where they execute their orders, and then send the fabric to Ladakh. Most of the fabric now goes as parcels, which are mailed out to Ladakh and reach there by air or road. If the dealers from Benaras are unable to go to Leh for some reason, the shopkeepers in Leh, as well as the monks, may send letters placing their orders with the suppliers in Benaras or they phone them; phone calls are more common nowadays. They do not need to see the samples because each fabric has its own name and particular design—the only thing they have to specify is the colour and the quantity. Over the years, the changes in the designs and colour palette of gos chen fabric have been minimal, and when weavers in Benaras attempt to make changes, then they usually find that the fabric does not sell. Buddhist Ladakhis, and particularly the clergy, want the same designs 14

Abdul Wahid Radhu (1997:176) accompanied his uncle, who was the leader of the Lopchak mission in 1942, and recalls how the two of them realized at that time that the end of the traditional caravan traffic between Ladakh and Tibet had begun.

from benaras to leh

Figure 14.8

343

Aminuddin Ansari, a trader from Benaras, shows samples of his silk- brocade to Lobsang Dorje. Lobsang has a cloth shop in Leh and sells a variety of fabrics that includes silk-brocades from Benaras, velvet from China and cotton cloth from Bhutan. Photo: Monisha Ahmed.

344

ahmed

and colours to be repeated. They state that this is their religious fabric and so it cannot be altered. Thus the patterns that are made are all generally copies of old ones and these can range from the eight auspicious Buddhist symbols to stylized depictions of clouds and flames, mountains, scrolling flowers and dragons amongst others. The fabrics’ affinity to the Buddhist faith is said to be auspicious for the wearer or the surroundings, depending on where the rgya ser is placed or hung.

The Process in Benaras

While weaving is done throughout Benaras, Pili Kothi in Alaipura is the area where gos chen is made. Almost everyone who lives in this area is involved in weaving, or the processes leading up to it, and the trade in some way or the other (Figure 14.9). Signboards, written in English and Tibetan script, guide customers to the shops in Pili Kothi where the main dealers of gos chen have their outlets. They also have looms in small rooms around the area, but the vast majority of their looms are in villages around Benaras. It is largely the more costly brocades, especially those that are made using real gold thread, that are woven in Pili Kothi. This is so that the fabric is woven under the dealers’ direct supervision to avoid pilfering of the costly metallic threads. Also the brocade pieces that have more complicated patterns and designs, such as the Mahakala image shown in Figure 1, are made in Pili Kothi. The weaving is done on a pitloom with a Jacquard mechanism, using punched cards for the pattern.15 Men are engaged in the weaving as well as the preparatory stages such as drawing the graph of the design, making the punch cards, bleaching and dyeing the silk threads. Women are involved in the preparation of the yarns for weaving. They twist the spun yarns to make a strengthened multi-ply yarn suitable for weaving, and then reel the silk threads on to the shuttles for the weaver. The thickness of the fabric requires that gos chen be woven in a narrow width and short lengths, generally not measuring more than 65cm by 500cm. 15

Initially the naksha drawloom was used; its complex mechanisms enabled weavers to create sinuous floral patterns in brocade (Kumar 1999:313). First the pattern was sketched on to mica and then transferred by a nakshaband (pattern maker) onto a thread module (the naksha). The naksha was hung above the loom and attached to the warp threads. By lifting the attached threads, the corresponding pattern was created by weaving the patterning weft threads into the warp.

from benaras to leh

Figure 14.9

345

Silk processing activities and preparation of the warp in Pili Kothi, Benaras. Photo: Monisha Ahmed.

Lengths are measured using the kha, a Ladakhi and/or Tibetan term that refers to a square the size of which is determined by the width of the fabric woven on the loom—this can vary from 60cm to 70cm. One strip of gya ser is generally not more than six or eight khas in length. The term kha is also used by weavers when they lay the warp on the loom, as well as all traders of the fabric when they measure cloth at the time of a sale. A large amount of the brocade that is made in Benaras today is not made using real silk fibres and gold thread. Most of it is made with artificial silk or synthetic fibres such as polyester, nylon, and viscose. This keeps the price of the fabric low, making it more affordable for a larger number of people. The pure fabric is mainly made for the clergy, and generally on the receipt of a confirmed order. Otherwise the cost is too exorbitant for both the weaver and supplier to bear.

Continued Importance of Silk-brocade in Ladakh

The trade in brocade continues to flourish in Ladakh, and today almost all the brocade in the market in Ladakh comes from Benaras. By contrast, the trade in kinkhwab has fallen as it is no longer widely used in Ladakh. The use of gos

346

ahmed

chen is also no longer confined to the clergy and affluent Ladakhis; though they are the only ones who can afford to buy the pure silk fabric. Most of the population is able to afford synthetic mixes, which are very popular nowadays and extensively used. In fact, no bride’s wedding attire is complete if she does not have a brocade cape. Cheap imitations made from synthetic fabrics come from China and are also widely available, but Ladakhis prefer the Benaras brocade. They say that, compared to the brocade that comes from Benaras, the Chinese variety is inferior quality and the fabric is thinner. In Benaras the making of gos chen is a very viable commercial activity. Most of the brocade that is made there is supplied not only to Buddhist communities in India and to the Himalayan Buddhist countries neighbouring India, but also exported to all those regions that have followers of the Buddhist faith, such as Southeast Asia, America and Europe. The only time production fell in Benaras was in the early 1960s, when the Chinese occupied Tibet. As a result of the uncertainty in the market at that time, a lot of the main weaving centres switched over from weaving brocade to weaving saris. As Tibetans were their main customers, many of them thought it was the end of rgya ser; they changed their looms, and destroyed their graphs and punch cards.16 Though the Kasim family also turned to weaving saris, they held on to their looms and patterns, and continued weaving gos chen, though on a smaller scale. Once Tibetan refugees settled in India and other parts of the world, new monasteries had to be built and so the demand rose tremendously. Naturally, the Kasims had the advantage, because they had not stopped weaving brocade. As a result of the increased demand for gos chen, many of the weavers who had stopped weaving this fabric in the 1960s have resumed it, as they realized its profitability. Probably as a tribute to the past, members of the Kasim family frame and hang the original samples of the brocade that were first brought to Benaras in their offices and shops (Figure 14.10). These are reverentially displayed alongside embroidered pieces of ‘Allah’ and pictures of them presenting brocade to His Holiness the Dalai Lama. They also acknowledge, with pride, the Dalai Lama’s words when he introduced them at a gathering by saying: “These are the people who make the fabric of our religion.”

16

Saris are woven on wider looms than those used for gos chen.

from benaras to leh

347

Figure 14.10 A piece of brocade is given prominence on a wall in Haseen Ahmed Kasim’s office, flanked by embroidered hangings of ‘Allah’ and pictures of the Dalai Lama. Photo: Monisha Ahmed.

Acknowledgements In the preparation of this paper and with help during fieldwork, I would like to thank Aijaz Ahmed and Lobsang Dorje in Leh, Sribhas Supakhar and members of the Kasim family in Benaras, and Rahul Jain and Janet Rizvi in New Delhi.

Conservation of Leh Old Town—Concepts and Challenges André Alexander and Andreas Catanese The Tibetan cultural regions across the Himalayas are famous for their large monastic settlements, but they have produced comparatively few cities. In Tibet itself, most of these have changed beyond recognition as a result of the very rapid advent of modernization in the region since the 1950s. Political events, such as the Chinese occupation of Tibet have further diminished Tibet’s urban heritage. By contrast, Leh—the former royal and current administrative capital of Ladakh—still preserves important monuments from the 15th, 16th and 17th centuries. The city is dominated by the former royal palace, a nine-storey stone structure erected in the early 17th century in the Himalayan style made famous by the slightly later Potala Palace in Lhasa. The Old Town, consisting of some 200 stone, mud and timber houses sandwiched between thick rammed earth walls, is located on the slope below the palace, still accessed by a number of ancient stupa gateways. The city is therefore a very important example of historic Tibeto-Himalayan urban architecture. Lhasa and Leh not only share a common architectural heritage: the historic districts of both cities have also faced many of the same problems in the 20th and 21st centuries. Both have had infrastructure deficits, such as a lack of piped water and drainage facilities, and both have been deserted by many of the original owners. In Lhasa this process has been enforced by the authorities for urban policy reasons, while in Leh the wealthier families simply moved of their own accord to the leafy green suburbs of the town. The Tibet Heritage Fund (THF), an international NGO which was founded in 1996, has experience of working in both Lhasa and Leh. Between 1996 and 2000 it conducted the Lhasa Old Town Conservation Project, which developed an approach of using local craftsmen to restore the area, with the full support of the local community. In 2003 a THF team came to Leh and, after carrying out an initial survey of the Old Town, decided to launch a similar project in Ladakh. This paper shares their experience. It begins with a summary of the history of Leh, and then discusses the challenges that the THF has faced, the solutions that it has found and the implications for the future conservation of Ladakh’s built heritage.

© koninklijke brill nv, leiden, ���4 | doi ��.��63/9789004271807_017

conservation of leh old town

349

In 1995 one of the doyens of Ladakh studies, the late Dr. Henry Osmaston (1922–2006), first suggested to André Alexander that he should go to Ladakh and map historic Leh. The occasion was the seventh seminar of the International Association for Tibetan Studies (IATS) held in Austria, where a THF team introduced the project of mapping historic Lhasa. Dr. Osmaston pointed to the need to record old Leh’s historic buildings as soon as possible, as Leh was changing fast and old buildings were disappearing all too quickly. This paper is dedicated to his memory.

Historic Leh

For centuries the kingdom of Ladakh was an important crossroads of the Central Asian caravan trade, sharing many cultural traits as well as language and religion with Tibet. Other Central Asian cultures have also left strong traces. Tibetans reached Ladakh as early as the second half of the seventh century and left behind inscriptions, for example near the Alchi bridge.1 In the early 10th century Skyilde Nyimagon (sKyid lde nyi ma mgon), a descendant of the last king of Tibet’s Yarlung dynasty who was assassinated around 842 CE, founded a Western Tibetan kingdom.2 That was, more or less, the start of the Ladakhi dynasty which ruled until 1834. The earliest known royal domain of this dynasty was in Shey­­in the Indus valley, some 15 km upstream from Leh. The early history of Leh itself is shrouded in mystery. The discovery of a Buddhist monument with a Tibetan inscription dated to the period between the eighth and the tenth centuries suggests the existence of a human settlement at Leh at the time, centuries before Leh entered recorded history. This is a two metre-tall image of a standing Maitreya engraved on a rock in the ruins of an old house at the edge of the Old Town, found during the THF’s community survey. It has since been recovered and permanently installed in front of the former Sankar Labrang (gSang mkhar bla brang) building in Manikhang (Maṇi khang) area of old Leh.3 Leh’s strategic location in the centre of the stretch of the Indus River that defines most of

1 See Denwood (1980, 2008, 2009) and Bruneau (2004). See also the unpublished paper by Takeuchi Tsuguhito, ‘Old Tibetan Rock Inscriptions near Alchi’ (kindly made available by Amy Heller and Quentin Devers), as well as the discussion on Ladakh’s early history in Phuntsog Dorjay’s article in this volume. 2 See Petech (1977:14–17). 3 See Alexander & van Schaik (2011). Again, see also Phuntsog Dorjay’s article in this volume.

350

Figure 15.1

alexander and catanese

View of Leh Old Town showing the royal palace as well as the monasteries, temples and private houses of the king’s ministers just below.

Ladakh may have proved attractive. It also has a moderate climate that allows the growing of wheat. The earliest recorded building activity dates to the reign of Drags Bumdé (Grags ’bum lde) in the second half of the 15th century. In 1461 the king received a delegation from Central Tibet sent by Gendündrup (dGe ’dun grub, 1391– 1474), who was later recognised as the First Dalai Lama,4 and subsequently began the building of Buddhist temples. Three Maitreya temples (Byams pa lha khang) are preserved from that time: the ‘Red Maitreya Temple’ (Byams khang dmar po) just below the Leh Palace, the ‘White Maitreya Temple’, also known as ‘Street Maitreya’ near the Manikhang area of Leh’s Old Town, and one more temple next to the Tsemo (rTse mo) tower mentioned below. The next confirmed building activity took place during the reign of Tashi Namgyal (bKra shis rnam rgyal, r.c. 1555–1575). He is credited with the Namgyal Tsemo tower above the Leh place, and, as proved by an inscription, with the protector temple below the tower. The rammed earth walls around Leh 4 Lozang Jamspal (1997:141); Petech (1977:21) mentions that an earlier king had built some stupas at Leh.

conservation of leh old town

351

town may also have been built during his reign.5 However, it is King Jamyang Namgyal (’Jam dbyangs rnam rgyal, r.c. 1595–1616) and his son Senge Namgyal (Seng ge rnam rgyal, r.c. 1616–1642) that are credited with having established Leh as the royal capital. They erected the nine-storey Leh Palace, and their ministers settled immediately below within a fortified area. This is today’s historic Old Town of Leh (Figure 15.1). The area was known as Gosum (sGo gsum), ‘Three Gates’, from the number of gates built there, of which only one is extant, and was originally inhabited by 40 families. Parts of the ancient city walls are still preserved, but most of them were broken down as the city expanded.

Survey of Old Leh

Modern Leh is a city of some 35,000 inhabitants. The population almost triples in size during the summer tourist season, when people from all over India come looking for work. To protect rural Ladakhis from economic competition for which they are not prepared, the government has barred outsiders from acquiring land and the latter require permits—and often a Ladakhi partner— to work or open businesses. Ladakhis therefore seem to benefit from tourism more than their cousins in Lhasa do. The annual arrival of more than 15,000 tourists wanting regular showers and flush toilets and producing mountains of garbage (water bottles, food wrappings, toilet paper etc.) puts a severe strain on Ladakh’s fragile ecology. However, tourism almost completely bypasses the Old Town. That is mainly due to the lack of a regular water supply: five public taps, operating mostly for only thirty minutes each morning, are the only water supply for 200 families in the historic Old Town. The THF’s first task was to conduct a survey of Old Leh, reviewing both the state of its buildings and the social and economic conditions of the people who inhabited them. In 2003 and 2004 a small team consisting of André Alexander, Miss Lharigtso, from Tibet, and Miss Diskit Dolker, the first Ladakhi urban planning graduate, knocked on almost 200 doors in the Old Town. Each time, we were invited for tea and biscuits, and sat down with the inhabitants. We filled out a questionnaire for each building, listing the building’s name (Ladakhi houses are known by a given name rather than an address), its age, general condition, obvious defects, and information about the inhabitants, their number, relation to the building, income, family history etc. We also listened to people’s general opinions and observations about life in the Old Town, and their problems. We recorded 187 historic (i.e. over a century old) or 5 As suggested by Howard (1989).

352

alexander and catanese

traditional buildings (i.e. constructed in the last hundred years with traditional material and designs). The survey findings revealed that, for example, over 55% of the historic building stock was in bad or poor condition, and that the average monthly household income of the residents of Old Leh was little more than US$100. The social data obtained during the survey pointed to the need for intervention to improve people’s livelihood and living conditions. That was matched by a desire, generally expressed by many community members, to reverse the decline of the Old Town. Drainage was seen as major problem: there were only a handful of open channels, and these frequently become blocked or frozen. Heavy rainfalls or a neighbour’s washing day might trigger the flooding of one’s ground floor. Houses in the Old Town generally had (and still have) no running water and, as mentioned above, residents had to rely on a handful of public taps. The toilets in the Old Town were all of the standard Himalayan composting type, needing no water, as flushing is done with a handful of sand. Technically, some aspects of improving conditions in the Old Town presented a significant challenge; especially building a drainage system into the sheer rock in a region that has nearly six months of sub-zero winter temperatures. However, such technical problems, both there and elsewhere in Ladakh, could be solved mostly with locally available technologies and materials. The challenge is to identify and use the best of traditional skills, which have slid into obscurity since the advent of subsidized cement and steel, and blend them with adequate modern technologies where necessary.

Three Models for Heritage Conservation in Ladakh

In contemporary thinking, there are three distinct models of conservation:

‧ ‘Pure’ conservation refers to the discipline of treatment, preventive care, and ‧

research directed toward the long-term safekeeping of cultural and natural heritage. This approach befits major historic monuments.6 Rehabilitation refers here to the activity of returning deteriorated objects, structures, neighbourhoods, or public facilities to good condition. This process may involve repair, renovation, conversion, expansion, remodelling, or reconstruction.

6 These definitions are based on The Art & Architecture Thesaurus by the Getty Research Institute. See: www.getty.edu/research/conducting_research/vocabularies/aat/.

conservation of leh old town

353

‧ Adaptive re-use means to make something suitable for a new use or purpose.

An example is represented by the conversion of outmoded or unused structures, such as buildings of historic value, to new uses or applications in new contexts.

Pure conservation, which seeks to ‘freeze a building in time’, is suitable for major monuments, such as the ancient temples on the Tsemo hill or the Leh Palace, as well as many of Ladakh’s temples and monasteries. However, this approach is out of the question for the majority of historic residential buildings within the former walled city of Leh. Ground floor rooms are dark and poorly ventilated, as they were used as stables and granaries in the nottoo-distant past. These houses can be preserved only if they can be modified to the changes in lifestyle of their owner/occupiers during their rehabilitation. The main emphasis in Old Leh has therefore been on adaptive reuse rather than on pure conservation.

The THF’s Approach

The information gathered from the survey was shared with the Ladakh Autonomous Hill Development Council (the region’s semi-autonomous government, usually known simply as the ‘Hill Council’) and subsequently used to launch the Leh Old Town Conservation Project. The THF proposed an integrated approach on the model of the Lhasa Old City Conservation Program. A key element in its strategy has been to work together with the local administration to create a planning framework (a new Masterplan for Leh) and to improve the infrastructure. The THF—and most local community members—expect government action to be slow in coming, and therefore restoration and upgrading activities were started almost immediately by the THF, and the government was politely informed. However, from early on, all local government departments expressed their support for our activities, and made official maps and surveys available to us. In 2006 the THF and the Hill Council signed a Memorandum of Understanding to work together to preserve historic Leh. In the same period the THF worked with local people to set up the Leh Old Town Initiative (LOTI), which is registered as an NGO under the India Society Act. The first draft of a Masterplan for historic Leh was submitted to the Hill Council, and to the Jammu & Kashmir Ministry of Tourism and Culture. Secondly, in a series of community meetings in the Old Town, the THF designed a model offering 50% co-financing for adequate rehabilitation of

354

alexander and catanese

homes, on the condition that indigenous labour and skills were mainly used. Drawing on the experience of comparable work in Lhasa, the THF also offered improvements such as: introducing bathrooms with drains; bringing more light into the often dark houses (which were built when glass was unknown or unavailable); improving the composting pits of traditional latrines; and increasing the efficiency of traditional clay mixes used on roof surfaces in the light of our experience and the skills of the best traditional craftsmen. Several house owners immediately took up the offer, so that at present there is a waiting list, as the THF’s finances only allow for a limited number of buildings to be upgraded each year. At the outset the THF spent several weeks identifying Ladakhi craftsmen to carry out the work. The THF team travelled to surrounding villages and interviewed many craftsmen, and finally hired a small group consisting of two masons, Jamyang Tarchin and Sonam Dorje, and two carpenters, Tsering Dorje and Tsering Puntsok. This became the first project in Leh and vicinity to work again with Ladakhi craftsmen. Several years later, all except one are still working for the Old Town project. For training and labour, preferential hiring is given to poor residents of Leh. The objective is to give employment to those inhabitants of the Old Town who have no land and little education, thereby giving them an economic opportunity. The THF next worked with this core group of artisans to restore a communally owned shrine, the 17thcentury Guru Lhakhang in the Old Town, to demonstrate the practicalities and desirability of restoring Leh’s historic building stock. After conducting several community meetings, a model rehabilitation area was next chosen, the Stagopilok residential lane, where housing and infrastructure were to be upgraded. After this pilot project, a community meeting was held in the garden of the Tak Guesthouse in the Old Town, and the Old Town community was presented with an offer: 50% co-financing for the rehabilitation of Old Town houses carried out under the aegis of LOTI and its craftsmen. Administrative costs or fees for architects (especially foreign ones) would not be charged. Immediately one resident, Haji Abdul Qadir, the owner of Sofi House, asked for his home to be restored accordingly. Since then, around 15 historic residential buildings, temples and mosques in the Old Town have been restored.

Conservation of Historic Architectural Structures

Ladakhi temples and residential buildings share the same basic architectural structure. An internal timber frame on stone foundations supports flat,

conservation of leh old town

Figure 15.2

355

A thick layer of ‘markalag’ being laid on the roof by mason Tsering Dorjey.

mud-covered roofs. Walls are built from stone rubble and sun-dried mud bricks. Layers of local clays and mud are traditionally used to create waterproof roofs and dust-free plastered interior surfaces. Many Ladakhis say that rainfall in the region has substantially increased in recent years, perhaps as a result of global warming. Be that as it may, most house owners complain of leaking roofs. Imported corrugated iron sheeting is readily available in the local market and has become very popular as roofing material in both old and new Leh. However, its aesthetic suitability and sust­ ainability may be doubted. In the high altitude deserts of Ladakh, where few building materials are available, the skills and knowledge of local craftsmen created a highly sophisticated way of making the best use of whatever was available. The THF approach therefore stipulates that the traditional materials are reused and adapted where necessary to improve their quality. For example, water-proofing of the flat mud roofs is traditionally done with ‘markalag’ (mar ka lag). ‘Markalag’, meaning ‘buttery mud’, is a kind of clay which is available almost cost-free in the areas around Leh. In the past a thin layer on a mud roof was sufficient to ensure dry living rooms. Today its impermeability must be improved and therefore the THF has added a five

356

Figure 15.3

alexander and catanese

Composition of ‘markalag’ samples as analysed by the University of Stuttgart.

centimetre-thick layer of ‘markalag’ between layers of mud (Figure 15.2). When it rains, the clay will absorb the initial humidity and then swell, creating a solid layer that is impermeable. After several years and after many heavy rainfalls, the roofs treated in this way are still waterproof. During the THF’s first year of activity in Ladakh samples of different soil qualities used in construction were collected. These included: the locallyfamous soil of Shey, which is used to make the best quality mud bricks; the yellow soil (ser sa) of Stakmo (near Thikse); the ‘stove soil’ (thab sa) used to build traditional stoves, but also for flooring; and ‘markalag’.7 Those samples were analyzed thanks to the kind help of Prof. Achim Bräuning of the University of Stuttgart, who compared the results with samples from Tibet (Figure 15.3). In Tibet a fine grained limestone (micrite) is used to waterproof roofs, and this is known in Tibetan as ‘arga’ (ar ka). ‘Arga’ is rare in Ladakh and the THF found only one sample at Mangyu. The analysis showed that ‘markalag’ is indeed 80% pure clay with some silt but no sand at all, while all the other soil samples, including the ‘arga’ from Tibet, contain sand and silt as well as a little bit of clay. The quality of the Shey soil appears to be related to the high content of calcium carbonate found within the silt, but further testing is necessary to confirm this. The order of layers in an improved Ladakhi roof starts with a ceiling of wooden beams, rafters and willow-stick joists (on which the THF places woven straw mats to prevent dust from falling through the joists) (Figure 15.4). Next comes a layer of Ladakhi ‘yagtses’ grass, a traditional stop-gap layer: if water makes it thus far, the grass can absorb water several times its volume. The grass 7 See also Hubert Feiglstorfer’s article in this volume.

conservation of leh old town

Figure 15.4

357

A section through a Ladakhi roof shows the main beams (‘madung’), the secondary beams (‘dungma’) and the ceiling joists (‘talloo’).

also serves to insulate the roof. Next comes a layer of rough soil, and then the layer of ‘markalag’ clay. The final layer of soil on the roof is applied wet, and its mix may include straw and even the dung of cows, donkeys or horses to increase its solidity, as people often walk on the flat roofs, which are used for gathering and for performing certain household chores. Parapets are being improved by capping them with finely-cut slate stone, a method commonly used in Tibet and elsewhere, and introduced to Leh by the THF. When necessary, modern materials may be used, even if their utilization is concentrated only in the most vulnerable spot of the roofs where, due to changing climate, traditional materials fall short of perfection. Where possible, modern materials are hidden by layers of plaster or mud so that the original design is not altered. For example, locally-available bitumen (tar paper) can be used to protect the inner edges of parapets against possible leaks. The traditional water spouts can be improved by fitting them with stainless tin pans in the crucial area between the wooden spout and the mud of the roof. It is helpful to use cement to paste the wooden spout solidly onto the roof, in the form of a cement pan in which the spout lies.

358

Figure 15.5

alexander and catanese

The interior of the restored Masjid Sharif. The original Mihrab was restored although it almost collapsed during the restoration works.

Improvements in the interior consist of adding windows or enlarging existing ones; as well as water-proofing surfaces which might come into contact with water (kitchen, bathrooms); and generally creating more durable and dust-free surfaces.

conservation of leh old town

359

One example is the Hor Yarkandi house, a residential building erected a hundred years ago by a trader from Yarkand, in the Chinese province of Xinjiang, to the north of the Karakoram pass. Here the THF added apricot juice to the floor mix in some rooms and cow dung in others. As a result, the floors are more durable and dust-free. One room was designed as a bathroom and cemented, and left with holes for piping to fit a tank on the roof for showers, and fitted with piping connected to the street drain. The composting vault of the toilet was concreted in the inside, to prevent urine seeping into the foundations, a very common problem in Tibet and Ladakh. The THF has also worked on Islamic buildings. In 2007 the 400-year old Masjid Sharif (Figure 15.5), Leh’s oldest Sunni mosque, was painstakingly restored, with great input from the Anjuman Moin ul-Islam Society. This mosque was founded by caravan traders in the Tsa Soma (Tshas so ma) gardens, an old camping ground for caravans from Central Asia. It is one of the last mosques to keep its original appearance, an Islamic building in TibetoLadakhi architectural style. Some 60 years after Ladakh’s historic caravan trade came to a complete halt due to unresolved border disputes with Pakistan and China, the Jammu & Kashmir Ministry of Tourism commissioned the THF to design and build the Central Asian Museum at Tsa Soma to commemorate the trade. The museum was opened in July 2011. The old Sankar Labrang house in the Manikhang area is an example of the adaptive reutilization of a historic building. Erected possibly several hundred years ago around a group of five stone-carved Buddhas slightly smaller than life-size, the upper floor served as residence of the caretaker monk of the White Maitreya Temple nearby.8 The mother monastery, Sankar Gonpa near Leh, abandoned the house two decades ago, and asked the Goba (Go pa) family to service the temple instead. The monastery was planning to eventually demolish the house and to construct modern shops there, but the THF persuaded the monastery to keep the house, and offered to restore it in return for a nominal and limited lease. After conversion, it became the Leh Heritage House (Figure 15.6), a gallery showing modern art and photography, and an exhibition about historic Leh. Maps of the Old Town, and information about the conservation project and about old Leh (as well as Italian espresso) are also available. The conversion plans included the gutting of the upper floor, which was split into several small rooms and a large exhibition room was created. The flooring and roofing was designed, the former to bear the trampling of many visitors, so slate stone was laid on the floors and roof. The paving has so far withstood 8 See Phuntsog Dorjay’s paper in this volume.

360

Figure 15.6

alexander and catanese

Sankar Labrang, now named Lala’s café, in Leh’s Manikhang area.

conservation of leh old town

361

several dancing parties with Ladakhi, Bollywood and Western music, some of it played live.

