German Military Geology and Fortification of the British Channel Islands During World War II [1st ed.] 9783319227672, 9783319227689

This volume discusses how the German armed forces made effective use of military geologists to assist their fortificatio

446 114 29MB

English Pages XIV, 406 [415] Year 2020

Report DMCA / Copyright

DOWNLOAD FILE

Polecaj historie

German Military Geology and Fortification of the British Channel Islands During World War II [1st ed.]
 9783319227672, 9783319227689

Table of contents :
Front Matter ....Pages i-xiv
Introduction (Edward P. F. Rose, John T. Renouf)....Pages 1-24
Geology (John T. Renouf, Edward P. F. Rose)....Pages 25-80
Geologists and the German Armed Forces (Edward P. F. Rose, Dierk Willig)....Pages 81-105
Jersey and the German Army (Edward P. F. Rose)....Pages 107-161
Jersey and the German Air Force (Edward P. F. Rose)....Pages 163-197
Guernsey and the German Army (Edward P. F. Rose)....Pages 199-254
Guernsey and the German Air Force (Edward P. F. Rose)....Pages 255-288
Alderney (Edward P. F. Rose)....Pages 289-334
Groundwater Investigations: German and British (Nicholas S. Robins)....Pages 335-356
Conclusion: Contemporary Context and Postwar Legacy (Edward P. F. Rose)....Pages 357-393
Back Matter ....Pages 395-406

Citation preview

Advances in Military Geosciences

Edward P. F. Rose  Editor

German Military Geology and Fortification of the British Channel Islands During World War II

Advances in Military Geosciences

Series Editors Peter Doyle London South Bank University, London, United Kingdom Judy Ehlen Haytor, Devon, United Kingdom Francis Galgano G67 Mendel Science Center, Villanova University, Villanova, PA, USA Russell Harmon North Carolina State University, Raleigh, NC, USA Edward P. F. Rose Royal Holloway, University of London, Egham, Surrey, United Kingdom

Advances in Military Geosciences is a book series which explores the interaction between current and historic military operations and earth science, including geography, geology, geophysics, soil science, ecology, hydrology, glaciology and atmospheric sciences. Military activities are almost always strongly integrated within a wide spectrum of geoscience. The decisive outcomes of land battles throughout history have been dictated in large part by the terrain and environmental setting. Modern military operations rely on a wide range of land-, air-, sea-, and space-borne intelligence and knowledge of dynamic terrain processes and conditions. In addition, the study of geo-based environmental science is critical to both the sustainable management of military reservations and installations, as well as the evaluation of how terrain and environmental conditions may impact military equipment and operations. Advances in Military Geosciences contains single and multi-authored books as well as edited volumes. Series Editors are currently accepting proposals, forms for which can be obtained from the publisher, Zachary Romano (Zachary.Romano@ springer.com). More information about this series at http://www.springer.com/series/15030

Edward P. F. Rose Editor

German Military Geology and Fortification of the British Channel Islands During World War II

Editor Edward P. F. Rose Department of Earth Sciences Royal Holloway, University of London Egham, Surrey, UK

ISSN 2522-8315     ISSN 2522-8323 (electronic) Advances in Military Geosciences ISBN 978-3-319-22767-2    ISBN 978-3-319-22768-9 (eBook) https://doi.org/10.1007/978-3-319-22768-9 © Springer Nature Switzerland AG 2020 This work is subject to copyright. All rights are reserved by the Publisher, whether the whole or part of the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations, recitation, broadcasting, reproduction on microfilms or in any other physical way, and transmission or information storage and retrieval, electronic adaptation, computer software, or by similar or dissimilar methodology now known or hereafter developed. The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication does not imply, even in the absence of a specific statement, that such names are exempt from the relevant protective laws and regulations and therefore free for general use. The publisher, the authors, and the editors are safe to assume that the advice and information in this book are believed to be true and accurate at the date of publication. Neither the publisher nor the authors or the editors give a warranty, expressed or implied, with respect to the material contained herein or for any errors or omissions that may have been made. The publisher remains neutral with regard to jurisdictional claims in published maps and institutional affiliations. This Springer imprint is published by the registered company Springer Nature Switzerland AG The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland

Preface

Major advances in military geosciences were stimulated by the two world wars. That is especially true for German armed forces during World War II, who made military use of some 400 geologists: by far the largest number by any nation in wartime, ever. However, the lessons learnt are largely concealed in unpublished reports now dispersed in archives within the UK and the USA as well as Germany, and these reports are easily readable only by people with some understanding of the German language as well as geology. This book brings together information from the disparate sources to provide a case history in English. It illustrates the kind of geoscientists that German forces used as ‘military geologists’, and what could be expected of them in wartime. In doing so, it focuses on a unique region: the Channel Islands—the only part of the British Isles to be occupied by German forces during World War II. After the war and until the end of the twentieth century, it was generally believed that only two German geologists had served on the Islands, and that all their reports had been destroyed prior to the final surrender. However, this book demonstrates that this was a misconception: at least 14 men in total are now known to have been used on the Channel Islands to provide professional expertise as geologists within the German armed forces, and over 50 of their reports have survived within Germany, the USA, and the UK. For the UK and France, the war had begun on 3 September 1939. Germany had invaded Poland on 1 September, and when its Führer Adolf Hitler rejected an ultimatum by the British and French governments demanding that the invading troops be withdrawn, they declared war. A British Expeditionary Force began moving to France on 4 September, to take up positions in the NE: a defensive line to guard against German attack through neutral Belgium. During the months of the ‘Phoney War’ that followed, the opposing sides built up their military might. The attack finally came on 10 May 1940. Despite spirited resistance, the British Expeditionary Force was driven back to the Channel coast and, between 27 May and 4 June, mostly evacuated from the vicinity of the port of Dunkirk. French and remaining British troops fought on until German victory in the Battle of France was conceded, and the French government was compelled to sign an armistice with that of Germany on 22 v

vi

Preface

June. Deemed indefensible in the face of overwhelming German might, the Channel Islands, close to the Normandy coast of France, were demilitarized and partly evacuated. German troops took possession of undefended Guernsey on 30th June, Jersey on 1st July, and Alderney on 2nd July 1940. As explained in Chap. 1 of this book, German forces had thereby seized an area of great scenic beauty, whose long history had led to the development of a unique island culture. Although English was the local language, the Islands were not legally part of England or indeed the UK. Originally part of the ancient Duchy of Normandy, then as now they had their own distinctive systems of government, issued their own Sterling banknotes, and issued their own postage stamps—governance, currency, and postage differing even between the major islands. German forces had to control both people and terrain of a distinctive character—and had access to a considerable legacy of fortifications constructed by the British in earlier centuries, intended to provide protection from potential invasion by the French. The rocks that are an obvious feature along most of the island coasts are similar to those that may be seen in nearby France, in Normandy and Brittany. They are mostly of great age, some formed at least 2000 million years Before Present, and many associated with a period of mountain building (the ‘Cadomian Orogeny’) that took place some 600 million years Before Present, for which evidence is found nowhere else in the British Isles. The British Geological Survey has therefore described the Channel Islands amongst its ‘classical areas of British geology’. Chapter 2 describes the long history of geological studies on the Islands that culminated in a considerable pool of knowledge available to German forces when they began their period of occupation. Although the early months of occupation were relatively uneventful, the situation changed from March 1941 onwards, as German aggressive might became focused against the Soviet Union. In June 1941, Hitler decreed that the Channel Island garrison was to be increased to a full division, to counter any potential attempt by the British to recapture the Islands whilst his main forces attacked eastwards. Moreover, he issued two programmes for the Islands’ fortification, one to last 14 months, the other 7 years. Later that year, on 20 October, he issued a directive that permanent fortification should be pressed forward energetically, to create an impregnable fortress. His intention was that the Islands would forever remain a fortified outpost of the German state, much as Gibraltar, a rocky peninsula jutting south from Spain at the western entrance to the Mediterranean Sea, then was for the UK. On 15 December, this was followed by a construction order for the Atlantic Wall: intermittent coastal fortifications that stretched from Norway in the north to the border of France with Spain in the south, to create a defensive western boundary for German-occupied Europe, including the Channel Islands. Chapter 3 describes how German armed forces were organized to apply their considerable engineering and geological skills to this formidable task. The geology of the Islands had an important bearing on fortification. Geological history had been a formative influence on Island surface features, and these in turn influenced the selection of many military construction sites. The best sources for quarrying construction materials, such as crushed stone for making concrete and