Conservation of Wall Paintings

Wall paintings are an important component of the Tibetan Buddhist heritage. The interior walls of temple halls, monastic assembly halls and shrines would be painted in their entirety. Traditionally, mineral pigments with animal glue as binder would be applied on a preparation of chalk on dry mud plaster. In Ladakh several temples—notably at Alchi—still preserve early murals executed in the tradition of the now-lost Buddhist civilization of Kashmir, while later murals are done in the traditional styles of West, Southwest and Central Tibet. Compared with Tibet, where many historic monasteries and their paintings were damaged during the Cultural Revolution, the small territory of Ladakh preserves an astounding number of early paintings. However, there were no professional local mural restorers prior to the THF project. Soon after setting up its office in the old Lakruk House just below the Leh Palace in 2005, the THF received requests for assistance from numerous monastic establishments, and saw numerous temples whose ancient murals were damaged by leaking roofs, structural faults or badly-executed restoration. Realizing the need for local experts in mural conservation, the THF set up its conservation training programme in the spring of the same year. Initially two young Ladakhis, Yangchen Dolma and Skarma Lotos, were trained in situ by international restorers. This programme has now become institutionalized thanks to the cooperation between the THF and the Conservation Department of Erfurt University of Applied Sciences in Germany. After two years of training in the field, the two Ladakhi students travelled to Erfurt to receive further training in the laboratories of the University. They also participated in a conservation project in a local museum. Back in Leh, the caretaker of the Red Maitreya Temple, Ven. Ngawang Tsering, requested the THF to carry out routine roof repairs to prevent rain leaks from soiling the 15th century three-storey image of Maitreya, the Future Buddha. During the work, the present author discovered two walls with original wall paintings (Figure 15.7) hidden behind a coat of whitewash applied in the late 1950s. The Red Maitreya Temple is said to have been founded by the Ladakhi king Dragspa Bumdé (Grags pa ’bum lde), generally known as Drags Bumdé, who has been mentioned above.

362

alexander and catanese

Figure 15.7

Anca Nicolaescu cleaning the mural paintings inside the Red Maitreya Temple.

The Romanian restorer Anca Nicolaescu, the Ladakhi team of trainees and German conservation students from Erfurt spent three years bringing back these paintings from underneath the coat of white-wash, and have conserved and partly retouched them. The style confirms that the paintings are indeed from the founding period in the 15th century, making them the oldest wall paintings in Leh. The Ladakhi trainees, supported by two more young girls, Sonam Dolma and Tsering Chorol, have since completed another three-year project, cleaning and consolidating late 13th century murals in the Tsatsapuri group of temples in Alchi village. They are currently working on restoring the Gangtok Tsuklakhang, the former royal palace monastery in Sikkim.

Looking Towards the Future

The THF Leh Old Town project can look back on the successful rehabilitation of several clusters of buildings and monuments in old Leh. That work received a 2006 UNESCO Asia-Pacific Heritage Award and a Dubai Best Practice Award from UN Habitat. Funding has been provided so far by the governments of Germany and Finland, but it is hoped that also the Central Government of

conservation of leh old town

363

India or the Ladakh government may find funds for Leh Old Town. A start was made in 2008, when the Ministry of Tourism and Culture of Jammu & Kashmir offered to fund the Central Asian Museum which has now been built by the THF in the Old Town of Leh. In the future the THF hopes to hand over its project to LOTI, the NGO consisting of local experts and community representatives. However, much remains to be done. A new motorable thoroughfare through the Old Town, designed to ease traffic jams in the Main Bazaar, is threatening to transform the lower part of historic Leh and turn it into an extension of the Main Bazaar. The local government has declared Old Leh a Heritage Zone, but so far has not said what that would mean. Nevertheless, several areas of Old Leh have been visibly improved through the project and a steady stream of applicants from all over Ladakh are requesting their historic village homes or temples to be restored or upgraded on the model of historic Leh.9 In recent years much has been done to conserve Ladakh’s architectural heritage, but its future is far from assured.

9 Further technical reports on some of these projects can be found on THF’s website, www. tibetheritagefund.org.

Revealing Τraditions in Εarthen Αrchitecture: Analysis of Εarthen Βuilding Μaterial and Τraditional Constructions in the Western Himalayas Hubert Feiglstorfer It is amazing to see earthen constructions from more than a millennium ago still standing and even being used as sacred buildings. Continuous maintenance is essential to keep these buildings alive and prevent their decay. Their long-term care requires shovelling snow from the flat roofs after every snowfall, and the yearly tuck-pointing of cracks in the exterior plaster and on the roofs, as well as the occasional replacement of the earthen layer on the flat roofs—if necessary even the replacement of individual wooden beams or of the entire roof construction. On the other hand, neglect can easily result in rapid weathering, and finally in the collapse of the buildings. The construction and maintenance of these buildings require an understanding of the properties of the different kinds of earth that are used as building materials, and which in turn help determine the technical standard of each construction. This knowledge is largely based on oral traditions which have rarely been assessed in a scientific context. The main purpose of the present paper is to fill this gap. The loss of traditional knowledge inevitably leads to a decline in the quality of new earthen buildings as well as the durability of older ones. In addition, heavier seasonal rainfall due to climate change causes heavy moisture penetration which can easily result in the saturation of the ceilings and walls. The seepage of water into the construction causes long-term deterioration and, in the case of sacred buildings, this may include damage to wall paintings. In this contribution, I report the results of the Austrian Academy of Sciences (ÖAW) expedition to Western Tibet (Ngari) and beyond in 2010, as well as research conducted in Ladakh in August and September 2011.

Geological Parameters

For a detailed examination of earth material, a wide range of samples was taken from all over the Western Himalayan region. The following observations

© koninklijke brill nv, leiden, ���4 | doi ��.��63/9789004271807_018

revealing traditions in earthen architecture

365

are based on laboratory analysis of a selection of different types of soil for comparative purposes. In general, the material analysed can be described as geologically young. Judging from the samples, the minerals were transported only for relatively short distances from their sources to the construction sites. The sample 60571 of the unplastered external wall of the Kardong monastery at Kyelong in Lahul and the samples from the Nubra valley are made up of loess-like sediments. The high amount of silt in several samples may be a result of wind transport. The analysis of the sample 6052 of the ‘markalak’ (mar ka lag) clay from the clay pit in Spituk suggests that soil from the Indus valley consists of chalk or glacial flour (so-called ‘glacial milk’), a fine-grained material, derived from ground bedrock by glacial erosion. The higher amount of chalk corresponds to a higher pH value and, accordingly, to a higher amount of carbonate. The isometric rounding of the mineral grains (they are roughly cube-shaped but with rounded edges) is likely due to river transport. This rounding may be found primarily in the samples from the Nubra valley and Upper Rupshu in the northeast of Ladakh where the geological material is transported by rivers and glaciers. The microscopically small fibres found among the mineral grains may be interpreted as foraminifera, which are indicators for marine sediments. They may be found in samples from the Spiti, Pin, Lahul and Nubra valleys. The grain size distribution points to supra-regional differences, but also to some regional similarities. There are loam samples with a poor sorting, for which the cumulative curve of the grain size shows no peaks, as well as loam samples with a good sorting and distinct peaks. A database of the features of all the investigated samples from all over the Western Himalayan region will facilitate our understanding of this issue. The systematic listing of geological features as the basis for the allocation of particular types of earth in certain areas will enhance our understanding of their development over this region. The following analysis draws on the findings of our laboratory research to discuss the geological parameters that contribute to the qualities of different kinds of earth as building materials.

1 The sample numbers refer to their listing in the laboratory notebook kept at the Institute of Applied Geology at the University of Natural Resources and Life Sciences in Vienna. They enable the quick location of the samples for further research. The material tests in the laboratories of the university were carried out under the supervision and guidance of Prof. Franz Ottner.

366

feiglstorfer

Indus Valley in Ladakh, in the Region of Leh The bedrock in this area is comprised of Indus molasse and marine sediments, which are most probably carbonates.2 In the Indus valley the soil is rich in fine material such as silt which cannot be found in the same proportions in samples from other areas. The Indus valley, including Rupshu, shows a thrust fault. The bedrock of the southern part of the valley contains turbidites (flysch sediments), while the bedrock of the northern part of the valley is comprised of molasse. The samples from Upper Rupshu in general show a homogenous distribution of grain size. However, the cumulative curve of the particle size distribution shows a sudden bend from fine sand to the coarse silt fraction. Accordingly, these samples represent pure and unblended loam material. Otherwise these fractions would be evenly distributed. The sand content in these samples is relatively high, with values between 12.4% and 14.4%, whereas the amount of clay (8.3% and 9%) and silt (14.4% and 12.4%) is relatively low. In that respect, all samples show similar characteristics. In the Indus valley loams with high contents of fine material such as silt occur and in that respect the samples distinguish themselves from samples from other areas. However, sample 6045 from Alchi in the Indus valley is coarser. No gravel could be found in samples 6073 from Nyarma or 6052 from Spituk, and the proportion of silt ranges between 56.9% and 64.6%. The amount of clay in the Nyarma sample is one of the highest in the samples examined. The relatively high proportion of sand in the two samples from Rupshu ranges between 12.4% and 14.4%. The silt and clay content are relatively low: the former between 8.3% and 9.0%, and the latter between 12.4% and 14.4%. Nubra Valley, the Region North of the Indus valley The bedrock of the northern area is comprised of molasse. Further north, granites and other granite-like rocks can be found. The turbidites (flysch sediments) in the Nubra valley are dominated by marl. The loam samples from the Nubra valley show a similar grain size distribution to the samples of the Indus valley, with high proportions of fine sand compared to fine silt. The proportion of silt ranges between 23.5% and 53%, and the amount of clay between 4.8% and 29.8%.

2 Fuchs (1980) and Fuchs (1984).

revealing traditions in earthen architecture

367

A small quantity of tiny pieces of coal in samples 6047 from Sumur and 6043 from Hunder points to an anthropogenic component which may have been added either deliberately or accidentally. Spiti Valley The bedrock in this valley is comprised of middle to strong metamorphic rocks, such as mica, schist and gneiss, as well as limestone.3 The bedrock in the Pin valley consists of Quaternary sediments with river sediments and slightly weathered material on top. Along the flanks of the valley, debris and moraines from the tributary valleys can be found. In both the Spiti and the Pin valleys, limestone is covered with alluvial deposits. Carbonate has been deposited from the tributary valleys.4 The loam samples from the Spiti and Pin valleys are relatively coarse, with a gravel content between 18.4% and 27.5%, the amount of clay ranging between 13.4% and 19.2%. The interior plaster in the temples at Tabo, with gravel contents of 1.1% and clay contents of 39.8%, differs from the other samples. Lahul, Kyelong and Surrounding Areas Granite and schist can be found in the area around Kyelong, resulting in high amounts of quartz, mica and feldspar in the samples. In Lahul we may find earthen building material with little coarse material and a good sorting of the sediment. The grain size distribution shows a slightly elevated content of the fine sand compared with the fine silt fraction. However, the grain size distribution in general is similar to that of the Spiti and Pin valleys, the gravel and clay contents being respectively 0.5% and 13.5%. In general, in all the areas examined the particular types of loam used for walls and for roofs were rather uniform. Even the mineralogical composition used for adobe bricks and for rammed earth did not differ much in several cases, probably because the material was taken from the same clay pit. One exception is the interior plaster used in Tabo which shows a relatively high amount of fine material. The cumulative curve of the grain size distribution suggests that two different types of loam were taken from clay pits which were situated close to each other. Interestingly, when comparing the interior plasters of Tabo and Nyarma, we can identify a great similarity in their mineralogical composition. In this case such mixtures probably reflect a certain tradition in working with earthen 3 Fuchs 1978. 4 Fuchs 1979.

368

feiglstorfer

material. Furthermore, it is worth mentioning that the loam used for the upper layer of roofs is not necessarily very fine, as quite coarse material could also be found, as shown by sample 6043 from the Lhakhang Marpo of Hunder. For the conservation of wall paintings as well as for new paintings, the following discovery is certainly essential: sample 6044, an interior plaster from Hunder, as well as sample 6073, an interior plaster from Nyarma, contain swellable clay minerals, such as smectite or vermiculite or even both, which expand when humid. Probably this swellability helps improve waterproofing. Bulk and Clay Mineral Analysis The results of the bulk mineral analysis show a relatively high quantity of sheet silicates in the samples. For example, in the valleys of Spiti and Pin, the samples show contents between 40% and 62%, in the Indus valley between 20% and 63% and in the Nubra valley between 22% and 31%. In Rupshu the contents are much lower, between 22% and 31%. In the Spiti, Pin and Indus valleys, as well as in Lahul, the amount of plagioclase, calcite and feldspar is relatively low. By contrast, in the Nubra valley higher amounts of quartz between 28% and 31% and of plagioclase between 31% and 36% may be found. Samples 6044, 6046 and 6047 from the Nubra valley show dolomite contents between 2% and 4%. In the other samples no dolomite could be found. In general, the clay mineral analysis shows a high amount of illite, between 70% and 82% in the valleys of Spiti and Pin, 76% in Lahul, 67% in the Indus valley, 61% to 81% in the Nubra valley and 78% to 79% in Rupshu. Swellable clay minerals could be found in only a few samples. In the samples of the valleys Spiti and Pin no swellable clay minerals were detected. Sample 6057 from Lahul contains 1% vermiculite, while samples 6073 from Nyarma and 6052 from Spituk show contents of swellable smectite of 3% and 7% respectively. Samples 6046 from Diskit and 6044 from Hunder have smectite contents of 1% and 2% respectively; sample 6046 from Diskit contains 5% of vermiculite; and samples from Rupshu contain 2% of smectite and vermiculite. The comparison of the different types of loam which contain swellable clay minerals shows that they are rich in silt. Sample 6057 from Lahul contains 69% silt, of which 32% is coarse silt; sample 6073 from Nyarma contains 56.9% silt, of which 41.8% is fine silt; sample 6052 from Spituk contains 84.6% silt, of which 39.5% is middle silt; sample 6046 from Diskit contains 45.7% silt; and sample 6044 from Hunder contains 46.6% silt.

revealing traditions in earthen architecture



369

Different Uses of Earth for Building Purposes in the Western Himalayas

Earth and soil in general are called sa in Tibetan, regardless of whether this refers to building material or any other kind of earth. Sand (Tib. bye ma, pronounced ‘jemà’) is added to thin sa which contains too much clay, together with water (Tib. chu) if this is necessary and available. Nowadays sieving is applied to separate particular grain sizes from one another. Generally, certain kinds of soil in the vicinity of particular villages are known to make good building material. For example, these include the soils of Shey, Matho, Spituk, Basgo, Likir or Lamayuru, as is commonly known all over Ladakh. However, the suitability of particular kinds of earth depends on the part of the building under construction. Each part, such as the roof, the plaster, the floor, the brick or the rammed wall, requires a specific kind of earth—or mixture of different types—as basic material. An understanding of the properties of the different kinds of soil and their suitability or otherwise for specific purposes, makes it easier to decide how to make the best use of local materials, thus also minimising transport costs. On the one hand, mention may be made of soil which is used for rammed earth (sa brgyang pa, pron. ‘sa gjangpa’), for which a coarse material with a lower content of clay is used. On the other hand, the soil that is used to shape adobe bricks (pa’u, pron. ‘pau’) contains a higher proportion of clay than the one used for rammed earth, but lower than that found in the earth used for plaster. The finest earth, if we exclude the one used to construct sculptures, is used for the interior plaster, usually with a higher content of clay. To avoid cracks in plaster containing a high amount of clay, it is mixed either with organic material such as straw, pine needles, certain fibres or animal hair, or with a more sandy earth. This process is known as thinning the clay. A large variety of different kinds of earth were collected during a series of field research trips, several of them being locally known under a specific term. For example, Likir is known as a potters’ village: this is primarily because of a specific kind of red soil that can be found in its area, similar to Basgo, which is erected on a hill of red soil. The clay used for pottery is known as rdza sa (pron. ‘dzasà’) (Figure 16.1) and the profession of the (clay) potter as rdza mkhan (pron. ‘dzakhèn’):5 this craft is rare in these Himalayan regions because of the 5 Hamid (1998:224).

370

Figure 16.1

feiglstorfer

Likir. Clay pit with rdza sa, the pottery soil. Photo: Hubert Feiglstorfer.

shortage of firewood. As in Basgo and Nye, thab sa (pron. ‘thapsa’, thab meaning ‘stove’) and rdza sa in Likir refers to two different kinds of soil. Thab sa (Figure 16.2), in its shiny greyish-black colour, is used primarily to build the traditional Tibetan stove, whereas rdza sa is used for pottery. For that purpose, it is mixed with fine sand with a regular distribution of the sizes of the grains. It is also used as plaster for the interiors of buildings. Thab sa can also be used to make a floor or a roof by ramming. For these purposes it is mixed with sand. On the other hand rdza sa, a reddish earth, is not suitable for this purpose on its own as it would crack, but it can be mixed with thab sa for further processing. A bottle or, in earlier days, a round stone is used to give it a shiny surface. Today these materials are no longer used to make roofs or floors. To avoid cracks, thab sa for the production of bricks has to be mixed with river sand or with the fine pottery sand which can also be dug in a pit close to the thab sa and rdza sa pits in Likir. The close location of all these different varieties of clay, including the pottery sand and lime, at a distance of a few dozen metres, is a special feature of the Likir clay pits. Because of its heat-resistant quality, thab sa from Likir has been transported up to 60 km away to Phyang, as the Indian army needs this material to coat the interior of metal oil drums to make Indian tandoori ovens. Thab sa is ground, sieved and mixed with sand in a dry constitution with a proportion of 1:2.

revealing traditions in earthen architecture

Figure 16.2

371

Likir. Clay pit with thab sa, the soil used for making traditional Tibetan stoves. Photo: Hubert Feiglstorfer.

372

Figure 16.3

feiglstorfer

Phugtal. Clay pit with rdza sa, locally known as ‘arga’. Photo: Hubert Feiglstorfer.

To this mixture water is added and after about half an hour the thab sa is ready. The proportion of rdza sa and sand for pottery is 4:2, the unit being measured with a full hand.6 The thab sa of Tunlung, by contrast, is not suitable for making floors as it is not water-resistant and would turn into a clayey mass. In Nye, the local soil is used for this purpose, and in Basgo the thab sa is mixed with sand but, as with the above-mentioned example in Tunlung, it is not waterresistant.7 Deposits of rdza sa in the villages of Shara and Shey were also 6 The interview was held together with Sonam Wangchuk and Rebecca Norman and the family of potters of Likir in August 2011 in the fields of Likir. Sonam Wangchuk and his wife Rebecca Norman run the SECMOL (Students’ Educational and Cultural Movement of Ladakh) campus at Phey, in Ladakh, experimenting with traditional building materials such as earth. The family of potters living in Likir, whom we met on their way home from their fields during harvest time, offered us their provisions from the field and lots of information on traditional pottery making. The house name of this family is Langdopa. Rigzen Wangyal, the young potter, was accompanied by his brothers Rigzen Angdu and Tsering Norbu as well as by his mother. 7 This interview was conducted together with Sonam Wangchuk, who translated into English, in Nye in August 2011 at Tsewang Norbu’s home. Sonam Wangchuk is well-known as the founder of SECMOL in 1994 and as having worked continuously with earth as a building material. Tsewang Norbu is a native of Nye. His profession is primarily that of a blacksmith. His

revealing traditions in earthen architecture

Figure 16.4

373

Looking over the red earth (’phred sa) mountain of Basgo. This colour determines the appearance of the local landscape as well of the building structures. Photo: Hubert Feiglstorfer.

reported to me. Close to the entrance of Phugthal monastery, a lama who had been at a monastery in Tibet years ago showed me a yellow-reddish soil which appeared to me to be similar to the rdza sa (Figure 16.3) that I found in Tunlung, and he described it to me as ‘arga’. The area around Basgo is marked with a roadside board as ‘Sasa’ and another board at the junction of the Leh-Kargil road with the road to Nye mentions ‘Zasna’, which may refer to the reddish soil and hill in combination with the syllable sna (‘nose’, ‘tip’, ‘spur’).8 The earth of the Basgo hill is called ’phred sa (pron. ‘thetsa’), ’phred probably referring to a slope (Figure 16.4).9 Sonam Wangchuk mentions that the soil of the Basgo mountain reaches a depth of about 180 metres below the surface, information which he received from his father was already producing stoves (thab, pron. ‘thap’). As the material in which Tsewang Norbu heats the metal, supported by bellows made of animal skin, has to be refractory, he was always confronted with the need for proper earth materials which are actually the same which he uses for making a thab, i.e. the thab sa. His nickname Garbo (mgar ba meaning ‘blacksmith’) refers to the commonalty of blacksmiths. 8 Terminological addition by Erberto Lo Bue, personal communication by e-mail, 30 August 2012. 9 Hamid (1998:172).

374

feiglstorfer

Figure 16.5

Lamayuru. Clay pit with a striped ‘markalak’. Photo: Hubert Feiglstorfer.

brother, who was drilling for water in this area professionally.10 Today these names are no longer related to a certain kind of earth or to pottery in general. The ‘chukalak’ (chu ka lag), considered by local people to be a lower grade of ‘markalak’, basically refers to the clayey silt that one can find at the bottom of irrigation ponds or flood plains, and this is perhaps the reason why it is called ‘chukalak’, chu meaning ‘water’ or ‘river’. Although Spituk clay is technically fine silt, it is called clay (‘markalak’) by common people, and the term ‘chukalak’ (or ‘thukalak’) is used to distinguish it from the coarser clayey silt found in similar places such as Phey.11 While the above-mentioned types of earth are of a comparatively fine quality, the soils of Leh and Alchi are generally described as being sandy. In Lamayuru one can find relatively fine ‘markalak’ (Figure 16.5) which is suitable for use as plaster, mixed with a certain amount of sand to avoid cracks. On the other hand, the local earth is relatively coarse, with quite large stones 10 11

Interview with Sonam Wangchuk in August 2011 in Basgo. This paragraph, concerning the ‘chukalak’ was added by Sonam Wangchuk in a correspondence in October 2011. Furthermore he adds: “However I have found in tests that chukalag around Phey has more real clay particles and coarse silt than the Spituk markala which appears to be uniform fine silt, so that’s why it is possibly like a mix of Basgo and Spituk fine silt”.

revealing traditions in earthen architecture

Figure 16.6

Wanla, Sumtsek. Interior view of a wall section beside the portal, showing a detachment of the plaster. Photo: Hubert Feiglstorfer.

375

376

feiglstorfer

up to several centimetres in size. This material is used for bricks without any further mixing, resulting in building material with a high proportion of stones for which I have been unable to find any comparison so far. Interestingly, it was difficult to break such a stony brick, suggesting that this earth contains a certain amount of clay. In Wanla different sites for digging earth for building purposes are known. The field soil in the area close to the village school is locally known as a source of good material for bricks. It is processed into bricks without any further mixing. For plastering purposes, the soil is mixed with river sand for stabilizing and as a crack-stopper. The soil that can be found along the river bank is also locally known as a high quality building material. To use it as plaster, it is mixed with straw and sand. In Urtsi a red soil similar to the one that can be found at Basgo is mixed with straw and sand for use as plaster. An interesting find occurred along the right wall on the ground floor of the Sumtsek (gsum brtsegs) at Wanla (Figure 16.6) as one enters the temple.12 In that area a section of plaster about four centimetres thick and of a relatively large expanse gapes away at a distance of several centimetres from the outside wall. The plaster contains a small amount of straw, but what was more fascinating to me was the fact that it was not primarily the straw that kept the plaster together, as the visible content of straw was not sufficient for that purpose. Instead, the flat and round pieces of sand of a size of up to two to four millimetres seemed to have taken over that technical function. That material, which is locally known as the main additive to the local soil, brings us back to the aforementioned sand of the river bed. This example underlines the importance of the quality of the additive, which has to be of the right size and shape, a circumstance that has always to be considered of high importance besides the quality of the earth itself.

Specific Additives and Organic Components

To improve the quality of particular kinds of clay, especially their strength and cohesiveness, in several areas the basic building material is refined by adding specific organic additives and mixing different qualities of soil with each other, according to particular local traditions. In several exterior walls, for example in some walls of the inner ambulatory of the main temple Tsuklakhang 12

Susan Nitsche who was working as a conservator of wall paintings in the Sumtsek in Wanla for the Achi Association at the time of my visit brought my attention to the consistency of the earth used in this crack.

revealing traditions in earthen architecture

Figure 16.7

377

Nyarma, Tsuklakhang, Nangkor. Exterior plaster mixed with a high amount of straw. Photo: Hubert Feiglstorfer.

(Tib. gtsug lag khang) of Nyarma, a relatively large amount of straw was mixed with the plaster (Figure 16.7). Local accounts mention the use of pulp from Ladakhi apricots (Tib. cu li, pron. ‘chuli’) mixed with soil to increase its inner cohesion. In many households the main part of the apricot which is used after picking and drying on flat roofs is the almond-like core for producing oil. The pulp is often treated as waste and in that way it finds an efficient use as a stabilizer of earth materials. Apricot sap is known to be added to the rdza sa. The ‘sasung’ oil extracted from mustard seeds is known to be used to improve the surface of Tibetan stoves made of earth. With as little as one litre it is possible to treat about two square metres of surface. Mustard seeds grow all over Ladakh, but only a few people sow them. Boiling of walnut shells is known to produce oil for the same purpose. Walnut is available in the regions of Lower Ladakh and Sham.13 A similar use was explained to me by a teacher, Radha Krishan of Ribba,14 a village in Kinnaur. In earlier days ‘chid’ (Kinnauri), a very fine local clay used as upper roof layer, was mixed with ‘gum’ (Kinnauri), the 13 14

Interview with Mr. Angchuk, an engineer originally from Basgo, in August 2011 at his home in Leh. Interview with Radha Krishan in 2002 at Ribba, where he taught at the time of the interview.

378

feiglstorfer

sap of the ‘chuli’ tree (Prunus armeniaca) and water. Today this natural ingredient is replaced by wood glue. Many of the samples from all the evaluated regions use straw as an additive to strengthen earth material. Primarily it may be found in earth which is used as plaster (rdo sho, pron. ‘do scho’), but in a few cases also in earthen material used for roof construction and adobe bricks. In areas up to about 3200 metres, which are richer in woodlands than the barren high plateaus above that altitude, the use of pine needles is common. Several samples from Lahul, for example from Keylong or from Purthi in the Pangi valley, also from Thangi in the Lower Kinnaur, contain needles from the Kail tree (Pinus excelsa) as found in Purthi or from the Chilghoza tree (Pinus gerardiana) as found in Thangi. Several samples do not show any additives, as in one sample from Dhankar, or in some interior plasters in Nyarma (6073), or also in the interior plaster covering the southern wall below the wall paintings in the Lhakhang Marpo in Hunder (6044). However, others, such as the adobe bricks used on the northern wall below the joist bearings in the Dromton Lhakhang (6053) in Tabo or the interior plaster in the same temple (6041) found about 1.5 metres above the floor level, contain a relatively high amount of straw. Some samples from the Spiti, Pin, Nubra valleys and from Lahul show tiny fibres among the mineral grains under the microscope. The existence of small pieces of coal in the samples of the earthen roof layer in the old part of the temples at Sumur (6047) and Hunder (6043) points to anthropogenic additives. Their presence in these samples may be either accidental or intentional, possibly the result of reusing sooted interior kitchen plaster which is known to provide building material of high quality. In Basgo it is also known as ‘khu sa’ and is used as upper roof layer instead of or together with ‘markalak’ to make the upper surface waterproof.

Terms Related to Earth as Building Material

Several terms in the Tibetan and Ladakhi languages relate to soil as earthen building material. One of the most often mentioned terms for the soil used for building purposes is ‘markalak’. Under this term it is known in Tibetan as well as in Ladakhi. This term is used in Ladakh, Zangskar, Spiti, Lahul, Kinnaur, as well as Purang, where it is also known as ‘narkala’. It describes, with some variations, fine earthen building material of a high quality. Since its grains are relatively fine and its texture is smooth, this material is commonly preferred for the final surfaces of walls, as the base coat for wall paintings and also for the upper layer of roof constructions. A closer look into its components shows

revealing traditions in earthen architecture

379

obviously different types of ‘markalak’ soil of different colours mainly due to the content of particular varieties of clay mineral. The Ladakhi and Tibetan terms translated into English as ‘mud’, namely ‘kalak’ (Lad. and Tib. ka lag), ‘dam’ (Lad. ldam,15 Tib. ’dam, meaning also ‘gesso’ and ‘clay’) and as ‘dakpàk’ (Tib.’dag bag), refer to soil material mixed with water, similarly to the explanation given by Bielmeier as mud as a mixture of earth and water used instead of clay. Bielmeier’s dictionary also gives an explanation of the term ka lag (cf. ldam ka lag, mar ka lag) as a sort of stone whose powder may replace cement.16 Since such an explanation does not clarify the question of the type of clay, contrary to the ‘markalak’, the specific features of ka lag remain unclear, pending chemical analysis, although three Tibetan dictionaries, as reported by the Dharma Dictionary, give the following four definitions: ‘mortar’, ‘mud’, ‘kind of rock [its powder usable in place of aar ‘dam] earth and water used instead of mortar’, and ‘earth and water used instead of mortar’. If implying a comparison with cement, ka lag might refer to thab sa or rdza sa, which is also known to become as hard as stone. On the other hand, dka rag (pron. ‘karak’) is known to be used as plaster and indeed is defined as ‘white earth to be smeared on houses etc., whitewash’ in Tibetan dictionaries. So it might be suggested that ka lag and dkar rag refer to mixtures of similar earthen materials with water.17 Materials Used for Whitewash Bielmeier18 mentions ‘markalak’ also as a fine ochreous earth, used as groundcolour when staining houses with dkar rtsi (pron. ‘kartsi’), i.e. lime used to whitewash the buildings’ walls.19 For this whitewashing we may differentiate between the use of particular kinds of soil mixed with water (dkar rag) on the one hand and the use of lime on the other. The lime used as building material can be divided into two further types. The process for extracting lime was 15 Hamid (1998:136). 16 Bielmeier et al. Forthcoming. 17 Bielmeier et al. Forthcoming: This dictionary mentions the pronunciation as qalaq, kalaq or kalak. The use of this term as mud or clay in Tabo as a general term might underline the interpretation given above. The definition of this term varies according to the qualities of wetted soil and slurry, described by the terms feuchte Erde and Lehmschlämme in German. 18 Bielmeier et al. Forthcoming. 19 Bielmeier et al. Forthcoming: The term dkar rag is also mentioned as whitewash used on the outside of buildings. This translation correlates with the one known in Ngari, according to a translation by Tsering Gyalbo in October 2011. For more meanings of this Tibetan term see for example the Dharma Dictionary.

380

feiglstorfer

described by the family of potters living in Likir, probably the only village in Ladakh which keeps the tradition of pottery making alive, as follows.20 The chemical composition has been added by the author. The raw material (Figure 16.8) is not clean, but is obtained from rocks which may be splintered easily into small pieces. After being washed, this raw material is dried and ground into powder. Afterwards, the powdered dkar rtsi (CaCO3) is heated in the same big metal pans that are used for roasting barley grains to produce parched flour (Tib. rtsam pa, rngan phye).21 During this step carbon dioxide (CO2) escapes and anhydrous lime (CaO) remains. Then the lime is cooled down with warm water to receive chloral hydrate (Ca(OH)2), the basis for whitewash. Finally a larger amount of water is added. At Khorchag, in Purang, troughs in the earth along the road in the centre of the village are used to contain the impure chloral hydrate, so that heavier impurities as well as unburned pieces of lime deposit on the bottom of the troughs. In other parts of Ladakh the raw material is not as easily accessible as in Likir. Also the quality of the lime differs from pit to pit. In Lamayuru, for example, dkar rtsi is traditionally found in the mountains south of the village. As these places are not accessible by road, the material has to be brought down on foot with the help of donkeys. The ascent lasts around four to five hours and the load on each donkey is limited to around 10 to 20 kg. For example, the two big chortens beside the monastery of Lamayuru as well as the monastery itself are painted with this kind of lime. As a result of the huge effort required to collect the raw material, local practice has changed and the purchase of ready-made lime at the market has become common. The lime from Kanji is likewise known for its high quality, but also in this case the access to the material, which is located about five kilometres away from Kanji towards the Leh-Kargil highway up along the mountains, is arduous. The dkar rtsi from Mangyu is known not to be very water resistant. It can be dug along the mountain northeast of Mangyu, on the far side of the river valley. Traditionally it is mixed with the fat of animal skins. The dmar rtsi from Mangyu, which is dug in the mountains west of Mangyu, is known to be of low quality. Instead the dmar rtsi from Wanla is used (Figure 16.9). It is known to be of a better quality, which means that it lasts longer than the darker material found in Mangyu.22 20 21 22

For information on the relevant interview see above, n. 7. Translation into Ladakhi according to Bettina Zeisler, University of Tübingen, to whom I am very grateful for her support. The information on the dkar rtsi and the dmar rtsi from Mangyu and Wanla was given by Sandeep Kumar from Uttarachand, close to Rishikesh. At the time of the interview, which was held in August 2011 at Mangyu, he worked as conservator for INTACH on the restoration of the wall paintings in Mangyu temples.

revealing traditions in earthen architecture

Figure 16.8

381

Likir. The white spot in the centre of the picture is dkar rtsi, a local species of lime. This brittle, slaty material has to be knocked out of the surrounding rock. Photo: Hubert Feiglstorfer.

382

Figure 16.9

feiglstorfer

Wanla, chorten. According to Sandeep Kumar, locally available dmar rtsi was used for the red colour. Photo: Hubert Feiglstorfer.

revealing traditions in earthen architecture

383

Figure 16.10 Basgo. Clay pit with a specific kind of white shining ‘markalak’. Photo: Hubert Feiglstorfer.

In the eastern areas, like Nyoma in Changthang, places for getting dkar rtsi are known. In Urtsi, in the Wanla area, a black stone (sa nag, pron. ‘sanak’, literally ‘black earth’) is used with the addition of water to colour the spen pad (apparently also span bad, pron. respectively ‘pénpè’ and ‘pénbè’) friezes running around parapets or the windows’ black decorative frames. Another explanation found in Bielmeier’s dictionary in relation to the white mineral used for whitewash refers to its colour as found in certain areas. For instance, on the hills west of Basgo ‘markalak’ appears as shining white (Figure 16.10), similar to the colour of the clay mineral known as kaolin, whereas in Spituk its colour is bright greyish brown and in Purang it has an even more grey colour. Since such different qualities of building material are called with the same name, we can state that the term ‘markalak’ is commonly used with reference to similar building materials sharing specific qualities, as mentioned above. The Traditional Tibetan Earthen Stove In conclusion, I would like to present instructions for making a traditional Tibetan stove (thab) as a sculptured piece of architecture. Its position in the

384

feiglstorfer

Figure 16.11

Basgo. Traditional Tibetan stove (thab), located in the house of Mr. Angchuk’s brother. Photo: Hubert Feiglstorfer.

living room is determined by quick accessibility of both dishes and ingredients by the cook, mostly the housewife, and by a good position for communication between the people working at the stove and those sitting on the floor. The central role played by the stove in a local Tibetan house is appropriately described by the Tibetan and Ladakhi word for kitchen, namely thab tshang (Figure 16.11), the word tshang meaning ‘home’, ‘household’, ‘family’, ‘dwelling’ and so forth. As the material of the stove has changed during the last hundred years successively from clay into metal, the tradition of making a thab, including the knowledge about finding the right material for its production, is getting lost. Just a few local people still have this knowledge and even fewer really carry on that craft. Together with Sonam Wangchuk from Phey it was possible to find one of the few remaining stove makers in Ladakh, Tsewang Norbu from Nye, near Tunlung, and to receive the following instructions for making a thab.23

23

This interview was conducted with Sonam Wangchuk, who translated into English, at the home of Tsewang Norbu in Nye in August 2011.

revealing traditions in earthen architecture

385

Figure 16.12 Tunlung. Clay pit with orange-brown rdza sa. Photo: Hubert Feiglstorfer.

The basic material is rdza sa. Its colour is yellowish brown and it can be found in Tunlung (Figure 16.12) on the slope beside the road, approximately opposite the water ‘mani’ prayer-wheel. The suitable material is the one which is soft and can be easily crushed between two fingers. Material appearing as lumps, hard like stone, should not be used. At the clay pit we found that the difference lies not so much in the material itself, but rather in its dryness or dampness, slightly humid material being easier to crush. There is another kind of soil, also known as rdza sa, which can be found at the foot of the hills close to Nye and this is said to be of a better quality than the material from Tunlung. Unlike the material from Tunlung, the clay from Nye is covered by a white crust from which the clay has to be separated. The appearance of this material is similar to a kind of rdza sa that can be found at Leh. Water is added to soften the clay. This process takes several hours with the soil from Nye, while it takes about one day with the soil from Tunlung. The first step in processing the raw material is to beat it by adding water which is a tiresome job. After beating the earth, more water is added but no sand, since enough of that is contained inside the raw material. After its grounding with a stone, it is ready to be modelled into a stove. With that material a full-size stove of a cuboid shape is obtained, a single operation taking a day by adding one splash of rdza sa after the other. After completing that basic shape, the holes and air channels of the stove are roughly dug out by hand, and

386

feiglstorfer

after that this mass of earth has to dry for about a week. In the meantime ornaments, often showing the eight auspicious symbols of Mahāyāna Buddhism (Tib. bkra shis rtags brgyad), are modelled along the surface of the walls and after five days the side walls of the stove are cut so as to give them a vertical shape. Two days later, after a total of one week, the stove should have dried and show no cracks before the preparation of its upper surface. The surface is smoothened with a trowel, then egg white is added and the process of smoothening is repeated. With the egg white a binding ground is obtained to add the last layer, consisting of sieved soot mixed with water. As with shoe polish, the soot paste is added as the outermost surface. Probably this last layer makes the surface water-repellent. Soot is, for example, mixed with the interior plaster to make walls waterproof in the Persian hammam, as I could observe in Yazd, in Iran. All in all, after about two weeks the stove can be lit for the first time, not earlier, so as to avoid any risk of cracking. Conclusions Local terms for certain types of soil often serve to categorize their use and their origin. Different kinds of earth have different qualities as building materials, either on their own or when blended with other types of soil or organic additives. The traditions that apply to their use as building materials should therefore not be seen as universally applicable, but more as local developments making use of the possibilities of locally available materials. In many cases the different uses of particular materials, and even the materials themselves, are known only to a few elderly people. Accordingly, one of our main tasks must be the documentation of traditions that are in danger of being lost. In this context, I deliberately do not insist on the stubborn preservation of traditions,24 since ultimately the locally defined architectural overall context of each building should determine our approach to this knowledge. On the one hand, the conservation of a historical building 24

My article, Research on earthen roof constructions of Buddhist temples in the Western Himalaya, submitted for publication in the proceedings volume of “Terra 2012”, the 11th International Conference on the Study and Conservation of Earthen Architecture, held from April 23rd to 27th 2012 in Lima, in Peru, deals with the construction of traditional Tibetan roofs in particular, in a technical and socio-anthropological context. In that respect the keeping of traditions has to be clearly differentiated from their understanding and from their reusing in a contemporary context.

revealing traditions in earthen architecture

387

may require the use of traditional knowledge in a conservative sense. On the other hand the appropriate balance between traditional and contemporary materials and techniques will depend on the conservator’s understanding of building structures as well as on his flexibility as designer.25 Similar considerations apply not only to conservators but also to planners of new buildings which are of course more numerous than historic buildings in need of repair. Each planner finds himself confronted with the question whether and how to use traditional materials and techniques. These often require heavy investments in labour as well as materials, and this points to the need for case-by-case solutions. A deeper understanding of the origin, composition and processing of the locally sourced materials that have been used for hundreds of years will help us judge how best to apply traditional techniques today. Acknowledgements My research in Ngari in 2010 was financially supported by the FWF Research Project P-21806-G19 Society, Power and Religion in pre-modern Western Tibet, under the auspices of the Institute for Social Anthropology and the Centre for Studies in Asian Cultures and Social Anthropology based at the Austrian Academy of Sciences (ÖAW). Research carried out in 2011 was also supported by the Österreichische Forschungsgesellschaft (ÖFG). Regarding the material science in this contribution, including mineral and geologic analyses, I am indebted to Prof. Franz Ottner, who teaches at the Institute of Applied Geology at the University of Natural Resources and Life Sciences in Vienna, for his continuous and longstanding support. Furthermore I am very grateful to Dr. Edith Haslinger, a scientist at the AIT Austrian Institute of Technology in Tulln, for peer reviewing the geological part of my contribution. I am indebted to many people for their long-lasting support and encouragement which was invaluable for the successful completion of this research work. In particular, I have to thank the Austrian Science Fund (FWF), project P-21806-G19 and Dr. Christian Jahoda who reviewed this contribution, except the geological and material scientific passages. For the final peer review I wish to thank Dr. Erberto Lo Bue and John Bray, as well as Sonam Wangchuk and Rebecca Norman who guided me to specific places in Ladakh and supported me in collecting earth samples and translating during interviews. 25

For a discussion of similar themes in relation to the conservation of the Leh Old Town see the contribution by Alexander & Catanese in this volume.

Conservation of Architectural Heritage in Ladakh John Harrison

Figure 17.1

Prayer flags and telephone mast in Leh: communicating with other worlds. Photo by John Harrison.

© koninklijke brill nv, leiden, ���4 | doi ��.��63/9789004271807_019

conservation of architectural heritage in ladakh



389

What is “architectural heritage”?

The question was posed at the opening session of the International Association for Ladakh Studies (IALS) 2007 symposium in Rome, the Eternal City. The symposium was held in the Faculty of Oriental Studies of the “Sapienza” University of Rome, in a group of historic buildings which have been restored and successfully adapted for modern educational use, surrounded by astonishing monuments which illustrate the history of one of the world’s great cities over a period of 2,000 years. In this context, and to an audience of educated Ladakhis and Westerners, the question may have seemed superfluous, but in Ladakh itself the concept of architectural heritage is still developing. It is not yet defined, nor widely accepted. And while it is in this state of flux, the architectural heritage of Ladakh is continuing to disappear—demolished, left to decay, irrevocably altered, ‘modernised’.

The Preservation of Historic Buildings

And why should historic buildings be preserved? This is another question which has been asked again and again over the centuries. The responses in favour of preservation have cited the value of old buildings as tangible evidence of a society’s history, its roots in a particular time and place, the reason why a culture is what it is in the 21st century. However, these are all rather nebulous concepts which are countered by the hard-nosed facts of escalating land values in prime commercial locations, the financial cost of staving off physical decay, and the tempting glitter of new materials and modern forms in a rapidly changing world. The Emperor Majorian of Rome in an edict of 458 CE tried to prevent the demolition of historic buildings for their building materials. It was not effective. By then the Empire was in decline, and the population of Rome dropped from a million at the beginning of the first millennium CE to 30,000 in the medieval period: the size of Leh today. Roman monuments continued to be used as quarries for 1500 years, right through the Renaissance and Baroque rebuilding of the city by the Popes. But an appreciation of Roman ruins and antiquity also developed during the Renaissance, from the poet Petrarch (1304–1374) to artists and architects Alberti (1404–1474), Raphael (1483–1520) and Michelangelo (1475–1564), and was then taken up by scholars and cognoscenti from England, France and Germany during the 18th century Age of Enlightenment.

390

harrison

Evermore wide-ranging movements for the preservation of historic buildings emerged in Europe in the 19th and 20th centuries, and from Britain’s expansion to colonial India with the founding of the Archaeological Survey of India (ASI), particularly under the directorship of Sir John Marshall from 1902 to 1928. The creation of the United Nations, including UNESCO, after the massive destruction of the Second World War, has led to a globalisation of heritage preservation, but also recently to suggestions that the 1964 International Charter for Conservation and Restoration of Monuments and Sites (the Venice Charter) is too Euro-centric, and does not take into sufficient account the very different cultures of Asia, Africa, Latin-America and Australasia. Subsequent documents on cultural heritage, such as the Burra Charter, which was approved in Australia in 1979, have tried to address these regional concerns. Further international institutions established by UNESCO include the International Centre for the Study of the Preservation and Restoration of Cultural Property (ICCROM), based in Rome; and the International Council on Monuments and Sites (ICOMOS). UNESCO also assesses applications for recognition of World Heritage Sites submitted by national governments. These are sometimes seen as national status symbols, like membership of the World Trade Organisation (WTO), or possession of a nuclear arsenal.

The Architectural Heritage of Ladakh

India has a number of World Heritage Sites, and a UNESCO office in Delhi, but the only legally protected historic sites and buildings are the National Monuments listed and sometimes owned by the ASI. The ASI has a school of archaeology and research laboratories, but limited funds to maintain both a vast number of monuments spread over a vast country, and a cumbersome bureaucracy. There are 13 National Monuments in Ladakh, including the Leh royal palace, the early temples and paintings at Alchi, monasteries such as Thikse, and the rock-carved Maitreya at Mulbek. But is this All? What is the Architectural Heritage of Ladakh? In Europe and the West the definition of architectural heritage initially included only major monuments, cathedrals, castles, and archaeological sites such as Stonehenge, but has gradually extended to include ordinary houses, early industrial sites, milestones, post-boxes, historic groups of buildings, neighbourhoods, villages, even entire historic towns.

conservation of architectural heritage in ladakh

Figure 17.2

391

Mulbek, colossal rock-carved statue of Maitreya, between the 9th and 11th century. Photo by John Harrison.

The Indian National Trust for Art and the Cultural Heritage (INTACH), an NGO “established in 1984 to identify and protect the many thousands of monuments and historic buildings that fall outside the purview of the Archaeological Survey of India”, has published a more expansive list of 73 heritage sites in Ladakh (Sharma 2003). (Zanskar will be published separately). Although it includes Nyarma, which Gerald Kozicz discusses elsewhere in this volume (Chapter 5), all the monasteries and a few important houses, there are surprising omissions: no Chigtan castle (Chapter 10), nor many other castle ruins, nothing in Gya (Chapter 3), no historic village sites, no Skurbuchan (Figure. 17.3), no Kanji (Chapter 8). Listing by INTACH confers no statutory protection to sites and buildings, but it does identify them locally and nationally as valuable elements to be preserved. The Namgyal Institute for Research in Ladakhi Arts and Culture (NIRLAC), a Ladakhi NGO, has compiled a much more comprehensive list, village by village throughout Ladakh and Zanskar, which includes every chorten and rock carving as well as historic buildings. Villagers and gobas (the village headmen) have been involved in the fieldwork, and every village goba will have a copy of the list, so that this exercise stands a good chance of local acceptance (NIRLAC 2011). The Ladakhi scholar Phuntsog Dorjay is studying early Buddhist rock art in Ladakh (Chapter 2 in this volume), and Tashi Ldawa Thangspa is making a comprehensive documentation of petroglyph carvings (Chapter 1).

392

harrison

Figure 17.3

Skurbuchan Khar. Drawing by John Harrison.

So with these inventories, and other exercises in the documentation of individual buildings and settlements, we are beginning to have a much better idea of what constitutes the architectural heritage of Ladakh, an essential first step in conserving that heritage.

Current Conservation Work

What is Actually Happening on the Ground Now? The ASI is currently involved in work at Lamayuru, Thikse and Phyang monasteries, and Shey Palace as well as continuing major work at Leh Palace. It has no further plans for continuing with the restoration of the huge earth-built Tiseru Stupa for lack of evidence for the original appearance of the upper storeys. Seven of the officially-designated National Monuments in Ladakh are monasteries or temples, all in active religious use, and this has led to a number of conflicts between what the ASI perceives as the international norms of conserv­ation, and the monastic community’s desire for development (preferably in reinforced concrete). Lamayuru has demolished most of the smaller traditional buildings on the approach to the monastery in order to extend the dance ground and erect a

conservation of architectural heritage in ladakh

393

huge four-storey hotel. One side of this building is occasionally used to hang a large thangka, but the view from outside is of a repetitive concrete frame competing in size with the monastery itself. Hemis ignored a sensitive study by INTACH of the repairs needed to its historic buildings, and demolished its entrance courtyard to build a concrete framed grandstand to accommodate more tourists watching the chams dance at the annual festival in July. The monastery has since added a large and unlovely concrete museum, is rebuilding the historic assembly hall without expert advice, and is seeking to be removed from the list of national monuments. It is sad to see such an important institution in the life of Ladakh actively destroying the physical remains of its own history. However, this is perhaps an indication of the cultural conflicts and confusions which face a modernising Ladakh today. From the Jammu & Kashmir (J&K) state government there are hints of money from the official budget for the promotion of tourism. The J&K masterplan for Leh, long in production, has been rejected by the Ladakh Autonomous Hill Development Council, and in any case said nothing about conservation. The Hill Council have now appointed consultants to produce their own masterplan . . . The Hill Council professes to support the conservation of the architectural heritage, and has entered into an agreement with the Tibet Heritage Fund (THF—see Chapter 15) for their work in Leh Old Town, but government action on the ground is hard to discern. From all parties lack of funding is cited as the reason for inaction. INTACH started its work in Ladakh with its Leh development plan drafted in 1988 (Chaturvedi 1989),1 then a detailed study of Hemis Monastery, neither of which bore fruit. More recently they have repaired two early chortens at Mangyu, made repairs at Saspotse temple to protect wall paintings, planned future work at Karsha monastery in Zanskar, and are now working at the Mangyu temples. Interestingly, architect Janhwij Sharma, when at INTACH, worked on a Likir village development plan, following the success of a pilot project at Raghurajpur in Orissa, in sustainable rural heritage tourism—controlling tourism funding and tourist activity for the benefit of the local community, rather than the free-for-all which is all too evident elsewhere, particularly in Leh. 1 The INTACH conservation plan was prepared by Romi Khosla and the Group for Rural and Urban Planning (GRUP). Romi Khosla had earlier published Buddhist Monasteries of the Western Himalaya (Kathmandu, Ratna Pustak Bhandar, 1979), the first architectural documentation of Ladakhi monasteries following his epic journeys through Lahoul, Spiti and Ladakh in the 1960s and 70s.

394

harrison

A further project was proposed at Samstanling village in Nubra, but the promised funding for these projects from the J&K state has not yet materialised. A Ladakh Chapter of INTACH, working independently of Delhi, was established in 2010 with Tsering Angchok of Basgo as Convenor, and is developing building and painting conservation projects in Igu, Gya, Apati and Chushot. NIRLAC has undertaken restoration work at Sumda Chung temple, funded by the World Monument Fund (WMF), which also grant-aided the impressive Basgo citadel works with the Basgo Village Welfare Committee. Basgo has recently been awarded the UNESCO Asia Pacific Heritage Award of Excellence. The Swiss-based Achi Association is continuing building and painting conservation work on early Drikung Kagyu buildings: at the small 14th century Tsuglagkhang in Kanji, and at the three-storey temple in Wanla. THF has been working at Alchi on the conservation of the buildings and wall paintings of the Tsatsapuri group of temples, and is planning the repair and upgrading of the impressive Lonpo House in Alchi. A Czech group started a programme of building repair and painting conservation at Diskit monastery in Nubra in 2006, and a team of Hungarians are repairing Zangla castle in Zanskar where their compatriot Alexander Csoma de Kőrös (1784?–1842) lived and studied in the early 19th century. Most of these projects are individual monuments. At Likir and Samstanling villages there was the beginning, unfortunately now shelved, of area conservation, addressing problems of infrastructure, employment and community life as well as historic buildings. At Basgo the WMF may next support this type of work in the village to complement the successful work on the citadel monuments. Leh is a much more complex problem. The J&K government missed the opportunity to develop and implement the 1988 INTACH plan, and since then everything has got a lot worse—congestion, pollution, traffic, inadequate and defective infrastructure, and an unplanned and uncontrolled explosive expansion of the commercial centre and the urban area, mostly tourism-generated. The J&K government keeps its distance, and the Hill Council appears to be out of its depth. In this chaotic context it is good to report that progress is being made in the conservation of the Old Town, on the steep rock face below the palace. Leh Old Town is important because it is such a remarkable survival; while commercial redevelopment has destroyed virtually all the historic 19th century buildings in Main Market below, the earlier Old Town has been bypassed and much has survived. Fortuitously its lack of road access, mains water and sewerage has meant that many old Leh families moved out to the suburbs in the later 20th century, leaving the big houses to decay. In a more accessible

conservation of architectural heritage in ladakh

FigURE 17.4

395

Leh Old Town, section from Tashi Khangsar House down to Abdul Wahid Choskor House behind the Jami Masjid. Drawing by John Harrison.

location these historic houses would have been rebuilt as modern structures by the owners, as has happened elsewhere in Leh and throughout the villages of Ladakh. In the wider Himalayan and Tibetan world too Leh is important as a surviving urban ensemble. In Tibet itself, where historically there were never many urban settlements, Lhasa, Gyantse and Shigatse have been largely redeveloped as modern Chinese cities, and the Barkor around the Jokhang temple in Lhasa has become a tourist shopping mall. In Nepal the Kathmandu Valley towns have lost much of their historic character in recent years. Only Bhaktapur, with a strong conservation planning policy, has survived relatively unscathed, whilst in Patan and Kathmandu the major monuments remain, but surrounded by gimcrack multi-storey concrete structures waiting for the next earthquake. In 2002 the Ladakhi architect Deldan Angmo and the writer looked at the possibility of an urban conservation project in Leh, and in 2003 the THF, which had previously worked on the historic centre of Lhasa, started the Leh Old Town Initiative (LOTI—see Chapter 15). Generally paying 50% of building costs with funds provided by Western governments and charitable donors, THF has refurbished a number of historic houses for private owners, completed two street paving and drainage schemes, repaired the Guru Lhakhang and the Gompa Soma courtyard, and uncovered wonderful 14th century wall paintings in Chamba Lhakhang. It has also repaired the small White Chamba Lhakhang, hidden in the centre of the Old Town, and is working on the Rupshu House, a

396

Figure 17.5

harrison

Lonpo House, Leh. Drawing by John Harrison.

great block of masonry below Gompa Soma which had lost its two upper floors and four ‘rabsal’ balconies. The old mosque off Chhutay Ratak (the old lane of aromatic bakeries behind the main mosque) and adjoining houses have been rebuilt, and a new museum of the Central Asian trade has obtained funding from the J & K state and is nearing completion on a site next to the mosque. Also in the Old Town, from 2003 to 2005 INTACH UK funding obtained by the writer secured the restoration of the Lonpo House for Chemre monastery. This was one of the important buildings identified in 2002 as being at risk. In a prominent position at the foot of the palace, the monks’ accommodation there had been abandoned after a fire and the collapse of the front wall, and it was likely that the entire structure would have fallen into ruin and been lost. The Himalayan Cultural Heritage Foundation is now housed in Lonpo House. Another building at risk was the historic Munshi House with its two delicately carved balconies standing out above the smaller houses below. The Munshi family had already left the Old Town and built a new house in the fields below, when in 1987 INTACH proposed restoring the empty house as a museum of folk art. Initial government support for the proposal lapsed as key officials were posted elsewhere, and the building continued to decay for another 20 years until the Ladakh Arts and Media Organisation (LAMO) started restoration and conversion of the building as an arts centre. With the rebuilding of

conservation of architectural heritage in ladakh

Figure 17.6

397

LAMO centre and the restored Munshi House, Leh. Drawing by John Harrison.

the ruined Gyaoo House next door and a new block built on the massive stone walls of an earlier structure, this has created a library and exhibition spaces for Ladakhi arts and crafts, and meeting and teaching spaces in the restored historic interiors. This project shows how a historic building, considered redundant as a family house in changing social and economic conditions, could be adapted and reinterpreted for new community uses and become an active neighbourhood centre. So two new museums are proposed in the Old Town for the first time in Leh. Perhaps this is a sign of a cultural rebirth for a neighbourhood which has seen a gradual decline both physically and socially. But it is important that the Old Town remains a living community while the buildings are restored and improved. It must not become just a museum quarter and a tourist honeypot.