Preface

vii

sand for cement, were determined by geologists. Geologists also advised on where best to site wells or infiltration galleries to make use of groundwater for secure water supplies—at a time of unusually low rainfall. German armed forces made operational use of far more military geologists during World War II than their British and American opponents worldwide. Indeed, although between 1941 and 1943 the German Army and the German Air Force in total made use of at least 14 uniformed geologists on the Channel Islands, during this period the British Army made use of only three military geologist staff officers in total, in all of its theatres of operation. Records of German geological activity were amongst the military documents systematically destroyed prior to surrender of the Islands at the end of the war, on 9 May 1945. However, duplicate copies of geological reports, many illustrated by thematic maps, have been discovered in recent years within archives outside the Islands. These reveal the names of the geologists involved and the nature of their work. Chapters 4 and 5 describe and illustrate how the work of military geologists from the Army and Air Force, respectively, contributed to the fortification programme on Jersey, Chaps. 6 and 7 rather differently to fortification on Guernsey, and Chap. 8 to fortification on Alderney. One of the legacies of this geological work was a detailed record of groundwater conditions on the Islands during the occupation, especially for the year 1942. Nothing like this had been attempted on the Islands during the previous centuries, nor had it been attempted in such detail in other ‘hard rock’ areas of the British Isles as a whole. As explained in Chap. 9, the German studies have provided benchmark information that may be used, together with information gathered during postwar studies by the British Geological Survey, to generate a relatively long-term understanding of groundwater conditions on the Islands—essential for forward planning of adequate water supplies as the Islands are further developed as a popular holiday resort, and the world faces challenges of climate change and changing rainfall patterns. In conclusion, Chap. 10 sets the German military geological work on the Channel Islands in a broader context. It briefly describes similar Atlantic Wall fortifications on the Normandy coast nearby, fortifications that, unlike those on the Islands, were bombarded and many of them extensively damaged prior to or during the Allied invasion that began on D-Day, 6 June 1944. Since Alderney has long been known as the ‘Gibraltar of the Channel’, it compares and contrasts Channel Island fortification with coeval British fortification of Gibraltar, especially with regard to the use of geology and geologists to facilitate military construction works. It briefly reviews the wartime use of military geologists by the British Army, as individual staff officers in field force general headquarters, from 1940 within a team formed by the South African Engineer Corps, and from 1943  in two teams for terrain analysis remote from operational areas, one in England and the other in India. Finally, it describes wartime use of geologists in the United States Army, as a Military Geology Unit of the US Geological Survey, its many members primarily although not exclusively located near Washington DC. However, although the US Army made more extensive use of military geologists than the British Army, their numbers were again small relative those serving with German armed forces: less than 100 actual geologists by the war’s end.