The Future

In spite of the commercial pressures for redevelopment in Leh, and the impact of uncontrolled tourism on the environment and society generally, there are now some grounds for optimism on the survival of the architectural heritage in Ladakh. Progress is being made, but apart from NIRLAC and Basgo, too much of the current work is still being undertaken by outsiders, from elsewhere in India and the West. This is perhaps acceptable in the short term, but if the

398

harrison

architectural heritage is to survive and become an essential part of the future, Ladakhis must be involved to a much greater extent. There must be more education, in schools, in communities, in the media, to increase public awareness and encourage participation. NIRLAC has been working in this field, with workshops for villages and craftsmen, and World Heritage Day. There were also the 2004 UNESCO/NIRLAC workshops on monastic heritage in Basgo and Hemis, and the Ladakh Ecological Development Group (LEDeG) workshop organised by architect Anuradha Chaturvedi. The Achi Association has started an educational programme funded by the Getty Foundation for Drikung Kagyu student monks and nuns and Ladakhi youngsters, with classes in Dehra Dun and fieldwork in Ladakh, so that a new generation will learn to appreciate its heritage and how it can continue to be used. Already in 2012 an enthusiastic young team from the recently registered Achi Association India is working with the village community to restore Skurbuchan Khar. The Himalayan Cultural Heritage Foundation (HCHF) was set up in 2009 by Dr Sonam Wangchuk, formerly with NIRLAC, and is based in the restored Lonpo House in Leh. As well as lectures and exhibitions they have held workshops for monks, craftsmen and young musicians, and are now working with the Hungarian architects and an Italian group on stupa restoration. In 2012 the new Ladakh Chapter of INTACH and the National Museum Institute from Delhi held a seminar on preservation in Leh and a training workshop at Phyang monastery aimed particularly at the monastic community. The International Association for Ladakh Studies (IALS) held its biennial symposia in Leh in 2009 and 2011 and attracted a record number of local participants and scholars from around the world, again showing Ladakhis the esteem in which their culture is held outside Ladakh itself. Concerns might be expressed about this multiplication of agencies working in the conservation field, and the danger of duplication of effort or perhaps even damage to monuments by amateur inexperienced enthusiasts (The early work of a stupa restoration group, or temple restoration by a famous Indian photographer, spring to mind). In 2011, prompted by André Alexander of THF, some moves were made towards co-ordinating the activities of all these groups, perhaps by establishing a heritage forum to meet regularly and discuss projects, research and techniques. Following André’s untimely death in early 2012, this idea is currently in abeyance, but one hopes that other players in the field will take it forward. We saw above that only the few national monuments are protected in law against alteration and demolition, and that even in these cases the ASI has often been unable to enforce compliance on strong private owners. Extending

conservation of architectural heritage in ladakh

399

the legislation, either locally or nationally, to include a much wider range of buildings, would therefore not be effective in protecting the architectural heritage of Ladakh. This can only be done by agreement, by the support of building owners and the wider public. The current restoration work in Leh Old Town is having a positive effect on local people’s perceptions of historic buildings and their reuse, and this is reinforced by every completed building, in spite of the absence of a framework of conservation planning and services from local government. Private owners in the Old Town have welcomed the opportunity to upgrade their houses and improve their living conditions, and are giving a very practical demonstration of the continuing value of historic buildings and the future of a very special historic place. Indeed, the future of the past.

Bibliography Administrative Committee of Archaeology of the Tibet Autonomous Region. 1991. The Site of the Ancient Guge Kingdom (Gu-ge’i gna’-grong rjes-shul). Beijing: Cultural Relics Publishing House. 2 vols (in Chinese, with English summary). Agajanov, S.G. 1998. “The States of the Oghuz, the Kimek and the Kipchak.” In History of Civilizations of Central Asia. Vol. IV: The Age of Achievement: A.D. 750 to the End of the Fifteenth Century. Part One: The Historical, Social and Economic Setting, pp. 61–76. Eds. Muhammad S. Asimov and Clifford E. Bosworth. Paris: UNESCO Publishing. Ahmad, Zahiruddin. 1968. “New Light on the Tibet-Ladakh-Mughal War of 1679–84.” East and We­st 18, 3–4:340–361. Ahmed, Monisha. 2002. Living Fabric. Weaving among the Nomads of Ladakh Himalaya. Bangkok: Orchid Press. ———. 2005. “Textile Arts of Ladakh—Nomadic Weaves to Silk-Brocades.” In Ladakh—Culture at the Crossroads”, pp. 66–81. Eds. Monisha Ahmed and Clare Harris. Mumbai: Marg Publications. ———. 2010. “Ladakh.” In The Berg Encyclopaedia of World Dress and Fashion—South Asia. Vol. 4, pp. 185–193. Ed. Jasleen Dhamija. Oxford: Berg Publishers. Ahmed, Monisha, and Clare Harris (Eds.). 2005. Ladakh—Culture at the Crossroads. Mumbai: Marg Publications. Ahmed, Monisha, and John Bray (Eds.). 2009. Recent Research on Ladakh 2009—Papers from the 12th Colloquium of the International Association for Ladakh Studies, Kargil. Kargil and Leh: International Association for Ladakh Studies. Aitchison, J.E.T. 1874. Hand-book of the Trade Products of Leh. Calcutta: Wynne & Co. Akasoy, Anna, Charles Burnett, and Ronit Yoeli-Tlalim (Eds.). 2011. Islam and Tibet. Interactions along the Musk Routes. Farnham: Ashgate. Alafouzo, Marjo. 2008. The Iconography of the Drinking Scene in the Dukhang at Alchi, Ladakh. Ph. D. Dissertation, University of London. Alam, Humera. 1994. “The Sikri Stupa in Lahore Museum. A Fresh Study of its Reliefs.” Lahore Museum Bulletin 7:103–118. Alexander, André. 2005. The Temples of Lhasa. London: Serindia. Alexander, André, and Pimpim de Azevedo. 1998. The Old City of Lhasa—Report from a Conservation Project. Berlin: Tibet Heritage Fund. Alexander, André, and Andrew Brannan. 1997. “Inner City of Lhasa 1948–1995 and the Lhasa Historic City Archives Project. In Tibetan Studies. Proceedings of the 7th Seminar of the International Association for Tibetan Studies, Graz 1995. Vol. I, pp. 1–5. Eds. Helmut Krasser, Michael Torsten Much, Ernst Steinkellner, and Helmut Tauscher. Wien: Verlag der Österreichischen Akademie der Wissenschaften.

bibliography

401

Alexander, André, and Andreas Catanese. 2008. The Restoration of Tsas Soma Mosque, Leh, Ladakh, India. Berlin: Tibet Heritage Fund. www.tibetheritagefund.org). Alexander, André, and Sam van Schaik. 2011. “The Stone Maitreya of Leh: the Rediscovery and Recovery of an Early Tibetan Monument.” Journal of the Royal Asiatic Society Series 3, 21, 4:421–439. Anderson, Christy (Ed.). 2002. The Built surface. 1. Architecture and the Pictorial Arts from Antiquity to the Enlightenment. Aldershot: Ashgate. Archaeological Survey of India. 1910. Annual Report of the Archaeological Survey of India, Frontier Circle for 1909–10. Peshawar: Government Press, North-west Frontier Province. Ardussi, John, and Henk Blezer (Eds.). 2002. Impressions of Bhutan and Tibetan Art. Proceedings of the Ninth Seminar of the IATS 2000. Vol. 3. Leiden: Brill. Aschoff, Jürgen. 1987. Tsaparang. Tibets grosses Geheimnis. Freiburg: Eulen Verlag. Asimov, Muhammad S., and Clifford E. Bosworth (Eds.). History of Civilizations of Central Asia. Vol. IV: The Age of Achievement: A.D. 750 to the End of the Fifteenth Century. Part One: The Historical, Social and Economic Setting, pp. 61–76. Paris: UNESCO Publishing. Azarpay, Guitty. 1981. Sogdian Painting. The Pictorial Epic in Oriental Art. Berkeley: University of California Press. Bailey, Harold W. 1936. “An Itinerary in Khotanese Saka.” Acta Orientalia 14:258–267. Baipakov, Karl 2000. “The Silk Routes Across Central Asia.” In History of Civilizations of Central Asia. Vol. IV: The Age of Achievement: A.D. 750 to the End of the Fifteenth Century. Part Two: The Achievements, pp. 221–226. Eds. Clifford E. Bosworth and Muhammad S. Asimov. Paris: UNESCO Publishing. Baker, Ian A., and Thomas Laird. 2000. The Dalai Lama’s Secret Temple: Tantric Wall Paintings from Tibet. London: Thames & Hudson. Bandini-König, Ditte, Martin Bemann, and Harald Hauptmann. 1997. “Rock Art in the Upper Indus Valley.” In The Indus: Cradle and Crossroads of Civilizations, pp. 29–70. Ed. Harald Hauptmann. Islamabad: Embassy of the Federal Republic of Germany, Islamabad. Barnes, Ruth. 1997. Indian Block-Printed Textiles in Egypt. The Newberry Collection in the Ashmolean Museum, Oxford. Vols I and II. Oxford: Clarendon Press. Barrett, Douglas. 1969. “Façade of a Miniature Shrine from Kashmir.” British Museum Quarterly 34, 1–2:63–67. Barthold, Wilhelm (Vasily V. Bartold). 1992. Turkestan Down to the Mongol Invasion. New Delhi: Munshiram Manoharlal Publishers. Reprint of 1968 ed. Bartholomew, Therese T. 1995. Mongolia—The Legacy of Chinggis Khan. London: Thames and Hudson. Beckwith, Christopher I. 1987. The Tibetan Empire in Central Asia, A History of the Struggle for Great Power among Tibetans, Turks, Arabs, and Chinese during the Middle Ages. Princeton: Princeton University Press.

402

bibliography

Beek, Martijn van, Kristoffer Brix Bertelsen, and Poul Pedersen. 1999. Ladakh: Culture, History, and Development between Himalaya and Karakoram. Proceedings of the Eighth Colloquium of the International Association for Ladakh Studies held at Moesgaard, Aarhus University, 5–8 June 1997. Aarhus: Aarhus University Press—New Delhi: Sterling Publishers. Beek Martijn van, and Fernanda Pirie (Eds.). 2008. Modern Ladakh. Anthropological Perspectives on Continuity and Change. Leiden: Brill. Beer, Robert. 1999. The Encyclopedia of Tibetan Symbols and Motifs. Boston: Shambala. Behrendt, Kurt. 2004. The Buddhist Architecture of Gandhara (Handbook of Oriental Studies/Handbuch der Orientalistik. Vol. 17). Leiden: Brill. ———. 2005. “Narrative Sequences in the Buddhist Reliefs from Gandhara.” In South Asian Archaeology 2001. Vol. 2, pp. 383–392. Eds. Jean-François Jarrige and Vincent Lefèvre. Paris: Éditions Recherche sur les Civilisations. Bell, Mark S. 1890. “The Great Central Asian Trade Route from Peking to Kashgaria.” In Proceedings of the Royal Geographical Society and Monthly Records of Geography 12, 2:57–93. London: Edward Stanford. Bellezza, John V. 1999. “Northern Tibet Exploration. Archaeological Discoveries of the Changthang Circuit Expedition 1999.” Asian Art Online Journal www.asianart.com/ articles/tibarcheo/index.html#r5. ———. 2000. “Images of Lost Civilization: The Ancient Rock Art of Upper Tibet.” Asian Art Online Journal. www.asianart.com/articles/rockart/index.html. ———. 2000 “Pre-Buddhist Archaeological Sites in Northern Tibet: An Introductory Report on the Types of Monuments and Related Literary and Oral Historical Sources (Findings of the Changthang Circuit Expedition, 1999).” Kailash 19, 1–2:1–142. ———. 2004. “Metal and Stone Vestiges: Religion, Magic and Protection in the Art of Ancient Tibet.” Asian Art Online Journal. www.asianart.com/articles/vestiges/index. html. ———. 2008. Zhang Zhung: Foundations of Civilization in Tibet: a Historical and Ethnoarchaeological Study of the Monuments, Rock Art, Texts, and Oral Tradition of the Ancient Tibetan Upland. Wien: Verlag der Österreichischen Akademie der Wissenschaften. Bhattacharyya, Benoytosh. 1985. Indian Buddhist Iconography. New Delhi: Cosmo. Bianca, Stefano (Ed.). 2007. Karakoram: Hidden Treasures in the Northern Areas of Pakistan, 2nd ed. Torino: Umberto Allemandi & Co. Bielmeier, Roland et al. (Eds.). Forthcoming. Comparative Dictionary of Tibetan dialects (CDTD). University of Berne. Blezer, Henk, with the assistance of Abel Zadoks (Eds.). 2002. Religion and Secular Culture in Tibet. Tibetan Studies II. PIATS 2000: Tibetan Studies: Proceedings of the Ninth Seminar of the International Association for Tibetan Studies, Leiden 2000. Vol. 2. Leiden: Brill.

bibliography

403

Bosworth, Clifford E., and Muhammad S. Asimov (Eds.). History of Civilizations of Central Asia. Vol. IV: The Age of Achievement: A.D. 750 to the End of the Fifteenth Century. Part Two: The Achievements, pp. 221–226. Paris: UNESCO Publishing. Brauen, Martin. 1997. The Mandala, Sacred Circle in Tibetan Buddhism. London: Serindia Publications. Bray, John. 1990. “The Lapchak Mission from Ladakh to Lhasa in British Indian Foreign Policy.” The Tibet Journal 15:75–96. ———. 1991. “Ladakhi History and Indian Nationhood.” South Asia Research 11(2): 115–133. ———. 1999. “August Hermann Francke’s Letters from Ladakh 1896–1906. The Making of a Missionary Scholar.” In Studia Tibetica et Mongolica (Festschrift Manfred Taube), pp. 17–36. Eds. Helmut Eimer, Michael Hahn, Maria Schetelich, and Peter Wyzlic. Swisttal-Odendorf: Indica et Tibetica Verlag. Revised version 2008 in: Tibet Journal 33, 1:3–28. ———. (Ed.). 2005. Ladakhi Histories. Local and Regional Perspectives. Leiden— Boston: Brill. New ed. 2011. Dharamsala: Library of Tibetan Works and Archives. Bray, John, and Nawang Tsering Shakspo (Eds.). 2007. Recent Research on Ladakh 2007. Leh: Jammu & Kashmir Academy of Art, Culture & Languages—International Association for Ladakh Studies. Bray, John, and Elena de Rossi Filibeck (Eds.). 2009. Mountains, Monasteries and Mosques. Recent Research on Ladakh and the Western Himalaya. Supplement 2 to Rivista degli Studi Orientali. Nuova Serie. Vol. LXXX. Pisa—Rome: Fabrizio Serra Editore. Bray, John, and Tsering D. Gonkatsang. 2009. “Three 19th Century Documents from Tibet and the lo phyag Mission from Leh to Lhasa.” In Mountains, Monasteries and Mosques. Recent Research on Ladakh and the Western Himalaya. Supplement 2 to Rivista degli Studi Orientali. Nuova Serie. Vol. LXXX, pp. 97–116. Eds. John Bray and Elena de Rossi Filibeck. Pisa—Rome: Fabrizio Serra Editore. Brentjes, Burchard. 1999. “Rock Art in Russian Far East and in Siberia. A Bird’s Eye View over a Continent.” In Tracce: Online Rock Art Bulletin 11. www.rupestre.net/tracce/ russib.html, downloaded on 11 July 2011. Bruneau, Laurianne. 2004. Les représentations rupestres du Ladakh/Zanskar (état de Jammu et Cachemire, Inde). 2 vols. Pre-doctoral dissertation. Université de Paris 1 Panthéon-Sorbonne. ———. 2010. Le Ladakh (état de Jammu et Cachemire, Inde) de l’Âge du Bronze à l’introduction du Bouddhisme: une étude de l’art rupestre. 4 vols. Université de Paris 1 Panthéon-Sorbonne. PhD thesis. Bussagli, Mario. 1978. Central Asian Painting. Geneva: Skira. Bu ston. 1932. The History of Buddhism in India and Tibet. Ed. and Transl. Eugene Obermiller. Heidelberg: Harrassowitz. Reprint ed. 1986. Delhi: Sri Satguru Publications.

404

bibliography

Cabezón, José Ignacio. 1993. A Dose of Emptiness, an annotated translation of the sTong thun chen mo of mKhas grub dGe legs dpal bzang. Delhi: Sri Satguru Publications. Cabezón, José I., and Jackson Roger R. (Eds.). 1996. Tibetan Literature. Northfield: Snow Lion. Chandra, Lokesh. 1999. Buddhist Iconography. Delhi: Aditya Prakashan. Compact ed. Chaturvedi, Anuradha. 1989. “Leh, Ladakh.” Architecture+Design 6, 1:82–89. Clarke, Graham E. 1977. “Who Were the Dards? A Review of the Ethnographic Literature of the North-Western Himalaya.” Kailash 5, 4:323–356. Clarke, John. 2004. Jewellery of Tibet and the Himalayas. London: V&A Publications. Combe, George A. (Ed.). 1926. A Tibetan on Tibet. Being the Travels and Observations of Mr Paul Sherap (Dorje Zödba) of Tachienlu. London: T. Fisher Unwin. Cozort, Daniel. 1986. Highest Yoga Tantra. Ithaca: Snow Lion Publications. ———. 1995. The Sand Mandala of Vajrabhairava. Ithaca: Snow Lion Publications. Csoma de Kőrös, Alexander. 1834. A Grammar of the Tibetan Language. Calcutta: Baptist Mission Press. ———. 1863. “Notices on the Life of Shakya, Extracted from the Tibetan Authorities.” Asiatic Researches 20, 2:285–317. Reprint ed. 1986. Tibetan Studies. Budapest: Akadémiai Kiadó. Cuevas, Bryan J. 2003. The Hidden History of the Tibetan Book of the Dead. New York: Oxford University Press. Cunningham, Alexander. 1854. Ladak, Physical, Statistical and Historical with Notices of the Surrounding Countries. London: W.H. Allen. Reprint editions 1997 Srinagar: Gulshan Publishers and 1998 New Delhi: Asian Educational Services. Czuma, Stanislaw. 1989. “Ivory Sculpture.” In The Art and Architecture of Ancient Kashmir, pp. 57–76. Ed. Pratapaditya Pal. Bombay: Marg. Dagyab, Loden Sherap. 1977. Tibetan Religious Art. 2 vols. Wiesbaden: Otto Harrassowitz. Dainelli, Giotto. 1932. “Between Baltistan and Ladak.” In The Italian Expedition to Himalaya, Karakoram and East Turkestan 1913–1914, pp. 228–250. Ed. Filippo De Filippi. London: E. Arnold & Co. Dalai Lama, 14th. 1991. Path to Bliss. Ithaca: Snow Lion Publications. ———. 1996. L’Unione di Beatitudine e Vacuità. Ithaca: Snow Lion Publication. PomaiaPisa: Chiara Luce. Dalai Lama, 14th, and Jeffrey Hopkins. 1989. The Kalachakra Tantra. Rite of Initiation for the Stage of Generation. A Commentary on the Text of Kay-drup-ge-lek-bel-sang-bo by Tenzin Gyatso, the 14th Dalai Lama, and the text itself. London: Wisdom Publications. Dani, Ahmad H. 1982. “Prehistoric Rock Carvings at Chilas.” Studia Iranica 11:65–72. ———. 1989. History of Northern Areas of Pakistan. Islamabad: National Institute of Historical and Cultural Research. Dankoff, Robert. 1972. “Kasgari on the Tribal and Kinship Organization of the Turks.” Archivum Ottomanicum IV:23–43.

bibliography

405

———. 1975. “Kasgari on the Beliefs and Superstitions of the Turks.” Journal of the American Oriental Society 95:68–80. ———. 1983. Yusuf Khass Hacib: Wisdom of Royal Glory (Kutadgu Bilig). A TurkoIslamic Mirror for the Princes. Chicago and London: The University of Chicago Press. ———. 1992. “Qarakhanid Literature and the Beginnings of the Turko-Islamic Culture.” In Central Asian Monuments, pp. 83–91. Ed. Hasan B. Paksoy. Istanbul: The Isis Press. Dargyay, Eva K., and Ulrich Gruber. 1985. “The White and Red Rong btsan of Matho Monastery (Ladakh)”. Journal of the Tibet Society 5:55–66. Davidovich, Elena A. 1998. “The Karakhanids.” In History of Civilizations of Central Asia. Vol. IV: The Age of Achievement: A.D. 750 to the End of the Fifteenth Century. Part One: The Historical, Social and Economic Setting, pp. 119–144. Eds. Muhammad S. Asimov and Clifford E. Boswoth. Paris: UNESCO Publishing. Davidson, Ronald M. 2002. “Reframing Sahaja: Genre, Representation, Ritual and Lineage.” Journal of Indian Philosophy 1, 30:45–83. Davidson, Ronald M. 2004. Indian Esoteric Buddhism. Delhi: Motilal Barnarsidass. Reprint ed. of 2002. New York: Columbia University Press. Davies, Robert H. 1862. Report on the Trade and Resources of the Countries on the NorthWestern Boundary of British India. 2 vols. Lahore: Government Press. De Filippi, Filippo. 1924. Storia della Spedizione Scientifica Italiana nel Himàlaia, Caracorùm e Turchestàn Cinese (1913–1914). Bologna: Zanichelli. Reprint ed. 1979. ———. 1932. The Italian Expedition to Himalaya, Karakoram and East Turkestan 1913– 1914. With chapters by Giotto Dainelli and John A. Spranger. London: E. Arnold & Co. De Rossi Filibeck, Elena. 2009. “Wedding Songs from Wam le.” In Mountains, Monasteries and Mosques. Recent Research on Ladakh and the Western Himalaya. Proceedings of the 13th Colloquium of the International Association for Ladakh Studies. Supplement 2 to Rivista degli Studi Orientali, Nuova Serie. Vol. LXXX, pp. 173–207. Eds. John Bray and Elena De Rossi Filibeck. Pisa—Roma: Fabrizio Serra Editore. Deambi, B.K.K. 1996. “The Sharada Script, Origin and Development.” In Jammu, Kashmir & Ladakh. Linguistic Predicament, pp. 80–100. Eds. P.N. Pushp and K. Warikoo. New Delhi: Har Anand. Denwood, Philip. 1980. “Temple and Rock Inscriptions at Alchi.” In David L. Snellgrove and Tadeusz Skorupski. The Cultural Heritage of Ladakh. Vol. 2, part 4, pp. 117–163. Warminster: Aris and Phillips. ———. 2005. “Early Connections between Ladakh/Baltistan and Amdo/Kham.” in Ladakhi Histories, pp. 31–39. Ed. John Bray. Leiden: Brill. South Asian ed. 2011. Dharamsala: Library of Tibetan Works and Archives. ———. 2007. “The Tibetans in the Western Himalayas and Karakoram, SeventhEleventh Centuries. Rock Art and Inscriptions.” Journal of Inner Asian Art and Archaeology 2:49–58.

406

bibliography

———. 2008. “The Tibetans in the West. Part 1. Geography.” Journal of Inner Asian Art and Archaeology 3:7–21. ———. 2009. “The Tibetans in the West. Part 2.” Journal of Inner Asian Art and Archaeology 4:149–160. Denwood, Philip, and Neil F. Howard. 1990. “Inscriptions at Balukhar and Char Zampa and General Archaeological Observations on Balukhar Fort and its Environs.” In Indo-Tibetan Studies: Papers in Honour and Appreciation of Professor David L. Snellgrove’s Contribution to Indo-Tibetan Studies, pp. 81–88. Ed. Tadeusz Skorupski. Tring: Institute of Tibetan Studies. Devers, Quentin, and Martin Vernier. 2011. “An Archaeological Account of the Markha Valley, Ladakh.” Revue d’Études Tibétaines 20:61–113. DeWeese, Devin. 1994. Islamization and the Native Religion in the Golden Horde: Baba Tukles and Conversion to Islam in Historical and Epic Traditions. Pennsylvania: Pennsylvania State University Press. Didier, Hugues. 1996. Les Portugais au Tibet. Les premières relations jésuites (1642–1635). Paris: Éditions Chandeigne. Second ed. 2002. Di Mattia, Maria Laura. 1996. “A Historical Profile of Ladakhi Religious Architecture.” The Tibet Journal 21, 2:90–127. ———. 2002. “Indo–Tibetan Schools of Art and Architecture in the Western Himalaya: The Instance of Ribba in Kinnaur.” In Impressions of Bhutan and Tibetan Art. Proceedings of the Ninth Seminar of the IATS 2000. Vol. 3, pp. 91–112. Eds. John Ardussi and Henk Blezer. Leiden: Brill. ———. 2007. “The Divine Palaces of the Buddha: Architectural Frames in Western Himalayan Art.” In Discoveries in Western Tibet and the Himalayas. Essays on History, Literature, Archaeology and Art. Proceedings of the Tenth Seminar of the International Association for Tibetan Studies, Oxford, 2003. Vol. 8, pp. 55–81. Eds. Amy Heller and Giacomella Orofino. Leiden: Brill. Dodin, Thierry, and Heinz Räther (Eds.). Recent Research on Ladakh 7. Proceedings of the 7th Colloquium of the International Association for Ladakh Studies held in Bonn/ Sankt Augustin, 12–15 June 1995. Ulm: Universität Ulm. Dollfus, Pascale. 2003. “De quelques histoires de klu et de btsan.” Revue d’Études Tibétaines 2:5–39. Dotson, Brandon. 2006. Administration and Law in the Tibetan Empire (the section on Law and State and its Old Tibetan Antecedents), D.Phil. thesis. Tibetan and Himalayan Studies, Oriental Institute, Oxford University. Dye, Joseph M. 1976. “Two Fragmentary Gandharan Reliefs in the Peshawar Museum: A Study of Gandharan Representations of the Four Encounters.” Artibus Asiae 38:219–245. Elias, Ney (Ed.) and E. Denison Ross (Transl.). 1895. The Tarikh-i-Rashidi of Mirza Muhammad Haidar, Dughlát. A History of the Moghuls of Central Asia. An English Version. London: Sampson Low, Marston and Co.

bibliography

407

Eracle, Jean. 1994. Thanka Dell’Himalaya Immagini Della Saggezza, Ivrea: Priuli e Verlucca. Esin , Emel. 1969. “‘AND’ The Cup Rites in Inner-Asian and Turkish Art.” In Forschungen zur Kunst Asiens. In Memoriam Kurt Erdmann, pp. 224–59. Ed. Oktay Aslanapa and Rudolf Naumann. Istanbul: Baha Matbaasi. ———. 1970–1971. “Oldrug-Turug: The Hierarchy of Sedent Postures in Turkish Iconography.” Kunst des Orients 7, 1:1–29. Feiglstorfer, Hubert. 2011. Buddhist Sacred Architecture in the Western Himalayas: 10th to 14th century. Ph.D. Dissertation, Vienna University of Technology. ———. 2012. “Research on Earthen Roof Constructions of Buddhist Temples in the Western Himalayas.” Paper presented at the 11th International Conference on the Study and Conservation of Earthen Architecture (Terra 2012). Lima, April 23–27, 2012. Fischer, Klaus. 1958. “Gandharan Sculpture from Kunduz and Environs.” Artibus Asiae 21:231–253. Fisher, Robert E. 1989. “Buddhist Architecture.” In The Art and Architecture of Ancient Kashmir, pp. 17–28. Ed. Pratapaditya Pal. Bombay: Marg. Flood, Finbar Barry. 2005. “A Royal Drinking Scene from Alchi. Iranian Iconography in the Western Himalayas.” In Image and Meaning in Islamic Art, pp. 73–97. Ed. Robert Hillenbrand. London: Altajir Trust. Fonia, R.S. 1993. “Ladakh Corridor to Central Asia. An Investigative Report of PreHistoric Cultures.” Journal of Central Asian Studies 4:35–40. Fontein, Jan. 1979. “A Rock Sculpture of Maitreya in the Suru Valley, Ladakh.” Artibus Asiae XLI, 1:5–12. Foucaux, Philippe É. 1848. Rgya tch’er rol pa, ou Développement des jeux, contenant l’histoire du Bouddha çakya-Mouni: traduit sur la version Tibétaine du bKah hgyour, et revue sur l´original Sanscrit (Lalitavistara). Paris: Imprimerie Nationale. Francfort, Henri‑Paul, Daniel Klodzinski, and Georges Mascle. 1990. “Petroglyphes archaiques du Ladakh et du Zanskar.” Arts Asiatiques 45:5–27. ———. 1992. “Archaic Petroglyphs of Ladakh and Zanskar.” In Rock Art in the Old World. pp. 47–192. Ed. Michel Lorblanchet. Papers Presented in Symposium A of the AURA Congress, Darwin (Australia) 1988. New Delhi: Indira Gandhi National Centre for the Arts. English version of preceding article. Francke, August Hermann. 1900, 1902. “Der Frühlings- und Wintermythus der Kesarsage. Ein Beitrag zur Kenntnis der vorbuddhistischen Religion Tibets.” Suomalais-Ugrilaisen Seuran Tomituksia/Mémoires de la Société Finno-Ougrienne 15, 1:1–31; 2:1–77. ———. 1901. “The Eighteen Agus.” Indian Antiquary 30, p. 564. ———. 1902. “Notes on Rock-carvings from Lower Ladakh.” Indian Antiquary 31:98–401. ———. 1901, 1902. “The Spring Myth of the Kesar Saga.” Indian Antiquary 30:329–341; 31:32–40 and 147–157.

408

bibliography

———. 1903. “More Rock-carvings from Lower Ladakh.” Indian Antiquary 32:361–363. ———. 1905–1941. A Lower Ladakhi Version of the Kesar Saga. Calcutta: Royal Asiatic Society of Bengal. Reprint ed. 2000. Delhi: Asian Educational Services. ———. 1906, 1907a. “Paladins of the Kesar Saga.” Journal of the Asiatic Society of Bengal 2, 1:467–90; 3, 2:67–77 and 261–388. ———. 1907b. A History of Western Tibet: One of the Unknown Empires. London: S.W. Partridge and Co. ———. 1907c. “Archaeology in Western Tibet. [Part II].” Indian Antiquary 36:85–98. ———. 1909. “Ten Historical Songs from Western Tibet.” Indian Antiquary 38:57–68. ———. 1914. Antiquities of Indian Tibet. Vol. I. Personal Narrative. Calcutta: Archaeological Survey of India—Superintendent Government Printing, India. Reprint eds. 1972. New Delhi: S. Chand & Co. and 1994. New Delhi: Archaeological Survey of India. ———. 1923. Tibetische Hochzeitslieder, übersetzt nach Handschriften von Tag-ma-cig. Hagen and Darmstadt: Volkwang Verlag. ———. 1926. Antiquities of Indian Tibet. Vol. II. The Chronicles of Ladakh and Minor Chronicles. Calcutta: Archaeological Survey of India—Superintendent Government Printing. Reprint ed. 1972. New Delhi: S. Chand and Co. ———. 1977. A History of Ladakh, with a Critical Introduction by S.S. Gergan & F.M. Hassnain (a reissue of Francke’s A History of Western Tibet, 1907). New Delhi: Sterling. Fuchs, Gerhard. 1978. Geological Map of the Pin Valley: Spiti. 1:50000. Vienna: Geological Survey of Austria. ———. 1979. Geological Sections across the Pin Valley area: Spiti. 1:50000. Vienna, Geological Survey of Austria. ———. 1980. Geologic-Tectonical Map of the Himalaya. 1:2000000. Vienna: Geological Survey of Austria. ———. 1984. Geological Map of Ladakh. 1:500.000. Vienna: Geological Survey of Austria. gCod pa don grub, and bSod los. 1996. Gling ge sar sgrung gi tshig rgyan nor bu’i bang mzdod [Jewel Treasury of Poetic Expressions from the epic of Ling Gesar]. Beijing: Nationalities Publishing House. Gergan, Yoseb [Bsod nams tshe brtan yo seb dge rgan]. 1976. Ladags rgyalrabs chimed ster [Bla dvags rgyal rabs ’chi med gter] (History of Ladakh) in Tibetan. Ed. S.S. Gergan [bSod nams skyabs ldan dge rgan]. New Delhi: Sterling Publishers. Gnoli, Gherardo, and Lionello Lanciotti (Eds.). 1988. Orientalia Iosephi Tucci memoriae dicata. 3 vols. Rome: Istituto Italiano per il Medio ed Estremo Oriente. Goepper, Roger. 1990. “Clues for a Dating of the Three-storeyed (Sumtsek) in Alchi, Ladakh.” Asiatische Studien 44, 2:159–176. ———. 1993. “The ‘Great Stupa’ at Alchi.” Artibus Asiae 53, 1–2:111–143. ———. 1996a. Alchi. Ladakh’s Hidden Buddhist Sanctuary: the Sumtsek. With photographs by Jaroslav Poncar. Boston: Shambhala.

bibliography

409

———. 1996b. “Early Buddhist Architecture in Alchi.” In On the Path to Void. Buddhist Art of the Tibetan Realm, pp. 82–97. Ed. Pratapaditya Pal. Bombay: Marg. ———. 2003. “More Evidence for Dating the Sumtsek in Alchi and its Relations with Kashmir.” In Dating Tibetan Art. Essays on the Possibilites and Impossibilities of Chronology from the Lempertz Symposium, Cologne, pp. 15–24. Ed. Ingrid KreideDamani. Wiesbaden: Ludwig Reichert Verlag. Goetz, Hermann. 1969. Studies in the History of Art of Kashmir and the Indian Himalayas. Wiesbaden: Harassowitz. Golden, Peter B. 1990. “The Karakhanids and Early Islam.” In The Cambridge History of Early Inner Asia, pp. 343–370. Ed. Denis Sinor. Cambridge: Cambridge University Press. ———. 1992. An Introduction to the History of the Turkic Peoples. Wiesbaden: Otto Harrassowitz. ———. 2005. “The Turks: a Historical Overview.” In Turks: A Journey of Thousand Years 600–1600, pp. 201–246. Ed. David Roxburgh. London: Royal Academy of Arts. Gombrich, Richard F. 1971. Precept and Practice—Traditional Buddhism in the Rural Highlands of Ceylon. Oxford: Clarendon Press. Grenet, Frantz. 2004. “The Sogdians at the Crossroads of Asia.” In The Silk Road. Trade. Travel, War and Faith, pp. 110–13. Ed. Susan Whitfield. London: British Library. Grünwedel, Albert. 1901. “Bilder zur Kesarsage.” Globus 79:281–283. Grünwedel, Albert, and James Burgess. 1901. Buddhist Art in India. London: B. Qaritch. Reprint ed. 1998. New Delhi: Academic Publishers. Gutschow, Niels. 2006. “Stūpa. Einführung in Geschichte, Typologie und Symbolik”. In Tibet—Klöster öffnen ihre Schatzkammern. Ed. Jeong-Hee Lee-Kalisch. Essen and München: Kulturstiftung Ruhr and Hirmer Verlag. Gutschow, Niels, Axel Michaels, Charles Ramble, and Ernst Steinkellner (Eds.). 1998. Sacred Landscape of the Himalaya. Proceedings of an International Conference at Heidelberg 22–27 May 1998. Vienna: Verlag der Österreichischen Akademie der Wissenschaften. Guy, John. 1998. Woven Cargoes: Indian Textiles in the East. New York: Thames and Hudson. Gyurme Dorje. 2005. The Tibetan Book of the Dead. London: Penguin Books. Halkias, Georgios T. 2011. “The Muslim Queens of the Himalayas: Princess Exchanges in Baltistan and Ladakh.” In Islam and Tibet. Interactions along the Musk Routes, pp. 231–252. Eds. Anna Akasoy, Charles Burnett, and Ronit Yoeli-Tlalim Farnham: Ashgate. Ham, Peter van. 2010. Heavenly Himalayas: The Murals of Mangyu and Other Discoveries in Ladakh. London: Prestel. Hamid, Abdul. 1998. Ladakhi-English-Urdu Dictionary with an English-Ladakhi index. Leh: Melong Publications.

410

bibliography

Harrison, John. 1996. Himalayan Buildings. Recording Vernacular Architecture. Mustang and the Kalash. Kathmandu: British Council and Goethe Institut. Reprint of 1995 ed. Islamabad: British Council. Hauptmann, Harald. 2003. Rock Carvings and Inscriptions along the Upper Indus, Unpublished presentation at the International Association for Ladakh Studies Colloqium at Leh, 2003. ———. 2005. “Pre-Islamic Heritage in the Northern Areas of Pakistan.” In Karakoram: Hidden Treasures in the Northern areas of Pakistan, pp. 1–40. Ed. Stefano Bianca. Torino: The Aga Khan Trust for Culture. ———. 2007. “Pre-Islamic Heritage in the Northern Areas of Pakistan.” In Karakoram: Hidden Treasures in the Northern Areas of Pakistan, pp. 21–39. Ed. Stefano Bianca. 2nd ed. Torino: Umberto Allemandi & Co. Heller, Amy. 1999. Tibetan Art—Tracing the Development of Spiritual Ideals and Art in Tibet 600–2000 AD. Italian transl. Arte Tibetana. Lo sviluppo della spiritualità e dell’arte in Tibet dal 600 al 2000 d C. Milan: Jaca Book. ———. 2006. “Rezeption und Adaption fremder ästhetischer Elemente in der Tibetischen Skulptur—eine Spurensuche.” In Tibet—Klöster öffnen ihre Schatzkammern, pp. 80–89. Ed. Jeong-Hee Lee-Kalisch. Essen and München: Kulturstiftung Ruhr and Hirmer Verlag. ———. 2009. Hidden Treasures of the Himalayas. Tibetan Manuscripts, Paintings and Sculptures of Dolpo. Chicago: Serindia Publications. ———. 2011. “Preliminary Remarks on the Donor Inscriptions and Iconography of an 11th-century Mchod rten at Tholing.” In Tibetan Art and Architecture in Context, Proceedings of the 11th Seminar of the International Association for Tibetan Studies, pp. 43–74. Eds. Erberto Lo Bue and Christian Luczanits. Andiast: International Institute for Tibetan and Buddhist Studies GmbH. ———. Forthcoming. “The Impact of Kashmiri Art on the Sculpture of mNga’ ris skor gsum, 11th to 13th Centuries: a Re-assessment of the Chronology of the Alchi Sumstek as a 13th Century Foundation”. Journal of Inner Asian Art and Archaeology. Heller, Amy, and Giacomella Orofino (Eds.). 2007. Discoveries in Western Tibet and the Himalayas. Essays on History, Literature, Archaeology and Art. Proceedings of the Tenth Seminar of the International Association for Tibetan Studies, Oxford, 2003. Vol. 8. Leiden: Brill. Herdick, Reinhard. 1997. “The Old Village Centre of Lamayuru in West Ladakh.” In Recent Research on Ladakh 7. Proceedings of the 7th Colloquium of the International Association for Ladakh Studies held in Bonn/Sankt Augustin, 12–15 June 1995, pp. 195– 221. Eds. Thierry Dodin and Heinz Räther. Ulm: Universität Ulm. Herrmann, Silke. 1991. Kesar-Versionen aus Ladakh. Wiesbaden: Harrassowitz. Hillenbrand, Robert. (Ed.). 1994. The Art of the Saljuqs in Iran and Anatolia. California: Mazda Publishers.

bibliography

411

Himalayan Cultural Heritage Foundation. www.heritagehimalaya.org. Horlemann, Bianca. 2007. “The Relations of the Eleventh-century Tsong kha Tribal Confederation to its Neighbour States on the Silk Road.” In Contributions to the Cultural History of Early Tibet, pp. 79–101. Ed. Matthew T. Kapstein and Brandon Dotson. Leiden—Boston: Brill. Howard, Kath. 1995. “Archaeological Notes on mChod-rten types in Ladakh and Zanskar from the 11th–15th Centuries.” In Recent Research on Ladakh 4 and 5. Proceedings of the Fourth and Fifth International Colloquia on Ladakh, pp. 61–78. Eds. Henry Osmaston and Philip Denwood. London: School of Oriental and Aftican Studies. Delhi: Motilal Banarsidass. Howard, Neil. 1989. “The Development of the Fortresses of Ladakh c. 950 to c. 1650 A.D.” East and West 39, 1–4:217–288. ———. 1995a. “The Fortified Places of Zanskar.” In Recent Research on Ladakh 4 and 5. Proceedings of the Fourth and Fifth International Colloquia on Ladakh, pp. 79–99. Eds. Henry Osmaston and Philip Denwood. Delhi: Motilal Banarsidass. ———. 1995b. “An Introduction to the Fortifications of Central Nepal.” European Bulletin of Himalayan Research 9:20–31. ———. 1999. “Ancient Painted Pottery from Ladakh.” In Ladakh: Culture, History, and Development between Himalaya and Karakoram. Proceedings of the Eighth Colloquium of the International Association for Ladakh Studies held at Moesgaard, Aarhus University, 5–8 June 1997, pp. 222–236. Eds. Martijn van Beek, Kristoffer Brix Bertelsen, and Poul Pedersen. Aarhus: Aarhus University Press. New Delhi: Sterling Publishers. ———. 2005. “Sultan Zain-ul Abidin’s Raid into Ladakh.” In Ladakhi Histories. Local and Regional Perspectives, pp. 125–146. Ed. John Bray. Leiden: Brill. South Asian reprint ed. 2011. Dharamsala: Library of Tibetan Works and Archives. Huber, Brigitte, Marianne Volkart, and Paul Widmer (Eds.). 2008. Chomolangma, Demawend und Kasbek. Festschrift für Roland Bielmeier zum 65. Geburtstag. Vol. 1. Beiträge zur Zentralasienforschung 12. Halle: International Institute for Tibetan and Buddhist Studies. Hughes, Richard. 2007. “Vernacular Architecture and Construction Techniques in the Karakoram.” In Karakoram: Hidden Treasures in the Northern Areas of Pakistan, pp. 96–132. Ed. Stefano Bianca. 2nd ed. Torino: Umberto Allemandi & Co. Huntington, John C., and Dina Bangdel. 2003. The Circle of Bliss—Buddhist Meditational Art. Columbus, Ohio: Museum of Art. Chicago: Serindia Publications. Huntington, Susan. 1990. Leaves from the Bodhi Tree. The Art of Pala India (8th–12th Centuries) and its International Legacy. Dayton: Dayton Art Institute in association with the University of Washington Press. Huntington, Susan L., and John C. Huntington. 1993. The Art of Ancient India. New York and Tokyo: Weatherhill Inc.

412

bibliography

Hutchison, J., and J. Ph. Vogel. 1933. History of the Punjab Hill States. Lahore: Punjab Government Printers. Icke-Schwalbe, Lydia. 1990. “The Ladakh collection of Herrnhut in Historical and Cultural Valuation.” In Wissenschaftsgeschichte und gegenwärtige Forschungen in Nordwest-Indien, pp. 36–52. Eds. Gudrun Meier and Lydia Icke-Schwalbe. Dresden: Museum für Völkerkunde. Ihara, Shoren, and Zuiho Yamaguchi (Eds.). Tibetan Studies. Proceedings of the 5th Seminar of the International Association for Tibetan Studies. Vol. 2. Narita: Naritasan Shinshoji. Jackson, David. 1996. A History of Tibetan Painting. Wien: Verlag Der Österreichischen Akademie Der Wissenschaften. ———. 2005. “Lineages and Structure in Tibetan Buddhist Painting: Principles and Practice Of an Ancient Sacred Choreography.” Journal of the International Association of Tibetan Studies 1:1–40. Jackson, David and Janice. 1998. Tibetan Thangka Painting—Methods & Materials. London: Serindia Publications. Jackson, Peter. 1990. The Mission of Friar William of Rubruck. His Journey to the Court of the Great Khan Mongke 1254–1255. London: The Hakluyt Society. Jackson, Roger R. 2004. Tantric Treasures. New York: Oxford University Press. Jamspal, Lozang. 1997. “The Five Royal Patrons and Three Maitreya Images in Basgo.” In Recent Reserch on Ladakh 6, Proceedings of the Sixth International Colloquium on Ladakh, Leh 1993, pp. 139–156. Eds. Henry Osmaston and Nawang Tsering. Bristol: University of Bristol Press. Delhi: Motilal Banarsidass. Jest, Corneille, and John Sanday. 1983. “The Palace of Leh in Ladakh: an Example of Himalayan Architecture in Need of Conservation.” Mountain Research and Development 3, 1:1–11. Jettmar, Karl. 1990. “Das westtibetische Zentrum der Kesarsage. Zur Rechtfertigung der These A.H. Franckes.” In Wissenschaftsgeschichte und gegenwärtige Forschungen in Nordwest-Indien, pp. 259–265. Eds. Gudrun Meier and Lydia Icke-Schwalbe. Dresden: Museum für Völkerkunde. ———. 2002. Beyond the Gorges of Indus, Archaeology before Excavation. Karachi: Oxford University Press. Jettmar, Karl, and Volker Thewalt. 1985. Zwischen Gandhara und den Seidenstrassen, Felsbilder am Karakorum Highway. Entdeckungen deutsch-pakistanischer Εxpedi­ tionen 1979–1984. Mainz: Philipp von Zabern. ———. 1987. Between Gandhara and the Silk Roads. Rock Carvings along the Karakorum Highway. Mainz: Philipp von Zabern. Jha, D.N. 2000. Ancient India in Historical Outline. New Delhi: Manohar. Jiang, Cheng’an, and Wenlei Zheng (Eds.) 2000. Precious Deposits. Historical Relics of Tibet, China. 5 vols. Beijing: Morning Glory Publishers.

bibliography

413

Kak, Ram Chandra. 1923. Handbook of the Archaeological and Numismatic Sections of the Sri Pratap Singh Museum, Srinagar. Calcutta: Thacker, Spink & Co. Reprint ed. 1985. Patna: Eastern Book House. ———. 1933. Ancient Monuments of Kashmir. London: India Society Reprint ed. 2000. New Delhi: Aryan Books. Kantowsky, Detlef, and Reinhard Sander (Eds.). 1983. Recent Research on Ladakh. Munich: Weltforum Verlag. Kaplanian, Patrick. 1981. Les Ladakhi du Cachemire. Paris: Hachette. ———. 1995. “L’homme dans le monde surnaturel au Ladakh.” In Recent Research on Ladakh 4&5, pp. 101–108. Ed. Henry Osmaston and Philip Denwood. London: School of Oriental and African Studies/New Delhi: Motilal Benarsidass. Karev, Yury. 2005. “Qarakhanid Wall Paintings in the Citadel of Samarqand: First Report and Preliminary Observations.” Muqarnas 22:45–84. Karma gling pa (Ed). 1992. Zab chos zhi khro dgongs pa rang grol las bar do thos grol gyi skor ba bzhugs so. Dharamsala: Tibetan Cultural Printing Press. Karmay, Heather. 1977. “Tibetan Costumes, 7th to 11th Centuries.” In Essais sur l’art du Tibet, pp. 64–81. Eds. Ariane MacDonald and Yoshiro Imaeda. Paris: Librairie d’Amérique et d’Orient. Kennion, Roger L. 1910. Sport and Life in the Further Himalaya. Edinburgh and London: Blackwood. Khan, Asfandyar. 2009. “The Origin of the Kesar Epic.” In Recent Research on Ladakh 2009, pp. 13–19. Eds. Monisha Ahmed and John Bray. Kargil and Leh: International Association for Ladakh Studies. Khan, Hashmatullah. 1939. Tarikh Jammun, Kashmir, Laddakh aur Baltistan. Lucknow: Noor Alimad Malik and Mohammed Tegh Bahadur. Khan, Kacho Sikandar. 1987. Qadim Laddakh tarikh va tamaddun. Yokma Kharbu: published by the author. Khosa, Sunil. 1998. Handbook of Indian Art. New Delhi: Sundeep Prakashan. Khosla, Romi. 1979. Buddhist Monasteries in the Western Himalayas. Kathmandu: Ratna Pustak Bhandar. Kitzinger, Ernst. 1946. “The Horse and the Lion Tapestry at Dumbarton Oaks.” Dumbarton Oaks Papers 3:1–107. Klimburg, Max. 2007. “Traditional Art and Architecture in Baltistan”. In Karakoram: Hidden Treasures in the Northern Areas of Pakistan, pp. 149–164. Ed. Stefano Bianca. 2nd ed. Torino: Umberto Allemandi & Co. Klimburg-Salter, Deborah E. 1988a. “Reform and Renaissance: a Study of Indo-Tibetan Monasteries in the 11th Century.” In Orientalia Iosephi Tucci memoriae dicata, pp. 683–702. Eds. Gherardo Gnoli and Lionello Lanciotti. Rome: Istituto Italiano per il Medio ed Estremo Oriente.

414

bibliography

———. 1988b. “The Life of the Buddha in Western Himalayan Monastic Art. Act One.” East and West 38, 1–3:189–214. ———. 1995. Buddha in Indien. Die frühindische Skulptur von König Ashoka bis zur Guptazeit. Wien: Kunsthistorisches Museum. ———. 1997. Tabo, a Lamp for the Kingdom. With contributions by Christian Luczanits, Luciano Petech, Ernst Steinkellner, and Erna Wandl. Milano: Skira. ———. 2002. “Ribba—the Story of an Early Buddhist Temple in Kinnaur.” In Buddhist Art and Tibetan Patronage Ninth to Fourteenth Centuries. Proceedings of the Ninth Seminar of the International Association for Tibetan Studies, Leiden 2000. Vol. 7, pp. 1–28. Eds. Deborah Klimburg–Salter and Eva Allinger. Leiden: Brill. Klimburg–Salter, Deborah, and Eva Allinger (Eds.). 2002. Buddhist Art and Tibetan Patronage Ninth to Fourteenth Centuries. Proceedings of the Ninth Seminar of the International Association for Tibetan Studies, Leiden 2000. Vol. 7. Leiden: Brill. Klimburg-Salter, Deborah, Kurt Tropper, and Christian Jahoda (Eds.). 2007. Text, Image and Song in Transdisciplinary Dialogue. Proceedings of the 10th Seminar of the International Association for Tibetan Studies, Oxford 2003. Leiden—Boston: Brill. Kossak, Steven M., and Jane Casey Singer. 1998. Sacred Visions. Early Paintings from Central Tibet. New York: The Metropolitan Museum of Art. Kotchnev, Boris D. 2001. “Les frontières du royaume des karakhanides.” In Études karakhanides, pp. 41–48. Tachkent and Aix en Provence: Édisud. Kozicz, Gerald. 2006. “Gya—100 Jahre nach A. H. Francke: eine Dokumentation der architektonischen und archäologischen Überreste eines buddhistisch geprägten Siedlungsraums im westlichen Himalaya.” Archiv für Völkerkunde 56:51–70. ———. 2007a. “The Architecture of the Empty Shells of Nyar ma.” In Discoveries in Western Tibet and the Himalayas. Essays on History, Literature, Archaeology and Art. Proceedings of the Tenth Seminar of the International Association for Tibetan Studies, Oxford, 2003. Vol. 8, pp. 41–54. Eds. Amy Heller and Giacomella Orofino, Leiden— Boston: Brill. ———. 2007b. “Documenting the Last Surviving Murals of Nyarma.” Orientations 38, 4:60–64. ———. 2010a. “The Architectural Matrix of the Alchi Sumtsek”. Indo-Asiatische Zeitschrift 14:31–41. ———. 2010b. Stupa Project. http://stupa.arch-research.at/cms/. Accessed on 21-10-2011. Kreide-Damani, Ingrid (Ed.). 2003. Dating Tibetan Art. Essays on the Possibilites and Impossibilities of Chronology from the Lempertz Symposium, Cologne. Wiesbaden: Ludwig Reichert Verlag. Kreijger, Hugo E. 2001. Tibetan Painting—The Jucker Collection. London: Serindia Publications. Krist, Gabriela, and Tatjana Bayerová (Eds.). 2010. Heritage Conservation and Research in India. Wien—Weimar: Böhlau.

bibliography

415

Krom, Nicolaas J. 1925. The Life of the Buddha on the Stupa of Barabudur according to the Lalitavistara–Text. The Hague: Martinus Nijhoff. Reprint ed. 1974 Varanasi: Bhartiya Publishing. Kumar, Ritu. 1999. Costumes and Textiles of Royal India. London: Christie’s Books. Lee-Kalisch, Jeong-Hee (Ed.). 2006. Tibet—Klöster öffnen ihre Schatzkammern. Essen and München: Kulturstiftung Ruhr and Hirmer Verlag. Lessing, Ferdinand D., and Alex Wayman. 1998. Introduction to the Buddhist Tantric System. Delhi: Motilal Barnarsidass. Lewis, Geoffrey (Transl.). 1974. The Book of Dede Korkut. London: Penguin Books. Ligeti, Louis (Ed.). 1978. Proceedings of the Csoma de Körös Memorial Symposium, Bibliotheca Orientalis Hungarica XXIII. Budapest: Akadémiai Kiadò. Linrothe, Robert. 1994. “The Murals of Mangyu: A Distillation of Mature Esoteric Buddhist Iconography.” Orientations 25, 11:92–102. ———. 1997. “Paving over Precious Heritage,” Ladags Melong 2, 1:20–23. ———. 1999. Ruthless Compassion. London: Serindia. ———. 2001. “Creativity, Freedom and Control in the Contemporary Renaissance of Reb gong Painting.” The Tibet Journal 26, 3–4:5–90. ———. 2002. “inVISIBLE: Picturing interiority in western Himalayan stūpa architecture.” In The Built Surface. 1. Architecture and the Pictorial Arts from Antiquity to the Enlightenment, pp. 75–98. Ed. Christy Anderson. Aldershot: Ashgate. ———. 2006. “Two Fieldnotes from Zangskar: A Kashmiri Sculpture in a Personal Shrine and The Etymology of ‘Kankani’ Chorten.” In Mei shou wan nian—Long Life Without End. Festschrift in Honor of Roger Goepper, pp. 167–180. Eds. Jeong-Hee LeeKalisch, Antje Papist-Matsuo, and Willibald Veit. Frankfurt a. M.: Peter Lang. ———. 2007a. “A Winter in the Field.” Orientations 38, 4:40–53. ———. 2007b. “Strengthening the Roots: An Indian Yogi in Early Drigung Paintings of Ladakh and Zangskar.” Orientations 38, 4:65–71. Linrothe, Robert, and Melissa R. Kerin. 2001. “Deconsecration and Discovery: The Art of Karsha’s Kadampa Chorten Revealed.” Orientations, 32, 10:52–63. Litvinsky, Boris A., Guang-da Zhong, and R. Shabani Samghabadi (Eds.). 1996. History of Civilizations of Central Asia. Vol. 3. The Crossroads of Civilizations: A.D. 250 to 750, Paris: UNESCO Publishing. Lo Bue, Erberto. 1983. “Traditional Tibetan Painting in Ladakh in the 20th Century.” International Folklore Review 3:52–72. ———. 1990. “Iconographic Sources and Iconometric Literature in Tibetan and Himalayan Art.” In Indo-Tibetan Studies, pp. 171–197. Ed. Tadeusz Skorupski. Tring: The Institute of Buddhist Studies. ———. 2005. “Lives and Works of Traditional Buddhist Artists in 20th Century Ladakh. A Preliminary Account”. In Ladakhi Histories. Local and Regional Perspectives, pp. 353–378. Ed. John Bray. Leiden—Boston: Brill. New ed. 2011. Dharamsala: Library of Tibetan Works and Archives.