viii

Preface

A legacy of German engineering work as a whole is the range of fortifications that still exist on the Islands. The Allied landings in Normandy that began on D-Day bypassed the Channel Islands. The Islands and their fortifications were finally surrendered at the end of the war, untested and intact. Much of the metalwork was soon removed in postwar years for recycling, but the concrete structures mostly remain, and representatives of the many types have now been conserved, sometimes refurbished, to serve as tourist attractions. Some of those attractions are now housed in wartime underground facilities originally designed to provide accommodation and storage safe from aerial or naval bombardment. Almost all the sites illustrated in this book may freely be viewed from the outside, if only from a distance, but many sites are open, at least periodically, to paying visitors: a remarkably compact and accessible testimony to the programme of intense fortification, for which details of admission are now readily available online. The book thus complements the many other books that describe the German occupation of the Channel Islands from the perspective of the people involved, either the British islanders or the German occupiers, or by reference to military events or military engineering works. It is an account, written by professional geologists of different but very relevant expertise, that focuses on geology and geologists and on their importance to the construction of this unique, supposedly impregnable, fortress. Published in 2020 to help mark both the 80th anniversary of the start of the German occupation and the 75th anniversary of its end, it is hoped that the book will be of interest not only to geoscientists in general but also to the residents of the Channel Islands, in helping them to understand factors that influenced the construction of some of the fortifications now mellowing into the natural landscape; to people interested in the history of World War II, whether from the German or Allied side, because of the special significance of this ‘impregnable fortress’; and to the many tourists that visit the Islands, to enjoy their unique culture, mild climate, sandy beaches, marine wildlife, long history, fine cuisine—and over 800 years of defensive fortification that peaked during the German occupation. The book draws on contemporary German records now preserved in the archives of the British Geological Survey at Keyworth, near Nottingham, in the UK; the Military Division of the Federal German Archives (Bundesarchiv-Militärarchiv) at Freiburg-im-Breisgau, Germany; the archives at the Geoinformation Centre of the modern German Army (Bundeswehr Geoinformationen Zentrum) at Euskirchen, Germany; and the National Archives and Records Administration at College Park, Maryland, in the USA. As indicated in the text, its core chapters bring together and significantly amplify data that the editor has published, frequently in co-authorship with Dierk Willig, in issues of the Channel Islands Occupation Review and elsewhere, between 2002 and 2014. Footnotes are used to explain geological terms not in common use and to indicate information sources not listed in the references that end each chapter. They are also used to indicate details of German nomenclature and documents that are ­unnecessary for the general reader but which provide the data required by more serious scholars.

Preface

ix

Permission to reproduce some of this material is indicated where appropriate in the captions to individual figures, and so gratefully acknowledged. The chapter authors also acknowledge the kind assistance and encouragement that they have received over many years from members of both the Jersey and Guernsey branches of the now well-established and currently very active Channel Islands Occupation Society, particularly the late Michael Ginns, and the Society’s ‘Review’ editors Matthew Costard, Trevor Davenport, and Paul and Iain Ronane. John Renouf kindly read drafts of Chaps. 4 and 5 for Jersey, Pierre Renier Chaps. 6 and 7 for Guernsey, and Trevor Davenport Chap. 8 for Alderney, but any remaining errors are the responsibility of the sole author in each case. Wendy Cawthorne, Assistant Librarian at the Geological Society of London, was particularly helpful in gaining access in England to some of the more obscure bibliographic references and biographical information, and Professor Hermann Häusler of the University of Vienna helpful in providing such information from sources in Austria and Germany. Grateful thanks are due to all of these kind colleagues. My co-authors have read chapters in this book additional to their own and made helpfully constructive comments; the Springer editors have guided preparation of the book in its earliest and final stages; but the final responsibility for any errors or omissions remains entirely my own. Egham, UK  Edward P. F. Rose

Contents

1 Introduction����������������������������������������������������������������������������������������������    1 Edward P. F. Rose and John T. Renouf 2 Geology������������������������������������������������������������������������������������������������������   25 John T. Renouf and Edward P. F. Rose 3 Geologists and the German Armed Forces��������������������������������������������   81 Edward P. F. Rose and Dierk Willig 4 Jersey and the German Army ����������������������������������������������������������������  107 Edward P. F. Rose 5 Jersey and the German Air Force����������������������������������������������������������  163 Edward P. F. Rose 6 Guernsey and the German Army ����������������������������������������������������������  199 Edward P. F. Rose 7 Guernsey and the German Air Force ����������������������������������������������������  255 Edward P. F. Rose 8 Alderney����������������������������������������������������������������������������������������������������  289 Edward P. F. Rose 9 Groundwater Investigations: German and British ������������������������������  335 Nicholas S. Robins 10 Conclusion: Contemporary Context and Postwar Legacy������������������  357 Edward P. F. Rose Index������������������������������������������������������������������������������������������������������������������  395