416

bibliography

———. 2005. “Yama’s Judgement in the Bar do thos grol chen mo: An Indic Mystery Play in Tibet.” The Tibet Journal XXX, 2:9–24. ———. 2007. “The Gu ru lha khang at Phyi dbang: a Mid-15th Century Temple in Central Ladakh.” In Discoveries in Western Tibet and the Himalayas. Essays on History, Literature, Archaeology and Art. Proceedings of the Tenth Seminar of the International Association for Tibetan Studies, Oxford, 2003. Vol. 8, pp. 175–196. Eds. Amy Heller and Giacomella Orofino. Leiden—Boston: Brill. Lo Bue, Erberto, and Franco Ricca. 1990. Gyantse Revisited. Firenze: Le Lettere. Lo Bue Erberto, and Christian Luczanits (Eds.) 2011. Tibetan Art and Architecture in Context, Proceedings of the 11th Seminar of the International Association for Tibetan Studies. Andiast: International Institute for Tibetan and Buddhist Studies GmbH. Lo Bue, Erberto, and Chiara Bellini. (Forthcoming). Arte del Ladak. Tesori di arte buddhista nel Tibet indiano dall’XI al XXI secolo. Milano: Jaca Book. Lorblanchet, Michel (Ed.). 1992. Rock Art in the Old World. New Delhi: Indira Gandhi National Centre for the Arts. Luczanits, Christian. 1996. “Early Buddhist Wood Carvings from Himachal Pradesh.” Orientations 2, 8:67–75. ———. 1998. “On an Unusual Painting Style in Ladakh.” In The Inner Asian International Style 12th–14th Centuries, Papers Presented at Panel of the 7th Seminar of the International Association for Tibetan Studies, Graz 1995, pp. 151–169. Eds. Deborah Klimburg-Salter and Eva Allinger. Wien: Verlag der Österreichischen Akademie der Wissenschaften. ———. 1999. “The Life of the Buddha in the Sumtsek.” Orientations 30, 1:30–39. ———. 2002. “The Wanla Bkra shis gsum brtsegs.” In Buddhist Art and Tibetan Patronage Ninth to Fourteenth Centuries. Proceedings of the Ninth Seminar of the International Association for Tibetan Studies, Leiden 2000. Vol. 7, pp. 115–125. Eds. Deborah Klimburg–Salter and Eva Allinger. Leiden: Brill. ———. 2003. “Art-historical Aspects of dating Tibetan art.” In Dating Tibetan Art, pp. 25–57. Ed. Ingrid Kreide-Damani. Wiesbaden: Ludwig Reichert Verlag. ———. 2004. Buddhist Sculpture in Clay. Early Western Himalayan Art, late 10th to early 13th centuries. Chicago: Serindia Publications. ———. 2005. “The Early Buddhist Heritage of Ladakh Reconsidered.” In Ladakhi Histories. Local and Regional Perspectives, pp. 65–96. Ed. John Bray. Leiden—Boston: Brill. New ed. 2011. Dharamsala: Library of Tibetan Works and Archives. ———. 2008. Gandhara: Das buddhistische Erbe Pakistans. Legenden, Klöster und Paradiese. Bonn: Kunst– und Ausstellungshalle der BRD. Luczanits, Christian, and Holger Neuwirth. 2010. “The Development of the Alchi Temple Complex. An Interdisciplinary Approach.” In Heritage Conservation and Research in India. 60 years of Indo-Austrian collaboration, pp. 79–84. Eds. Gabriela Krist and Tatjana Bayerová. Wien—Weimar: Böhlau.

bibliography

417

Macdonald, Alexander W., and Pema Tsering. 1979. “A Note on Five Tibetan Thangkas of the Ge-sar Epic.” In Die Mongolischen Epen, pp. 150–157. Ed.Walther Heissig. Wiesbaden, Harrassowitz. Malyon, Tim. 1986. “The Oldest Wooden Statue in India.” In The Arts of Kashmir, pp. 59–62. Ed. Pratapaditya Pal. Mumbai: Marg. Mani, B.R. 1998–1999. “Rock Carvings and Engravings in Ladakh: New Discoveries.” Pragdhara 9: 65–74. Marx, Karl. 1891, 1894, 1902. “Three Documents Relating to the History of Ladakh: Tibetan Text, Translation and Notes.” Journal of the Asiatic Society of Bengal 60, 3:97–134; 63, 2:94–107; 71, 1: 21–34. ———. 1897. “Eine ärtztliche Missionsreise im Jahre 1890.” In Eben‑Ezer, Gedenkbuch der Familie Julius Marx: 55–99. Ed. Gustaf Dalman. Leipzig: privately published. Meier, Gudrun, and Lydia Icke-Schwalbe (Eds.). 1990. Wissenschaftsgeschichte und gegenwärtige Forschungen in Nordwest-Indien. Dresden: Museum für Völkerkunde. Minorsky, Vladimir. 1982. Hudud al-Alam: The Regions of the World. Cambridge: E.J.W. Gibb Memorial Trust. First ed. 1937. London: Messrs Luzac & Co. Moorcroft, William and George Trebeck. 1841. Travels in the Himalayan Provinces of Hindustan and the Panjab, in Ladakh and Kashmir, in Peshawar, Kabul, Kunduz and Bokhara . . . from 1819–1825. 2 vols. Ed. H.H. Wilson. London: John Murray. Reprint ed. 1986. New Delhi: Nirmal Publishers and Distributors. Mori, Fabrizio. 1992. “Rock Art and Cultural Evolution in the Prehistory of the Sahara.” In Rock Art in the Old World, pp. 3–8. Ed. Michel Lorblanchet. New Delhi: Indira Gandhi National Centre for the Arts. Namgyal, Phuntsok (Ed.). 2001. Ntho-ling Monastery. Beijing: China Publishing House. Namgyal Institute for Research on Ladakhi Art and Culture (NIRLAC). 2008a. Legacy of a Mountain People. Inventory of Cultural Resources of Ladakh. Vol. 1: Leh, Khaltse. New Delhi: NIRLAC. ———. 2008b. Legacy of a Mountain People. Inventory of Cultural Resources of Ladakh. Vol. 2: Leh, Leh – Kharu. New Delhi: NIRLAC. ———. 2008c. Legacy of a Mountain People. Inventory of Cultural Resources of Ladakh. Vol. 3: Leh, Durbuk – Nyoma. New Delhi: NIRLAC. Nawab, Sarabhai M. 1980–1985. Jain Paintings. Vol. I. Ahmedabad: Messrs Sarabhai Manilal Nawab. Nebesky-Wojkowitz, René de. 1996. Oracles and Demons of Tibet. The Cult and Iconography of the Tibetan Protective Deities. Delhi: Book Faith India. Original ed. 1956. The Hague: Mouton. Neumann, Helmut F., and Heidi A. Neumann. 2008. “The Portal of Khojar Monastery in West Tibet.” Orientations 39, 6:62–73. Neuwirth, Holger. 2009. “Frühe Buddhistische Architektur im westlichen Himalaya.” Institut für Architekturtheorie, Kunst– und Kulturwissenschaften, Technische

418

bibliography

Universität Graz. www.bks.tugraz.at/~neuwirth/neuweb/fwf_fsp_netz/start.htm (accessed Jan 2010). Norberg-Hodge, Helena, and Gyelong Thupstan Paldan. 1991. Ladakhi-English EnglishLakakhi Dictionary. Delhi: LEDeG & Ladakh Project. Odak, Osaga. “Kenya Rock Art Studies and the Need for a Discipline.” In Rock Art in the Old World, pp. 33–47. Ed. Michel Lorblanchet. New Delhi: Indira Gandhi National Centre for the Arts. Orofino, Giacomella. 1990. “A Note on Some Tibetan Petroglyphs of the Ladakh Area.” East and West 40, 1–4:173–201. Osmaston, Henry, and Nawang Tsering (Eds.). 1997. “The Five Royal Patrons and Three Maitreya Images in Basgo”. In Recent Reserch on Ladakh 6, Proceedings of the Sixth International Colloquium on Ladakh, Leh 1993. Bristol: University of Bristol Press. Delhi: Motilal Banarsidass. Osmaston, Henry, and Philip Denwood (Eds.). 1995. Recent Research on Ladakh 4 and 5. Proceedings of the Fourth and Fifth International Colloquia on Ladakh. London: School of Oriental and African Studies. Delhi: Motilal Banarsidass. Otavsky, Karel. 1998. “On the Classification of the Textiles in Terms of Art History.” (Translation from the German original “Zur kunsthistorischen Einordnung der Stoffe”, pp. 119–214). In Entlang der Seidenstrasse: Frühmittelalterliche Kunst zwischen Persien und China in der Abegg-Stiftung, pp. 1–52. Eds. Karel Otavsky with the collaboration of Sheila S. Blair. Riggisberg: Abegg-Stiftung. Otavsky, Karel, and Sheila S. Blair (Eds.). 1998. Entlang der Seidenstrasse: Frühmittelalterliche Kunst zwischen Persien und China in der Abegg-Stiftung. Riggisberg: Abegg-Stiftung. Pabongka Rinpoche. 1997. Liberation in the Palm of Your Hand. Boston: Wisdom Publications. Pal, Pratapaditya. 1975. Bronzes of Kashmir. Graz: Akademische Druck– u. Verlagsanstalt. New York: Hacker Art Books. ——— (Ed.). 1986. The Arts of Kashmir. Mumbai: Marg. ———. 1989. “Kashmir and the Tibetan Connection.” In The Art and Architecture of Ancient Kashmir, pp. 117–135. Ed. Pratapaditya Pal. Bombay: Marg. ———. (Author and Ed.). 1989. The Art and Architecture of Ancient Kashmir. Bombay: Marg. ———. 1990. Art of Tibet. Los Angeles: Los Angeles County Museum of Art and Mapin Publishing. ———. 1993. Indian Painting. Vol. I. 1000–1700. A Catalogue of the Los Angeles County Museum of Art Collection. Los Angeles: Los Angeles County Museum of Art. ———. (Ed.). 1996. On the Path to Void. Buddhist Art of the Tibetan Realm. Bombay: Marg. Pal, Pratapaditya with contributions by Amy Heller, Oskar von Hinüber, and Gautama V. Vajracharya. 2003. Himalayas. An Aesthetic Adventure. Chicago, Berkeley

bibliography

419

and Ahmedabad: The Art Institute of Chicago, University of California Press and Mapin Publishing. Pal, Pratapaditya, and Lionel Fournier. 1982. A Buddhist Paradise. The Murals of Alchi, Western Himalayas. Hong Kong: Ravi Kumar—Visual Dharma Publications. Pallis, Marco. 1995. Peaks and Lamas. Delhi: Book Faith India. Reprint ed. Panchen Lama, 1st. 2003. The Guru Puja. Delhi: L.T.W.A. Publications. Panglung, Jampa. 1983. “Die Überreste des Klosters Ñar ma in Ladakh”. In Contributions on Tibetan Language, History and Culture. Proceedings of the Csoma de Körös Symposium Held at Velm-Vienna, Austria, 13–19 September 1981. Vol. 1, pp. 281–287 and pls. VIII–XV. Eds. Ernst Steinkellner and Helmut Tauscher. Wien: Arbeitskreis für Tibetische und Buddhistische Studien Universität Wien. Papa-Kalantari, Christiane. 2007. “The Art of the Court: Some Remarks on the Historical Stratigraphy of Eastern Iranian Elements in Early Buddhist Painting of Alchi, Ladakh.” In Text, Image and Song in Transdisciplinary Dialogue. Proceedings of the 10th Seminar of the International Association for Tibetan Studies, Oxford 2003, pp. 167–228. Eds. Deborah Klimburg-Salter, Kurt Tropper, and Christian Jahoda. Leiden—Boston: Brill. Parpola, Asko, and Petteri Koskikallio (Eds.). 1995. South Asian Archaeology 1993. Helsinki: Suomalainen Tiedeakatemia. Paul, Jürgen. 2001. “Nouvelles pistes pour la recherche sur l’histoire de l’Asie centrale à l’époque karakhanide (Xe–début XIIIe siècle).” In Cahiers d’Asie centrale: 9 Études karakhanides: 13–34 Tachkent: IFEAC; Aix-en-Provence: Edisud. ———. 2002. “The Karakhanids.” In The Turks 2. Middle Ages, pp. 71–78. Ed. Hasan Celal Güzel, Cem Oguz, and Osman Karatay. Ankara: Yeni Turkiye Publications. Pedersen, Poul. 2005. “Prince Peter, Polyandry and Psychoanalysis.” In Ladakhi Histories. Local and Regional Perspectives, pp. 293–308. Ed. John Bray, Leiden—Boston: Brill. New ed. 2011. Dharamsala: Library of Tibetan Works and Archives. Pedersen, Poul, and Neil Howard. 2009. “Prince Peter’s Journey from Manali to Ladakh, 5th June–22nd August 1938.” In Recent Research on Ladakh 2009, pp. 55–71. Eds. Monisha Ahmed and John Bray. Kargil and Leh: International Association for Ladakh Studies. Pema Dorjee. 1996. Stūpa and its Technology A Tibeto-Buddhist Perspective. Delhi: Motilal Banarsidass. Petech, Luciano. 1939. A Study on the Chronicles of Ladakh (Indian Tibet). Supplement to Indian Historical Quarterly 15. London. Reprint ed. 1999. New Delhi: Low Price Publications/DK Publishers. ———. 1977. The Kingdom of Ladakh: c. 950–1842 A.D. Roma: Istituto Italiano per il Medio ed Estremo Oriente. ———. 1978. “The ’Bri-guṅ-pa Sect in Western Tibet and Ladakh”, pp. 313–325. In Proceedings of the Csoma de Körös Memorial Symposium, Bibliotheca Orientalis Hungarica XXIII. Ed. Louis Ligeti. Budapest: Akadémiai Kiadò.

420

bibliography

———. 1997a. “Western Tibet: Historical Introduction.” In Tabo, a Lamp for the Kingdom, pp. 229–255. Author and Ed. Deborah E. Klimburg-Salter, with contributions by Christian Luczanits, Luciano Petech, Ernst Steinkeller, and Erna Wandl. Milano: Skira. ———. 1997b. “A Regional Chronicle of Gu ge pu hrang.” (Review article). The Tibet Journal 22, 3:106–111. Petech, Luciano, and Christian Luczanits. 1999. Inscriptions from the Tabo Main Temple. Texts and Translation. Roma: Istituto Italiano per l’Africa e l’Oriente. Phuntsog Dorjay. 2005. Development of Buddhist Art in Ladakh from 800 AD to 1200 AD. Ph.D thesis submitted to the Postgraduate Department of History, University of Jammu. Jammu. Phylactou, Maria. 1989. Household Organization and Marriage in Ladakh, Indian Himalaya, Ph.D. thesis. London School of Economics and Political Science, University of London. Pieper, Jan. 1980. “Stūpa Architecture of the Upper Indus Valley” In The Stūpa. Its Religious, Historical and Architectural Significance, pp. 127–136. Eds. Anna L. Dallapiccola and Stephanie Zingel-Avé-Lallemant. Wiesbaden: Franz Steiner Verlag. Poell, Heinrich. 2004. “Wooden Temple Doors in Ladakh, 12th–14th centuries CE.” Journal of the Asiatic Society of Mumbai 79:191–204. ———. 2011. “The Evolution of the Ashta–Maha–Pratiharya Iconography and its Impact on later Buddhist Art.” Journal of Bengal Art 16:9–37. Ponlop, Dzogchen. 2008. Mind Beyond Death. Ithaca: Snow Lion. Pushp, P.N., and K. Warikoo (Eds.). 1996. Jammu, Kashmir & Ladakh. Linguistic Predicament. New Delhi: Har Anand. Radhu, Abdul Wahid. 1997. Islam in Tibet and Tibetan Caravans. Ed. Gray Henry Louisville: Fons Vitae. Ray, Sunil C. 1970. Early History and Culture of Kashmir. Delhi: Munshiram Manoharlal. Reedy, Chandra L. 1997. Himalayan Bronzes, Technology, Style, and Choices. Newark and London: University of Delaware Press and Associated University Press. Reynolds, Valrae. 1981. “From a Lost World—Tibetan Cos­ tumes and Textiles.” Orientations 12, 3:6–22. ———. 1997. “Luxury Textiles in Tibet.” In Tibetan Art: Towards a Definition of Style, pp. 118–131. Eds. Jane Casey Singer and Philip Denwood. London: Laurence King. Rhie, Marylin, Robert A.F. Thurman, Gennady Leonov, and Kira Samosyuk. 1991. “Wisdom and Compassion. The Sacred Art of Tibet.” Arts of Asia 21, 5:80–96. Rhie, Marylin M., and Robert A.F. Thurman. 1999. Worlds of Transformation, New York: Tibet House. Ricca, Franco, and Erberto Lo Bue. 1993. The Great Stupa of Gyantse. London: Serindia.

bibliography

421

Rice, Tamara Talbot. 1969. “Some Reflections on the Subject of Arm Bands.” In Forschungen zur Kunst Asiens. In Memoriam Kurt Erdmann, pp. 262–77. Eds. Oktay Aslanapa and Rudolf Naumann. Istanbul: Baha Matbaasi. Richards, Donald S. 2002. The Annals of the Saljuq Turks. London: Routledge and Curzon. Richardson, Hugh E. 1985. “The Skar-cung rdo-rings.” In A Corpus of early Tibetan Inscriptions, pp. 72–81. London: Royal Asiatic Society. Rizvi, Janet. 2005. Ladakh—Crossroads of High Asia. New Delhi: Oxford University Press. Rockhill, William W. 1884. The Life of the Buddha and the Early History of His Order, derived from Tibetan works in the Bkah–hgyur and Bstan–hgyur. Followed by Notices on the Early History of Tibet and Khoten. London: Kegan Paul, Trench, Trübner & Co. Ltd. 1884. Reprint ed. 1972. Varanasi: Orientalia Indica. Sachau, Edward C. 2002. Alberuni’s India. New Delhi: Rupa and Co. Reprint of 1888 ed. Samuel, Geoffrey. 1992. “Gesar of Ling, the Origins and Meaning of the East Tibetan Epic.” Tibetan Studies. Proceedings of the 5th Seminar of the International Association for Tibetan Studies. Vol. 2, pp. 711–722. Eds. Shoren Ihara and Zuiho Yamaguchi. Narita: Naritasan Shinshoji. Sander, Lore. 1994. “A Graffito with the Quintessence of Buddhist Doctrine from Ladakh.” In Festschrift Klaus Bruhn zur Vollendung des 65. Lebensjahres, pp. 561– 570. Eds. Nalini Balbir and Joachim K. Bautze. Reibek: Verlag für Orientalische Fachpublikationen. Sangs rgyas rgya mtsho. 1989. dGa’ ldan chos ’byung Baiḍūrya ser po. Zi ling (Xining): Krung-go Bod kyi Shes rig dPe skrun khang. Sauvaget, Jean. 1951. “Une Représentation de la citadelle seljoukide de Merv.” Ars Islamica 15–16:128–32. Schroeder, Ulrich von. 2001. Buddhist Sculpture in Tibet. Hong Kong: Visual Dharma. Schiefner, Anton. 1869. “Des Missionärs Jäschke Bemühungen um die Erlangung einer Handschrift des Gesar.” Mélanges Asiatiques 6:1–12. Schmidt, Isaac Jakob. 1839. Die Thaten Bogda Gesser Chan’s, Des Vertilgers der Wurzel der zehn Übel in den zehn Gegenden. St Petersburg: W. Gräff. Schuh, Dieter. 2008. “Die Herrscher von Baltistan (Klein-Tibet) im Spiegel von Herrscherurkunden aus Ladakh.” In Chomolangma, Demawend und Kasbek. Festschrift für Roland Bielmeier zum 65. Geburtstag. Vol. 1, pp. 165–225. Eds. Brigitte Huber, Marianne Volkart, and Paul Widmer. Halle: International Institute for Tibetan and Buddhist Studies. Schwieger, Peter. 1997. “Kah-thog-rig-’dzin Tshe-dbang-nor-bu’s diplomatic mission to Ladakh in the 18th century.” In Recent Research on Ladakh 6. Proceedings of the Sixth International Colloquia on Ladakh, Leh 1993, pp. 219–230. Eds. Henry Osmaston and Nawang Tsering. Bristol: University of Bristol Press. Delhi: Motilal Banarsidass.

422

bibliography

———. 1999. Teilung und Reintegration des Königreichs von Ladakh im 18. Jahrhundert. Der Staatsvertrag zwischen Ladakh und Purig aus dem Jahr 1753. Bonn: VGH Wissenschaftsverlag. Segraves, Julie M. 2001. “An Interview with Lionel Fournier.” Orientations 32, 1:68–75. Seidenstücker, Karl. 1914. Süd-buddhistische Studien. I. Die Buddhalegende in den Skulpturen des Ananda-Tempels zu Pagan. Supplement 9 to Jahrbuch der Hambur­ gischen Wissenschaftlichen Anstalten 32. Hamburg: Lütcke & Wulff. Sela, Ron. 2003. Ritual and Authority in Central Asia: the Khan’s Inauguration Ceremony. Indiana: Indiana University Research Institute for Inner Asian Studies. Shah, Ahmad. 1906. Pictures of Tibetan Life. Benares: E.J. Lazarus. Sharma, Janhwij. 2003. Architectural Heritage, Ladakh. Delhi: INTACH. Shastri, Lobsang. 1994. “The Marriage Customs of Ru-thog (mNga’-ris).” In Tibetan Studies: Proceedings of the 6th Seminar of the International association for Tibetan Studies, Fagernes 1992, pp. 755–767. Ed. Per Kvaerne. Oslo: Institute for Comparative Research in Human Culture. Sheikh, Abdul Ghani. 1995. “Ladakh’s Relations with Central Asia.” In Recent Research on Ladakh 7. Proceedings of the 7th Colloquium of the International Association for Ladakh Studies held in Bonn/Sankt Augustin, 12–15 June 1995, pp. 447–456. Eds. Thierry Dodin and Heinz Räther. Ulm: Universität Ulm. ———. 2005. “Islamic Architecture in Leh.” In Ladakh—Culture at the Crossroads, pp. 31–43. Eds. Monisha Ahmed and Clare Harris. Mumbai: Marg Publications. ———. 2007. “Transformation of Kuksho Village.” In Recent Research on Ladakh 2007, pp. 163–170. Eds. John Bray and Nawang Tsering Shakspo. Leh: Jammu & Kashmir Academy of Art, Culture & Languages—International Association for Ladakh Studies. Sims, Eleanor. 2002. Peerless Images. Persian Painting and its Sources. New Haven and London: Yale University Press. Sims‑Williams, Nicholas. 1993. “The Sogdian Inscriptions of Ladakh.” In Antiquities of Northern Pakistan, Reports and Studies 2, pp. 157–163. Ed. Karl Jettmar. Mainz: Verlag Philip von Zaubern. Singer, Jane Casey. 1997. “The Sublime Image: Early Portrait Painting in Tibet.” In First Under the Heaven: The Art of Asia, pp. 126–141. London: Hali Publications. Singer, Jane Casey, and Philip Denwood (Eds.). 1997. Tibetan Art: Towards a Definition of Style. London: Laurence King. Sinor, Denis. 1998. “The Kitan and the Kara Khitay.” In History of Civilizations of Central Asia. Vol. IV: The Age of Achievement: A.D. 750 to the End of the Fifteenth Century. Part One: The Historical, Social and Economic Setting, pp. 227–242. Eds. Muhammad S. Asimov and Clifford E. Bosworth. Paris: UNESCO Publishing. Sinor, Denis, and Sergei G. Klyashtorny. 1996. “The Turk Empire”. In History of Civilizations of Central Asia. Vol. III. The crossroads of civilizations: A.D. 250 to 750,

bibliography

423

pp. 327–347. Eds. Boris A. Litvinsky, Guang-da Zhong, and R. Shabani Samghabadi. Paris: UNESCO Publishing. Siudmak, John. 1995. “A Group of Three Standing Female Deities from Kashmir and their Antecedents in Late Gandharan art.” In South Asian Archaeology 1993, pp. 681–694. Eds. Asko Parpola and Petteri Koskikallio. Helsinki: Suomalainen Tiedeakatemia. Snellgrove, David. 1987. Indo-Tibetan Buddhism. Indian Buddhists and Their Tibetan Successors. Boston: Shambhala. Snellgrove, David L., and Tadeusz Skorupski. 1977. The Cultural Heritage of Ladakh. Vol. 1. Central Ladakh. Warminster: Aris & Phillips. Snellgrove, David L., and Tadeusz Skorupski, with contributions by Philip Denwood. 1980. The Cultural Heritage of Ladakh. Vol. 2. Zangskar and the Cave Temples of Ladakh. Warminster: Aris & Phillips. Sonavane, Vishwarao. 1992. “Significance of the Chandravati Engraved Core in the Light of Prehistoric Art of India.” In Rock Art in the Old World, pp. 273–283. Ed. Michel Lorblanchet. New Delhi: Indira Gandhi National Centre for the Arts. Stein, Marc Aurel. 1961. Kalhaṇa’s Rājataraṅgiṇī, A Chronicle of the Kings of Kaśmīr. Delhi: Motilal Banarsidass. Reprint of 1892 ed. Bombay: Education Society Press. Stein, Rolf A. 1956. L’Épopée tibétaine de Gesar dans sa version lamaïque de Ling. Paris: Presses Universitaires de France. ———. 1958. “Peintures tibétaines de la vie de Gesar.” Arts Asiatiques 5, 4:243–271. ———. 1959. Recherches sur l’épopée et le barde au Tibet. Paris: Presses Universitaires de France. Steinkellner, Ernst, and Helmut Tauscher (Eds.). 1983. Contributions on Tibetan Language, History and Culture. Proceedings of the Csoma de Körös Symposium Held at Velm-Vienna, Austria, 13–19 September 1981. Vol. 1. Wien: Arbeitskreis für Tibetische und Buddhistische Studien Universität Wien. Stoddard, Heather. 2003. “’Bri gung, Sa skya and Mongol patronage: a reassessment of the introduction of the Newar ‘Sa skya’ style into Tibet.” In Dating Tibetan Art, pp. 59–72. Ed. Ingrid Kreide-Damani. Wiesbaden: Ludwig Reichert Verlag. Stoye, Martina. 2008. “Das Leben des Buddha in der Gandhara–Kunst.” In Gandhara: Das buddhistische Erbe Pakistans, pp. 184–190. Ed. Christian Luczanits. Bonn: Kunst– und Ausstellungshalle der Bundesrepublik Deutschland. Striedter, Karl Heinz. 1992. “Rock Art Research on the Djado Plateau (Niger): A Preliminary Report on Arkana.” In Rock Art in the Old World, pp. 113–128. Ed. Michel Lorblanchet. New Delhi: Indira Gandhi National Centre for the Arts. Takeuchi, Tsuguhito. “Old Tibetan Rock Inscriptions near Alchi.” Unpublished research paper. Tashi Ldawa Thsangspa. 2007. “A Short Note on some Petroglyphs of the Nubra Valley.” In Recent Research on Ladakh 2007, pp. 53–60. Ed. John Bray and Nawang

424

bibliography

Tsering Shakspo. Leh: Jammu & Kashmir Academy of Art, Culture & Languages— International Association for Ladakh Studies. Tashi Rabgias (Bkra shis rab rgyas). 1984. Mar yul la dvags kyi sngon rabs kun gsal me long. History of Ladakh Called The Mirror Which Illuminates All. Leh: C. Namgyal and Tsewang Taru. Teichman, Eric. 1922. Travels of a Consular Officer in Eastern Tibet. Cambridge: Cambridge University Press. Thakur, Laxman S. 2001. Buddhism in the Western Himalaya. A Study of the Tabo Monastery. New Delhi: Oxford University Press. Theuns-de Boer, Gerda. 2008. A Vision of Splendour, Indian Heritage in the Photographs of Jean-Philippe Vogel, 1901–1913. Leiden: Kern Institute. Thub bstan dPal ldan. 1990. dPe thub dgon dGa’ ldan dar rgyas gling gi chags rabs kun gsal me long. Leh: dPe thub Monastery. Thubstan Paldan. 1997. The Guide to the Buddhist Monasteries and Royal Castles of Ladakh. Delhi: Dorjee Tsering at Jayyed Press. Tibet Heritage Fund. www.tibetheritagefund.org. Annual reports 2003–2012. Trewin, Mark. 1990. “A Cross-cultural Study of Instrumental Music in Ladakh.” In Wissenschaftsgeschichte und gegenwärtige Forschungen in Nordwest-Indien, pp. 273– 276. Eds. Gudrun Meier and Lydia Icke-Schwalbe. Dresden: Museum für Völkerkunde. Tropper, Kurt. 2007. “The Historical Inscription in the Gsum brtsegs Temple at Wanla, Ladakh.” Text, Image and Song in Transdisciplinary Dialogue. Proceedings of the Tenth Seminar of the International Association for Tibetan Studies, Oxford, 2003, pp. 105–150. Ed. Deborah Klimburg-Salter, Kurt Tropper, and Christian Jahoda. Leiden– Boston: Brill. Tsering Mutup. 1983. “Kesar Ling Norbu Dadul.” In Recent Research on Ladakh, pp. 9–28. Eds. Detlef Kantowsky and Reinhard Sander. Munich: Weltforum Verlag. Tsering, Nawang. 2009. Alchi: The Living Heritage of Ladakh: 1000 Years of Buddhist Art. Leh-Ladakh: Central Institute of Buddhist Studies; New Delhi: National Museum; Leh-Ladakh: Likir Monastery. Tsong kha pa. 1977. Tantra in Tibet. London: Gorge Allen and Unwin Ltd. Tsong kha pa. 2000. The Great Treatise on the Stages on the Path to Enlightenment, Ithaca-New York: Snow Lion Publications. Tucci, Giuseppe. 1932. Indo-Tibetica. Vol. I. Mc’od rten e ts’a ts’a nel Tibet indiano e occidentale. Roma: Reale Accademia d’Italia. ———. 1933. Indo-Tibetica. Vol. II. Rin c’en bzaṅ po e la rinascita del buddhismo nel Tibet intorno al mille. Roma: Reale Accademia d’Italia. ———. 1935. Indo-Tibetica. Vol. III. I templi del Tibet occidentale e il loro simbolismo artistico. Part I. Spiti e Kunavar. Roma: Reale Accademia d’Italia. ———. 1949. Tibetan Painted Scrolls. 3 vols. Roma: La Libreria dello Stato.

bibliography

425

———. 1971. Opera Minora. 2 vols. Roma: Giovanni Bardi Editore. ——— . (Ed. and Transl.). 1971. Deb t’er dmar po gsar ma. Tibetan Chronicles by bSod nams grags pa. Roma: Istituto Italiano per il Medio ed Estremo Oriente. ———. 1973. Transhimalaya. Geneva: Nagel. New Delhi: Vikas. ———. 1988a. Stupa: Art, Architectonics and Symbolism. Transl. Uma Marina Vesci. Ed. Lokesh Chandra. New Delhi: Aditya Prakashan. (English version of Indo-Tibetica, vol. I). ———. 1988b. Rin-chen-bzang-po and the Renaissance of Buddhism in Tibet around the Millennium. New Delhi: Aditya Prakashan. (English version of Indo-Tibetica, vol. II). ———. 1988. The Temples of Western Tibet and their Artistic Symbolism. Part 1. The Monasteries of Spiti and Kunavar. Delhi: Aditya Prakashan (English version of IndoTibetica, vol. III, part I). ———. 1989. The Temples of Western Tibet and their Artistic Symbolism. Part 2. Tsaparang. Delhi: Satapitaka Series (English version of Indo-Tibetica, vol. III, part II). ———. 1998. Il libro Tibetano dei morti (Bardo Tödöl). Milano: SE. Reprint of 1981 ed. Torino: UTET. Uray, Géza. 1990. “The old name of Ladakh”. Acta Orientalia Academiae Hungarica 44, 1–2:217–224. Vernier, Martin. 2007. Exploration et documentation des pétroglyphes du Ladakh. 1996– 2006. Como: Fondation Carlo Leone et Mariena Montandon. Vernier, Martin, Laurianne Bruneau, and Quentin Devers. 2011. “Archaeological Heritage at Stake.” Ladakh Studies 26:13–14. Vets, Hilde, and Amandus Van Quaille. 1998. “Lamayuru, the symbolic architecture of light” in Sacred Landscape of the Himalaya. Proceedings of an International Conference at Heidelberg 22–27 May 1998, pp. 85–94. Eds. Niels Gutschow, Axel Michaels, Charles Ramble, and Ernst Steinkellner. Vienna: Verlag der Österreichischen Akademie der Wissenschaften. Vitali, Roberto. 1990. Early Temples of Central Tibet. London: Serindia. ———. (Ed.). 1996. The Kingdoms of Gu.ge Pu.hrang, according to mNga’.ris rgyal.rabs by Gu.ge mkhan.chen Ngag.dbang.grags.pa. Dharamsala: Tho.ling gtsug.lag.khang lo.gcig.stong ’khor.ba’i rjes.dran.mdzad sgo’i go.srig tshog.chung. ———. 1997. “Nomads of Byang and mNga’-ris-smad: A Historical Overview of Their Interaction in Gro-shod, ‘Brong-pa, Glo-bo and Gung-thang from the 11th to the 15th Century”. In Tibetan Studies. Proceedings of the 7th Seminar of the International Association of Tibetan Studies, Graz 1995. Vol. II, pp. 1023–1036. Eds. Helmut Krasser, Michael Torsten Much, Ernst Steinkellner, and Helmut Tauscher. Wien: Verlag der Österreichischen Akademie der Wissenschaften. ———. 1999. Records of Tho.ling: A Literary and Visual Reconstruction of the “Mother” Monastery in Gu.ge. Dharamsala: High Asia.

426

bibliography

———. 2003. “A Chronology of Events in the History of mNga’ ris skor gsum.” In The History of Tibet. Vol. I, pp. 90–105. Ed. Alex McKay. London and New York: Routledge Curzon. Vohra, Rohit. 1993. “Antiquities on the Southern Arteries of the Silk Route: Ethnic Movements and New Discoveries.” In Anthropology of Tibet and the Himalaya: 386–405. Ed. Charles Ramble & Martin Brauen. Zürich: Ethnological Museum of the University of Zürich. ———. 1994. “Sogdian inscriptions from Tangtse in Ladakh.” In Tibetan Studies. Proceedings of the 6th Seminar of the International Association for Tibetan Studies. Fagernes 1992: 920–929. Ed. Per Kvaerne. Oslo: Institute for Comparative Research in Human Culture. ———. 1995. “Arabic Inscriptions of the Late First Millennium A.D. from Tangtse in Ladakh.” In Recent Research on Ladakh 4 and 5: 419–429. Eds. Henry Osmaston and Philip Denwood. London: School of Oriental and African Studies. ———. 1999. “Tamgas and Inscriptions from Tangtse in Ladakh.” In Studia Tibetica et Mongolica (Festschrift Manfred Taube): 279–304. Ed. Helmut Eimer, Michael Hahn, Maria Schetelich and Peter Wyzlic. Vol. 34. Indica et Tibetica. Swisttal-Odendorf: Indica et Tibetica Verlag.  ———. 2005. Petroglyphs in Purig Area in Ladakh. Himalayan Studies. Ladakh Series. Vol. V. Grosbous (Luxembourg): published by the Author. VV. AA. Forthcoming. Drukpa Heritage—The Drukpa Heritage in Ladakh. Delhi: Drukpa Publications. Walravens, Hartmut (Ed.). 2001. Briefe und Dokumente / Albert Grünwedel. Wiesbaden: Harrassowitz. Walravens, Hartmut, and Manfred Taube. 1992. August Hermann Francke und die Westhimalaya Mission der Herrnhuter Brüdergemeine. Stuttgart: Franz Steiner Verlag. Wandl, Erna. 1997. “The Representation of Costumes and Textiles.” In Tabo. A Lamp for the Kingdom, pp. 179–87. Deborah E. Klimburg-Salter, with contributions by Christian Luczanits, Luciano Petech, Ernst Steinkeller, and Erna Wandl. Milano: Skira. Warmington, E.H. 1947. The Commerce between the Roman Empire and India. Delhi: Munshiram Manoharlal. Wayman, Alex. 1990. The Buddhist Tantras—Light on Indo-Tibetan Esotericism. Delhi: Motilal Barnarsidass. Wei, Huo, and Yongxian Li. 2001. The Buddhist Art in Western Tibet. Chengdu: Sichuan People’s Publishing House (in Chinese, with English summary). Wessels, C. 1924. Early Jesuit Travellers in Central Asia. 1603–1721. The Hague: Martinus Nijoff. Reprint ed. 1992. New Delhi and Madras: Asian Educational Services. Whalley, V.B. 1983. “Desert Varnish.” In Chemical Sediments and Geomorphology: 197– 226. Eds A. Goudie & K. Pye. London: Academic Press.

bibliography

427

Whitfield, Susan (Ed.). 2004. The Silk Road. Trade. Travel, War and Faith. London: The British Library. Willis, Janice. 1995. Enlightened Beings. Boston: Wisdom Publications. Winkler, Jakob. 2002. “The rdzogs chen murals of the Klu khang in Lhasa”. In Religion and Secular Culture in Tibet. Tibetan Studies II. PIATS 2000: Tibetan Studies: Proceedings of the Ninth Seminar of the International Association for Tibetan Studies, Leiden 2000. Vol. 2, pp. 321–343. Ed. Henk Blezer, with the assistance of Abel Zadoks. Leiden— Boston—Köln: Brill. Wu, Junkui, and Jianlin Zhang. 1987. “Xizang Ritu xian gudai yanhua diaocha jianbao.” In Wenwu 2: 44–50. Cited in Francfort et al. (1990, 1992). Yablonsky, Gabrielle. 2000. “Sculpture in Bhutan: the Tshogs Zhing in the Paro Museum.” In Impressions of Bhutan and Tibetan Art. Proceedings of the Ninth Seminar of the IATS 2000. Vol. 3, pp. 49–67. Eds. John Ardussi and Henk Blezer. Leiden: Brill. Zab chos zhi khro dgongs pa rang grol las bar do thos grol gyi skor ba bzhugs so. See under Karma gling pa. Ziegler, Verena. 2008. Das Leben des Buddha Íåkyamuni am Holzportal des dKar chung lha khang in Nako. M.A. thesis. University of Vienna. Zwalf, Wladimir. 1996. Catalogue of the Gandharan Sculpture in the British Museum. London: British Museum Press.

Index ’Bebs ma  239 ’Bhag dar skyabs  165 ’Bri gung pa  87, 159, 164–166 ’Bri gung ti se lo rgyus  165 ’Brom ston pa  325–326 ’Brug mo  170 ’Bum seng ge  236 ’cham  329, 331 ’chi kha’i bar do (intermediate state of the moment of death)  279–280 ’Do sde  227 n. 5 ’du khang see dukhang ’Jig rten mgon po  159, 162, 165 ’Od lde  188, 226 ’Phan po see Phenpo Aṅgaja  326 abhaya mudrā  251 Abhidharmakośa  320 Achi Association  225, 376 n. 12, 394, 398 Agni (Tib. Me lha)  251 Ajanta  181 n. 23 Akshobhya (Akṣobhya)  55 Vajrasattva  281 (Fig. 11.4) Alaipura  344 Alberti, Leon Battista  389 Al-Bīrunī  187–188 Alchi (A lci)  1, 3–4, 6–9, 16, 35–36, 38, 42, 44, 58, 61, 75, 87, 101, 121, 126, 139, 141–145, 149–155, 157–168, 175, 178, 180–182, 184–186, 189, 193–194, 199, 202–204, 206, 208–212, 214, 216–217, 219–225, 265–266, 335, 343, 361–362, 366, 374, 390, 394 Dukhang (’Du khang)  7, 159, 167–169, 172–178, 181 n. 22, 182 n. 25, 183–184, 189–190 Sumtsek (gSum brtsegs)  49, 152, 159, 223–224, 266, 335 Alexander, André  viii, 10, 117, 140 n. 23, 225, 313, 349, 351, 398 Al-i Afrasiyab  187 al-Qaqaniya  187 Altar covering  331 amalaka  213 Amitabha (Amitābha)  55, 293

Amitāyus  326 Amoghasiddhi  55, 293 Amoghavajra  236 Andree, Richard  301 Antinoë, Egypt  183 anuttarayogatantra  164, 314, 324, 327 apron  329 Archaeological Survey of India, ASI  3, 260, 277, 303, 304 n. 15, 390–391 arches Kashmiri  195 stepped  195 trefoil  204 n. 14, 206, 221 n. 26 trilobed  204 arga (ar ka)  356, 373 armbands  181 Arslan  187, 189 Asaṅga  326 Assembly Hall  7, 9, 167, 169, 177–178, 180–182, 361, 393 Atiśa  325, 327 Avalokiteśvara  40, 48–49, 53, 55, 58, 60–61, 144 n. 7, 275–276, 325–326 Avantipur  200 n. 8 Avantisvamin  204 n. 14 Azevedo, Francisco de  88–90, 91 n. 17 Ba sgo, Bab sgo see Basgo Babu Pindi Lal see Lal, Babu Pindi bagdas  172, 185 Bakula Rangdröl Nyima Rinpoche (Ba ku la Rang grol nyi ma Rin po che), Bakula Rin po che  8, 274–275, 277, 293, 297 Balasaghun  189 Baltistan  8, 40–42, 65, 84, 86–87, 188, 257–259, 264–265 Bar do mural painting  279, 292 Bar do thos grol chen mo (Tibetan Book of the Dead)  9, 276, 278, 279 n. 12, 280 n. 13 Barkor  395 Basgo (Ba sgo, Bab sgo)  228 n. 5, 251, 262 bCu gcig zhal  232 n. 12 bDe ldan rnam rgyal  337 n. 6

index Be ri see Beri Bellini, Chiara  viii, 7, 277, 293, 297 Bema, petroglyphs  20 Benaras see Varanasi Beri (Be ri)  96–97 Bha ra dan dur  187 Bhagan  166 Bhai ṣajyaguru  326 Bhaktapur  395 bhaṭṭaraka (Tib.rje btsun)  188 n. 34 Bhattavaryan  188 Bilge Qaghan  172 bKa’ brgyud, bKa’ brgyud pa see Kargyu, Kargyupa  73 bKa’ gdams pa  133 n. 17, 229, 323 n. 10, 324 n. 10, 327 bKra shis rnam rgyal  89–90, 92, 160, 169 n. 3, 350 Bla ma Blo bzang rdo rje ’chang  315 Bla ma mchod pa  9, 314 n. 1, 315, 316 n. 3, 320, 322, 323 n. 9, 324 n. 11 Bla ma Ye shes dPal  236 Blancke, Kristin  viii, 9 Blo bzang ’jam dpal lhun grub  323 Blo gros mchog ldan  231 Bodhisattva  40, 42, 49, 53, 55, 58, 65, 116, 126, 129, 133, 144 n. 7, 155–156, 199, 202, 206, 208 n. 17, 235 n. 15, 316, 325–327 Brag khung kha ba chen  226 Brahmā (Tib. Tshangs pa)  246 Brahmaur  203, 211, 214, 223–224 brāhmī  30, 38, 44 Bray, John  vii, 8–9, 67, 88 n. 13, 93, 158, 273, 297, 387 bronzes  35, 38–39, 42, 49, 55, 58, 61, 219 Bru zha  188 n. 35 bSod nams grags pa  87, 227 n. 5, 229 Buddha archery  199, 202 biography  3, 199, 200 n. 9, 208 depiction of life story  193, 199–201, 208 n. 17, 223 Enlightenment  210 episodes from life story  193, 199, 201 four encounters  199, 201 n. 10 great departure  199, 202 throwing the elephant  199, 225 wearing jewellery  195, 204 wrestling  199

429 Buddha Heruka  285 (Fig. 11.9) Buddhist symbols  331, 338, 344 Burra Charter  390 Byama Khumbu (Suru Valley)  36, 40, 55 Byang go la see Changgo La Byang la see Chang La Cakrasaṃvara  324 canopy  79, 237 cape  163, 176–178, 334, 338, 346 carving, wood  194–195, 209, 211, 214, 219, 223 n. 30, 224, 254, 265–266, 273 castle architecture  2 Central Asian animal style  271 Central Asian Trade  58, 396 Chaba Mission  337 Chamba  60, 211, 214, 224 Chamba Lhakhang, Leh  395 Chandan  258, 261 chang  96–98, 176, 276, 286 n. 15 Chang La  60, 97 Changgo La  94, 97, 99 Changspa (Changs pa)  9, 48–49, 53, 298–299, 303–305, 306 n. 19, 309–312 Changthang  30, 96, 98–99, 383 Chaturvedi, Anuradha  393, 398 Chemchok (Che mchog) Heruka  285, 286 n. 15 Chemre (lCe ’bre, lCe bde)  396 Chengzu  226 n. 4 Chenrezik (sPyan ras gzigs)  8, 60, 275, 277, 292–293, 297 Chhutay Ratak  396 Chigtan, Chiktan (Cig gtan, Cig ldan)  8, 73, 210–211, 223, 254, 255–262, 264, 266–269, 271, 273, 391 Chilas  17–18, 91, 188 Chilling  23, 25 Chinese inscription  27 Choskhor (Chos ’khor)  6 chos kyi rgyal po  234, 295 chos nyid bar do (intermediate state of ultimate reality)  279, 280 n. 13, 284 Chos rab  236 Chos rje Don grub rin chen  236 Chos sde dKar rgyas  227 n. 5 column capital  216, 219, 221–222 fluted  193

430 conservation architectural  11 models of  352 wall paintings  361–362, 368 construction materials and tradition   11–12, 34, 44, 48, 53, 66, 76, 79, 101, 139–141, 145, 147, 149–152, 157–158, 191, 194, 212, 221, 224, 234, 257–258, 260, 264–266, 268, 277, 335 n. 4, 356, 364–365, 369, 378, 386 n. 24 corbelled roof  149–151 Csoma de Kőrös, Alexander  299, 394 cup offering and hunt  183, 185–186, 189 rites  170–171 cupbearers female  170, 174 male  170 ’Du khang dkar po  234 Dainelli, Giotto  267 Dalai Lama, XIV  92, 130, 227 n. 5, 229, 315, 320, 322, 323 n. 9, 337, 346, 350 Dangra Yumtsho  96 Dards  36, 49, 90–91 Dargo (Dar go)  255, 262 dBang gi gShin rje  237 dBang grub  93–95 dBang phyug chos rje  229 n. 9 dBus byang see Üchang De mur  188 Dede Korkut  170, 174–175 defensible settlement  71, 73 Deldan Angmo, architect  395 Deldan Namgyal  337 n. 6 Deogarh  210 devata  195, 203 dGa’ ldan byang rtse  229, 323 n. 9 dGa’ ldan grva tshang  229 dGa’ ldan snyan rgyud  324 n. 10 dGe lugs pa see Geluk Dhṛtarāṣṭra  326 dharma  65, 234–235, 237, 278 n. 9, 314, 322 n. 8, 325, 327 n. 14, 379 dharmapāla  237, 320 Dharmaraja (dharmarāja)  237, 287, 295, 325 dhoṭī̄  39, 42, 53, 55, 61, 126, 156, 182, 195, 197, 204, 242 n. 21, 246, 249

index Digar, Nubra Valley  36, 60–61 Dīpankararakṣita  236 Diskit  61, 351, 368, 394 dKar sa  101, 227 n. 5 dKyil khang  231 dNgos grub mgon  165 Dogra invasion of Ladakh  8, 275, 311 Dome (mDo smad)  94–95 Domkhar, petroglyphs  20, 27, 271 door hanging  331 dPal ’khor chos sde  233 n. 13 dPal gShin rje gshed nag po’i rgyud rtog pa gsum pa  235 n. 17 dPal ldan lha mo  113, 234, 238, 244–245 dPe thub see Spituk Drag gi gShin rje  237 Dragspa Bumdé (Grags pa ’bum lde)  361 Dras  35–36, 38, 40, 48, 86, 91 dress, women’s  176, 178, 338 Drikung, ’Drigung (’Bri khung, ’Bri gung) 7, 97, 126, 165–166, 277 drinking scene  167–169, 171, 173–176, 179, 183, 185, 189 Du ram  94, 97 dukhang (’du khang)  7, 73–74, 159, 167, 314 see also Alchi Dukhang Dulan  97, 99 Dungkar (Dung dkar)  178 n. 15 Dunhuang  336 earth architecture  386 n. 24 earthen building material  10, 367, 378 Eastern qaghanate  187 entry hall  176 Everding, Karl-Heinz  274 n. 1 façade, tripartite  214, 221 n. 26 foundation scene  169 Francke, August Hermann  3–4, 9, 14, 16, 35, 42, 44, 49, 68–69, 74, 76, 79, 82–84, 90–92, 102 n. 3, 109 n. 6, 123 n. 12, 128, 176, 254–255, 257–258, 260, 264, 271, 273, 300–306, 308–309, 311–312 Gandhara  200–201 gandharva  42 Gaṇeśa  251 Ganzhou  96, 170 Garuḍa  126, 137, 152, 250, 271

index Gauri (Gaurī)  286 Gawa Marser (dGa’ ba dmar gser)  286 (Fig. 11.12) geology  387 Geluk (dGe lugs)  7–8, 87, 105 n. 4, 144, 226, 227 n. 5, 228–229, 231–234, 314 n. 1, 315, 320 n. 6, 322 n. 8, 323 n. 10, 324–325, 327 Gendündrup (dGe ’dun grub)  350 Gesar (Ge sar)  82, 90, 94–99, 170, 176, 305, 307 n. 21 Ghaznavids  181, 188 Gilgit  8, 40, 86, 91, 188 n. 35, 255 Gling see Ling Gling tshang see Lingtsang Globus  301–302, 304–305 gnod  244 gNubs chen  188 gnya’ bo  176 goatee  179 Goloks  96 Gompa Soma, Leh  395–396 Gongka Maru (Gong kha dmag ru)   94–95, 97 Gopa  199 gos chen  292, 331, 338–340, 342, 344, 346 Grags ’bum lde  226, 230–231, 350 Grags ldan ’od  164 Grünwedel, Albert  300–302, 304 gSang mkhar see Sangkar gSang phu pa lHa dbang blo gros see lHa dbang blo gros gsang sgrub  237 gSer khang  231 gShin rje chos kyi rgyal po  237 n. 20 gShin rje dkar po (Sita Yamarāja)  237 gShin rje dmar po (Rakta Yamarāja)  237 gShin rje nang grub (Antarasādhana Yamarāja)  239 gShin rje ser po (Pīta Yamarāja)  237 gShin rje sngon po (Nīla Yamarāja)  237 Guge (Gu ge)  6, 83–84, 90, 93, 95–97, 99, 141, 178, 186–188, 190 Guhyasamāja  324 Gupta period  195, 203, 208 n. 16, 210–211, 222 n. 26, 223 Gu ru lHa khang, Phyi dbang  126, 232, 251 Guru Lhakhang, Leh  354, 395

431 Gya (rGya)  6, 18, 68–71, 73–74, 76, 79–80, 82–86, 88–95, 97–99, 131 n. 16, 249, 294, 331, 339–340, 344–346, 391, 394 Gya mur orr  86 Gyalse Rinpoche  332 (Fig. 14.3), 336 Gyade  98 Gyal Khatun (rGyal Khatun)  257, 268 Gyantse (rGyal-rtse)  233, 234 n. 14, 395 Gye sar  90, 93–95 Gyurme Dorje  279 n. 12, 287 n. 16 gZi De khyim  188 hairstyle  176, 178, 180–181, 183, 185, 195, 197, 203, 208 n. 16, 218, 223, 251 Haji Kasim  340 Haji Mohammad Ishaque  340 haloes  174 Hankar  73, 266, 269 Hanle (Wam le)  1, 83–85, 258 Hashmatullah Khan  256, 259–260 hat  331, 334 Hayagrīva (rTa mgrin)  49 headband  183, 185 Hemis, Himis (He mis)  100, 259, 329, 336, 393, 398 Henasku  73 heritage conservation, models  11 Herrnhut, Völkerkundemuseum  302 n. 11, 308, 309 n. 23, 311, 313 Heruka  239, 244, 280 n. 13, 285 Hevajra  324–325, 327 Hevajratantra  328 Himachal, Himachal Pradesh  7, 65, 176, 203, 211, 222 n. 26, 223 Himalayan Cultural Heritage Foundation, HCHF  396, 398 Ho  254–255, 259 Hor  98, 187, 298, 300, 306 n. 19, 312 Hor nag mo A lan  187 Hor ser see Yellow Uighurs Hudud al-Alam  86 Hundar, Nubra Valley  36 iḍā  235 n. 16, 237 iconography  6–7, 42, 49, 60, 142, 146, 153, 167 n. 1, 168, 170, 171 n. 8, 172, 189, 199–200, 206, 208, 222, 233, 235 n. 16, 246, 311, 320 n. 6, 321–322, 324, 328

432 Indian National Trust for Art and Cultural Heritage, INTACH  267, 391 Indra (Tib. brGya byin)  249 Inner Asia  171 n. 9, 183, 186 inscriptions  3, 7, 16, 30–31, 34 International Centre for the Study of the Preservation and Restoration of Cultural Property, ICCROM  390 International Council on Monuments and Sites, ICOMOS  390 Iranian influence  171 Īśāna (Tib. gNod sbyin dbang ldan)  248 Ishvari (Īśvarī)  286 Islamic iconography  172 ivory  200 n. 9, 202–203, 221 n. 26 Jain  182, 347 Jambhala  244, 253, 326 Jammu and Kashmir, J&K  8, 108 n. 5, 256 n. 3, 260, 299, 353, 359, 363, 393 Jamyang Namgyal (’Jam dbyangs rnam rgyal), King  257–259, 268, 351 Jäschke, Heinrich August  299, 309 jātaka  200 n. 10, 201 jewellery  111, 178, 195, 197, 204, 213, 216 Jñānakaragupta  236 Jokhang  308, 395 Jokisaari, Laura  277, 297 Kabul  91, 98 Kadam (bKa’ gdams)  117 Kālī  245 Kalimpong  339–340 Kalon (bKa’ blon)  9, 298–300, 303, 306 n. 19, 312 Kanji  191 n. 1, 380, 391, 394 kapāla  238, 243, 248, 325 Kar gling zhi khro  276, 280 Kargil, Kargilo (dKar skyil, dKar gyil lo)  5, 12, 14, 36, 40–42, 55, 63, 254–255, 336, 342, 373, 380 Kagyu, Kagyupa (bKa’ brgyud, bKa’ brgyud pa)  159, 162 Karma Heruka  285 (Fig. 11.9) Karma pa  229 Karsha  101, 121, 133 n. 17, 139, 393 Kartse Khar (Suru Valley)  40 Kashgar  40, 170, 187, 336

index Kashmir  1, 3, 5–8, 17, 35, 38–40, 42, 49, 55, 58, 61, 66, 83–83, 108, 141, 154, 165, 187–188, 193, 195 n. 4, 200–204, 210–211, 221, 224, 256 n. 3, 260, 273, 299, 303, 336–338, 340, 342, 353, 359, 361, 363, 393 Kathmandu  393 n. 1, 395 Kaya Dorjay Angdus  297 Kazakhstan  28 Kennion, R.L.  311 Kere Kharu, petroglyphs  33 (Figs. 1.19, 1.20) Kern Institute see Leiden, Kern Institute Kesar see Gesar Keurima  286 Kha rag  97 khaṭvāṅga  325, 243 Krodharāja sMe brtsegs see sMe brtsegs Khalatse (Kha la rtse, Kha la tse)  12, 14 n. 4, 16, 32, 34, 36, 38, 40, 44, 66, 75, 260, 265, 278 n. 10, 303 Kham (Khams)  9 n. 2, 95, 293–294, 298, 300, 306 n. 19, 312 khan  3, 8, 88, 171, 174, 187, 189, 256, 260, 273, 337 n. 6 Kharoshti  38, 66 Khotan  40, 61, 63, 187–188, 336 Khrig se, Khri rtse see Thikse khrom tshogs  315 Khyi tra  234 Kinkhwab  340 n. 11, 345, 338–339 Kinnaur  3–5, 148, 193 n. 3, 223, 276, 377–378 kīrtimukha  195, 211 Klu dkyil see Likir Kök Turk  172 Könchok Gyatso, Lama  275 n. 4, 277 n. 9, 292, 297 Könchok Rigzin, Lama  275 n. 2, 297 Kṛṣṇayamārikarmasarvasiddhikaraṇanāmata ntra  235 n. 17 Kṛṣṇayamārisarvatathāgatakāyavākcittakṛṣṇ ayamārināmatantra  235 n. 17 Kubera  253 Kumāratantra  246 Kun dga’ rgyal mtshan  227 n. 5 Küntuzangmo (Kun tu bzang mo)  280 Küntuzangpo (Kun tu bzang po)   279–280, 285

index Kushana  38, 40, 44 Kushok Bakula Rinpoche (sKu shogs Ba ku la Rin po che)  228 n. 6 Kutadgu Bilig  170 Kyelang (Lahul)  303 La dvags rgyal rabs  3, 14, 82, 89 n. 14, 97–98, 165, 176, 186, 226 n. 2, 229–230, 257, 298, 300 Lachuse  7, 191, 193–194, 200, 202–204, 208, 209 n. 18, 210–211, 214, 221–225 Ladakh Arts and Media Organisation, LAMO 396 Ladakh Autonomous Hill Development Council, LAHDC  353, 393 Ladakh Ecological Development Group, LEDeG  398 Lahaul  65 Lal, Babu Pindi  3, 8–9, 260, 264–266, 303–304 Lalitavajra  235 n. 17, 236 Lalitavistara  199 n. 6 Lam rim  314 n. 1, 315, 321–322, 324 Lamayuru (Bla ma g.yu ru)  7–8, 73, 100–101, 108 n. 5, 121, 133 n. 17, 191 n. 1, 232 n. 12, 274–275, 277, 278 n. 10, 279, 286 n. 15, 292, 297, 369, 374, 380, 392 lan tsa  248 lantern ceiling  129–130, 133, 139, 150 lay people  169, 311 lCam sring  234 Leh (Sle, Gle)  169 n. 3 Leh Old Town Initiative, LOTI   353–354, 363, 395 Old Town  10, 12, 49, 225, 350, 353, 362, 387 n. 25, 393–394, 399 Palace  266, 268, 350–351, 361, 392 urban conservation  395 Leiden, Kern Institute  304, 313 lha bab (‘Divine Descent’) mchod rten  76, 79 lHa chen Bha gan  89 lHa chen dPal gyi mgon  5, 53, 83–85, 93 lHa chen Khri gtsug lde  75 lHa dbang blo gros  230–231, 233 lha khang  8, 126, 223, 227–228, 232–235, 269, 275, 350 lHa khang dKar po  233 lHa khang dMar po  233

433 lHa khang So ma, A lci  126, 232 lHun gyis grub pa’i rig ’dzin  284 (Fig. 11.8), 294 Lhasa urban conservation  395 Likir (Klu dkyil)  229, 369–370, 372 n. 6, 380, 393–394 Ling (Gling)  9, 94–99 Lingtsang (Gling tshang)  95–99 Lo Bue, Erberto  vii–viii, 4, 9 n. 2, 67, 76, 225–226, 227 n. 5, 278 n. 10, 297, 313, 373 n. 8, 387 Lo tsa ba mchod rten (Lotsava’s Translator’s stupa)  3, 38, 49, 75–76, 79–80, 100, 109 n. 7, 123, 131, 141, 232, 265 Lonpo House, Leh  394, 396, 398 Lopchak Mission  338, 342 Lotsava, Lotsawa (lo tsa ba)  6, 100, 108 n. 5, 117, 141–142, 144, 148, 211, 219, 222–223, 274 Lukhang Temple  276 Madhyamaka  316, 323 n. 10, 325, 327 Mahakala (Mahākāla)  111, 234, 244, 326, 329, 344 Mahākāla Ṣaḍbhuja  326 Mahāvairocana  325 Mahāvairocanābhisaṃbodhitantra  325 Mahāvajrabhairava  235 n. 17, 2360 Mahāvajrabhairavatantra  251 Mahāyāna  235 n. 17, 336, 386 Mahottara Heruka  285 n. 15 Maitreya  10, 35, 42, 48, 53, 55, 58, 60–61, 63, 65, 155 (Fig. 5.21), 160, 199, 251, 316, 349–350, 359, 361, 390 Majorian  389 mala (mālā)  195, 204, 206, 215 mandala (maṇḍala)  58, 111, 126–127, 130, 139 n. 20, 152, 164, 211, 236–237, 285 n. 15, 286, 320 Mang nang  231 Mangyu (Mang rgyu)  42, 101, 108, 121, 167 (Map 7.1), 169, 175, 183, 186, 189, 195, 203, 211, 219, 221 n. 26, 223, 237, 242 n. 21, 356, 380, 393 Manikhang (Maṇi khang), Leh  349–350, 359 Manjushri, Mañjuśrī̄  53, 63, 126 n. 14, 144 n. 7, 206, 211, 219, 222–223, 235, 316, 322 n. 8

434 Mar yul  3, 76, 82–84, 86–87, 141, 165, 227 Māra (Tib. bDud)  235 markalak (mar ka lag)  10, 13, 355, 365, 374, 378–379, 383 Markha (dMar kha) Valley  73, 266, 269 Marshall, Sir John  390 Marx, Karl  9, 16 n. 1, 300 Maryul (Mar yul)  82–84, 86–90, 91 n. 17, 92–93 Masrur  203, 211 Maya (Māyā)  127 (Fig. 4.25) mchod rten (stupa) walls  75 mDo sde rin chen  227 n. 5 mDo smad see Dome medieval history, early  168, 186 mGon dkar  234 mgon khang  7, 73, 169 n. 3, 226, 231–234, 237, 239, 246, 251 Michelangelo  389 Mirza Haidar Dughlat  88 mKhas grub rje, mKhas grub dGe legs dpal bzang po  230, 233 n. 13, 236, 320 n. 6 mNga’ ris  5, 82–84, 90, 93, 96–99, 186–188, 231 mNga’ ris rgyal rabs  90, 93, 99, 186–188 mNga’ ris skor gsum  5, 82 Mon  76, 264 Mongol court ceremonies enthronement ceremony  171 Moorcroft, William  9, 42, 89, 92, 338 mtho rtsegs  315 mud, shuttered  74, 262, 266–267 Mughal  337–338 Mulbek (Mul bhe)  5, 35–36, 42, 44, 48, 55, 63, 390 Munshi House, Leh  396 murals, secular  168, 189 Muslim  1, 8, 10, 171 n. 10, 188–189, 256, 338 Nāga  271, 275 n. 4 Nāgārjuna  325 Nako  148, 200–201, 203, 208–211, 223 Nalanda  182 Nam mkha’ ba, Nam mkha’ pa‘  229–230 Nam mkha’ dpal ba  229, 231 n. 11 Nam mkha’i dbang po phun tshogs lde   231 Namgyal dynasty  334

index Namgyal Institute for Research in Ladakhi Arts and Culture, NIRLAC  391 Nanda  325 nang sgrub  237 Nang sgrub Srin gi gdong can  237 Naraka infernal deities  276 Nāro Ḍākinī  243 Naropa (Nāropa)  126 n. 14, 324–325, 327 Nātha Mahākāla  244 national monuments  390, 392–393, 398 neolithic sites  16 Nestorian inscriptions  31 Newar stylistic features  234 Ngari Korsum (mNga’ ris skor gsum)  5 Nges pa thang  93–94, 96–97 niche, in Kashmirian architecture   193 n. 3 Nimo  36, 44 Nirṛti  248 nomadic customs  171 northern areas of Pakistan  17, 36, 86, 187–188 Nubra  5, 16, 18, 20, 26, 36, 60–61, 63, 65, 84–85, 88, 266, 365–366, 368, 378, 394 Nyar ma  6, 141, 232, 234 Oghuz  170–171, 174–175 oral history  90 tradition  14, 170, 229–230, 254, 256, 259, 364 Padmapani (Padmapāṇi)  53 Palgyigon (dPal gyi mgon) see lHa chen dPal gyi mgon Padma Heruka  285 (Fig. 11.9) Padmavajra  236 painting, mural  73, 145, 275–276, 292–293, 295, 297–298, 321 Pala  182 Paleolithic sites  16 Pallis, Marco  278 n. 10 Panchen Lama, I  315 Pandrethan  204 n. 14 Panggong Lake  96 Panjikent  169 n. 3 Panthaka  326 Parihasapura  195 n. 4 parivāra  234, 253

index

435

Pārvatī  248 Pashkyum (Pas kyum, Pas skyum) 255–257 Pashmina  337 Patan  395 peaceful deities  279, 282–284 perak (pe rag)  178 peristyle  221 n. 26 Persian painting  186 Petech, Luciano  3–4, 35, 83–84, 89 n. 14, 166, 188, 228 n. 6, 229 n. 8, 350 n. 4 Petrarch  389 petroglyph, petroglyphs  4–5, 8, 12, 15–18, 23, 27, 30–32, 34, 36, 38, 44, 271, 391 Pha bong kha bDe chen snying po  322, 323 n. 9 Phenpo  97 Phugtal (Phug tal)  7 Phyag rgya chen po’i rig ’dzin  294 Phyang (Phyi dbang)  65, 126, 232, 370, 392, 398 phyi nang  237 phyi sgrub  237 physiognomic details  179 piṅgalā  235 n. 16, 237 Pili Kothi  344 pillar cover  331 Pishachi (Piśacī)  286 Pit loom  329 plaits  178, 180–181 post-Gupta period  195, 203, 208 n. 16, 210, 223 princely imagery  186 Pṛthivī Devī (Tib. Sa’i lha mo)  249 Punjab  336, 338, 340, 342 Purang (Pu hrang)  5–6, 82, 90, 141, 165, 378, 380, 383 Purig (Pu rig, Bu rig)  73, 84, 89, 92, 254, 259

Ranbirpur  75, 92 Raphael  389 Ratna Heruka  285 (Fig. 11.9) Ratnasambhava  55 rDo rje grags pa  236 Rematī  126 n. 14, 244–246 Renmudong, petroglyph  30, 271 rGya sde see Gyade rGya see Gya rgya ser  331, 339–340, 344, 346 rGyal tshab Dharma rin chen  325 rGyas pa’i gShin rje  237 rGyud sTod  227 n. 5 Ribba  211, 223–224, 377 Rinchenzangpo (Rin chen bzang po)  3, 5–6, 80, 87, 100, 224 Ringdom  191 n. 1 rJe dPal ldan Shes rab grags pa  228 rNam par smin pa’i rig ‘dzin  294 rNam rgyal rtse mo (Namgyal Peak)   169 n. 3, 237 robe, female and male  55, 176, 180, 182, 218, 334, 337–338 rock art  2, 4, 16–18, 20, 27, 31, 34, 36, 55, 271, 391 Rome  11, 158, 273, 323 n. 9, 389–390 Rong pa Shes rab seng ge  236 rTa dbag brgyad  253 rTsa hrang blo stang  231 rTse lde see Tsede Rudok (Ru thog)  30, 60, 83–84, 176, 258 ruler  7–8, 82, 97–98, 165, 171–174, 180 n. 20, 183, 185–186–7, 189, 226, 256, 259–260 Rumtse (Rum rtse)  6, 68–69, 71, 73–75, 79–80, 82, 131 n. 16 Rupshu House, Leh  395 Rva Chos Rab see Chos Rab Rva rDo rje grags pa see rDo rje grags pa

Qara Khan  187 Qarakhanids  170–171, 187–188

Sa la gnas pa’i rig ’dzin  284 (Fig. 11.8) Sabu  60, 75, 131 n. 16 Sad na legs  82 Sakti  36, 58, 60, 88 Sakyamuni, Shakyamuni (Śākyamuni) 145, 242 n. 21, 315, 322, 323 n. 9, 325 Samantabhadra (Samantabhadra)  276 n. 5, 280–281 Samantabhadri (Samantabhadrī)  280

Rab brtan Kun bzang  233 n. 13 Radhu  338 n. 7, 342 Radhu, Abdul Wahid  342 n. 14 Raghurajpur  393 rājalīlā  216, 219 rākṣasa  237

436 Samstanling  394 Sang nang ba  93–95 Sangkar (gSang mkhar)  48–49, 53, 58, 228 n. 6 Sangs rgyas rgya mtsho  227 n. 5, 229–230 Sankar Labrang (gSang mkhar bla brang), Leh  349, 359 Sanskrit  5, 91, 100, 188, 199 n. 6, 237, 323 n. 9 Sapienza University of Rome  389 Sarasvatī  325 Sarnath  339 Sarvavid  325 Sasanian  181, 183 n. 30, 186 Sasoma, petroglyphs  18, 20, 26 Saspol, Saspola (Sa spo la)  36, 44, 75, 154, 166, 227 n. 5 Saspotse  158 (Fig. 5.25), 393 sbyor chos  315 scarf  42, 154, 176, 183 Schiefner, Anton  299 Scythian artistic style  28 Seljuq  180–181, 183 Seng ge lha khang, Lamayuru  232 Sengge Namgyal (Seng ge rnam rgyal) 258–259 ser sa  356 Shang rong  232 n. 12 Shangara  338, 342 Sharma, Janhwij  393 Shes rab bzang po  227 n. 5, 228, 231, 233–234 Shey (Shel)  145 (Fig. 5.5), 227 n. 5, 228, 231, 233–234 Shigatse  395 Shyok  84 Siddhartha  199 Sikri, stupa  200–201 n. 10 silk-brocade  331, 334–335, 337, 339 n. 10 Sita Candra  247 Sita Gaṇapati (Tib. Thsogs dbag dkar po) 251 Sita Mahākāla  244 Sita Varuṇa (Tib. Chu la dkar po)  250 Sitātapatrā  326 Sitatārā  326 Śiva see Īśāna Six samsaric realms  289

index Sixteen Drops of the bKa’ gdams pas   324 n. 10 Skardu  40, 88, 91, 257, 259–260 sKu ’bum  233 Skurbuchan  391, 398 Skyilde Nyimagon (sKyid lde nyi ma mgon) 5, 349 sMe brtsegs  244 sNang grags  93–95 Sogdian inscriptions  31 Sonam Wangchuk  372 nn. 6–7, 373, 374 nn. 10–11, 384 n. 23, 387, 398 Southern Silk Road  187 Spiti  3–5, 65, 84–85, 176, 193 n. 3, 269, 303, 365–368, 378, 393 n. 1 Spituk (dPe thub)  7, 100, 226, 228–230, 232–234, 237, 239, 246, 320, 322–323, 326–328, 365–366, 368–369, 374, 383 srid pa bar do—intermediate state of becoming  279–280, 287, 294 Śrīraktayamāritantrarāja  235n. 17 Śrīvajramahābhairavatantra  235 n. 17 Śrīvajrabhairavakalpatantrarāja  235 n. 17 sTag mo lha khang  227 n. 5, 228 Stagtshang Raspa see Taktsang Repa Stok (sTog)  14, 108 n. 5, 109, 117, 139, 140 n. 22, 264 sTon pa lha khang  234 stupa (stūpa, Tib. mchod rten)  16, 38, 42, 44, 49, 53, 55, 63, 65–66, 70–71, 74–76, 79–80, 82, 100–101, 109 n. 7, 113 n. 8, 123 nn. 11, 13, 126 n. 14, 128 n. 15, 131, 133 n. 17, 141–150, 152–158, 178 nn. 15, 17, 181 n. 22, 182, 191, 200, 201 n. 10, 214, 222, 233, 234 n. 14, 348, 350 n. 4, 392, 398 steppe  8, 18, 28, 149, 170–172, 174–175, 181, 189, 271 sul ma  176, 334, 337 Sumda  42, 101, 195, 211, 219, 223, 394 Sumda Chung  101, 144 (Fig. 5.3), 195, 211, 219, 223, 394 Sumtsek (gSum brtsegs)  7, 9, 38, 42, 49, 126, 152, 159–160, 163–166, 193 n. 3, 199, 200 n. 7, 208, 211, 214, 216, 219, 221–224, 226, 335, 376 Sumur, Nubra Valley  36, 63, 367, 378 Sūrya  251 Suvarnagotra  84 susumṇā  237

index Tabo (lTa bo, lTa pho)  42, 167, 176–177, 178 n. 15, 180–182, 203, 224, 231, 269, 367, 378, 379 n. 17 Taklung Thangpa chenpo (sTag lung Thang pa chen po)  179 n. 18 Taktsang Repa (sTag tshang Ras pa)  91, 259 Tangtse, petroglyphs  18, 20, 23, 28, 30, 60, 271 Tantra  275, 276 n. 5, 324–325, 327 Tārā  267, 273, 325 Tashi Challi  340 n. 13 Tashi Namgyal (bKra shis rnam rgyal) 160, 165–166, 169 n. 3, 350 Tathagata (Tathāgata)  109 n. 7, 282, 293 Te’u ser po see Tiseru Tegin  188 Temisgam  71, 73, 166 Temisgang (gTing mo sgang)  262 Temur  188 Tengri Nor  96–97 textiles  9, 154, 182, 335–336 luxury  331, 334, 338 patterns  335 thab sa  356, 370, 372, 373 n. 7, 379 thang ka  179 n. 18, 239, 292, 320, 331 The Kingdoms of Gu.ge Pu.hrang  93, 186 Thikse (Khrig se, Khri rtse)  7, 13, 87, 109, 113, 126 n. 14, 129, 138–140, 144, 227 n. 5, 228, 356, 392 Thok Jalung  84, 96 Tholing (mTho lding)  167 (Map 7.1), 178 nn. 15, 17, 181 n. 22, 182 n. 29, 187 n. 33, 188, 201, 208, 210 n. 19, 211, 231 Thubstan Paldan (Thub bstan dPal ldan)  229–230 Tibet Heritage Fund, THF  10, 49, 53, 117, 268, 348, 393 Tie-resist dyeing  335 n. 5 Tilopa  324, 326–327 Timber lacing  8, 264–266 Tingmosgang  44, 133 n. 17, 337 Tirith, Nubra Valley  36, 63 Tiseru (Te’u ser po)  233, 392 tourism  3, 351, 353, 359, 363, 393–394, 397 trade  5–6, 9, 32, 35, 39–40, 44, 48, 58, 60–61, 86, 95, 99, 191 n. 9, 299 n. 3, 331, 335–340, 342, 344–345, 349, 359, 396

437 Tramo Marmo (Pra mo dmar mo)  294 Tramenma (Phra men ma)  286 Transoxiana  170 Treaty of Tingmosgang  337 Trebeck, George  9, 89, 338 tribhaṅga  204 Tsangkhan Malik (lTsang mkhan ma lig) 255 Tsede  90, 93–95, 98 Tsering Angchok, Basgo  394 Tsering Malik (Tshe ring ma lig)  8, 256–259, 266 Tseuri (Tse’u ri)  286 (Fig. 11.11) Tsha tsha Phu ri  232 Tshazar  73 Tshe dbang rnam rgyal  90–91 Tshe la dbang ba’i rig ’dzin  294 Tshe ring dbang ’dus  321 tshogs zhing  9, 314–315, 316 n. 3, 320–324, 326–328 Tshul khrims ’od  159, 163–164, 222 Tshul khrims rdo rje  92 Tshurbu  96–97 Tsong kha pa  226–227, 229–231, 236, 239, 315, 316 n. 3, 320 n. 6, 321–322, 323 n. 10, 324 Tucci, Giuseppe  3–4, 16, 75, 100, 101, 109 n. 6, 148, 193 n. 3, 210 n. 19, 229 n. 7, 276, 285 n. 15, 286 n. 15, 287 n. 16 Tumel (Kargil)  36, 40, 42 Turkestan  1, 298 Turkic funerary monuments  172 traditions  170 tribal confederation  170 Turko-Islamic written sources  170 Turks  170–171, 180 n. 20, 186–190 turquoises  178 Üchang  96–98 Uḍḍiyāna  235 n. 17 Uighurs  99, 170 Yellow  96 United Nations Educational, Scientific and Cultural Organisation, UNESCO  362, 390, 394, 398 Utpala, king  176 Vairochana (Vairocana)  293 Vaiśravaṇa (Tib. rNam thos sras or rNam sras) 253, 325

438 vajra  127, 156 Vajra Heruka  285 (Fig. 11.9) Vajrabhairava (Tib. rDo rje ’jigs byed)  228 n. 5, 233–237, 315, 324, 326 Vajradhara (Vajradhāra)  242, 316, 321–322, 323 n. 9, 324 Vajradharma  325–326 Vajrapāṇi  61, 144 n. 7, 325 Vajravidhāraṇā  326 Varanasi  9–10, 329, 331, 336, 338–340, 342, 344–347 Vasubandhu  320, 326 Vāyu (Tib. Rlung lha)  251 Vemacitrin (Tib. Thags bzang ris)  246 Venice Charter  390 vestibule  191, 193, 219, 221, 223 Vets, Hilde  277, 297 Vidyadhara (Vidyādhara)  279, 284, 293 vihāra  188 Virudhaka (Virūḍhaka)  326 Virūpākṣa  325 Viṣṇu  250 Vogel, Jean Philippe  303–304 Wanla (Wan la, Wan le)  71, 126, 165, 188 n. 36, 211 n. 22, 212, 214, 216, 223, 232 n. 12, 266, 376, 380, 383, 394 Wara (Wa ra)  9 n. 2 Western Himalayas  65, 128, 140 Winkler, Jakob  276 n. 5 wisdom lights  282, 293 woodcarvings  194, 209, 214, 224 wool  337 World Heritage Sites  390 World Monuments Fund, WMF  267 World Trade Organisation, WTO  390

index wrathful deities  276 nn. 5, 7, 278 n. 12, 279–280, 282, 284, 285 n. 15, 286, 294 Xinjiang  1, 170, 171 n. 9, 180 n. 20, 298, 359 Xuan Zang, Xuanzang  86, 183 Yamī  237–239 Yama (Lord of Death)  235, 237–239, 248 n. 23, 287 Yamāntaka  235, 237 Yamarāja  237, 239 Yamāri  235 n. 17 Yarkand  40, 63, 84–85, 88, 336, 338, 359 Yarlung dynasty  98, 169, 180 n. 19, 349 Yasin  40, 188 n. 35 Ye shes mkha’ ’gro ma  236 Yeshé Ö (Ye shes ’od)  5 Ye shes seng ge  236 yi dam  239, 314–316, 324–326 Yogācāra  316, 323 n. 10, 326–327 Yogatantra  164 Yongle  226 n. 4 Zangla  73, 101, 123, 133, 138–140, 394 Zangskar, Zanskar (Zangs dkar)  16, 18, 30, 65, 73, 84–85, 88, 101, 105 n. 4, 191 n. 1, 234, 254, 265, 298–299, 378 Zhang zhung  38, 84, 87 Zhi ba’i gShin rje  237 zhi khro  235, 276, 278 n. 12, 280, 285 n. 15, 287 n. 16 Zhiwa Marmo (Zhi ba dmar mo)  294 Ziling Tsho  96 Zorawar Singh  8, 259, 299, 305