xi

About the Contributors

John  T.  Renouf  Born on Jersey, John Renouf began research studies in the University of London by a PhD thesis on the older rocks of western Brittany in France. During a subsequent Jersey-based career, first at the Jersey Museum and subsequently to retirement at the Department of Education, his research interests widened to take in Channel Islands’ geology as it can be used to enhance archaeological and historical work and, particularly, to more recent geology in the islands and the significance of past sea levels above and below that of the present. Nicholas S. Robins  Nick Robins is a graduate of the University of Southampton, but his degrees include an MSc in hydrogeology and a DSc, both from the University of Birmingham. He is author or co-author of over 100 technical papers and articles concerning aspects of hydrogeology of the UK, the Channel Islands, and elsewhere, notably Sub-Saharan Africa and the Caribbean; he is also author/editor of nine books on a variety of hydrogeological topics. Now working part-time as a hydrogeologist, on retirement from the British Geological Survey, his other recent activities include the role of Editor-in-Chief for the two book series published by the International Association of Hydrogeologists. Edward P. F. Rose  Ted Rose graduated from Oxford University with a ‘first’ in geology and a doctorate for a thesis on Libyan echinoid palaeontology and Cenozoic stratigraphy, before lecturing on geology at London University: at Bedford College from 1966, at Royal Holloway from 1985 until retirement from a senior lectureship to an honorary research fellowship in 2003. Commissioned into the Territorial Army in 1962 via the Oxford University Officers Training Corps, he served from 1969 to 1990 as a military geologist (from 1974 senior military geologist) in the Royal Engineers (Volunteers), from 1987 to 1990 in the rank of colonel. He is author or co-author of over 120 book chapters or journal articles on the history of military applications of geology (additional to publications on echinoid palaeontology and the geology of Gibraltar); he received the Sue Tyler Friedman Medal of the Geological Society of London in 2014.

xiii

xiv

About the Contributors

Dierk  Willig  Dierk Willig graduated from the University of Würzburg (Dr. rer. nat.) for a thesis on the history of military applications of geology. He has developed a career as a civilian military geologist within the modern German Army and also a strong record of publication on topics relating to historical aspects of military geology. He leads the Geology/Hydrology/Geophysics Division at the Bundeswehr Geoinformation Centre of the German Army and holds the reserve army rank of lieutenant colonel.

Chapter 1

Introduction Edward P. F. Rose and John T. Renouf

Abstract  The Channel Islands, British since 1204, lie close to the Normandy coast of northern France. They have a temperate marine climate and rural rather than urban landscapes with many features attractive to tourists. The largest islands (Jersey, Guernsey, Alderney and Sark) are fringed with spectacular coastal scenery and dotted with sites of prehistoric or historic interest, the latter principally fortifications to deter invasion. A small Roman fort on Alderney, probably constructed in the fourth century AD, is the earliest of these to be stone-built. British fortification to deter invasion from France began with the construction of castles on Jersey and Guernsey between the thirteenth and seventeenth centuries. It continued during the seventeenth to early nineteenth centuries with the development of coastal redoubts for batteries of artillery and towers to be defended by small arms or cannon fire. It culminated in a massive programme of perimeter forts constructed on Alderney during the mid nineteenth century: part of arguably the greatest programme of national fortification in the British Isles. All of these fortifications, being massive in construction and strategic in their position, were to varying degrees adapted by German forces after their seizure of the islands in the summer of 1940, to help counter potential recapture by the British.

1.1  Regional Setting The Channel Islands are British Crown Dependencies situated in the south of the English Channel, over 100  km from England and some 25 to 50  km from the Normandy coast of France (Fig. 1.1). Seven islands are permanently inhabited: the E. P. F. Rose (*) Department of Earth Sciences, Royal Holloway, University of London, Egham, Surrey, UK e-mail: [email protected] J. T. Renouf Le Cotil des Pelles, La Route du Petit Port, St. Brelade, Jersey e-mail: [email protected] © Springer Nature Switzerland AG 2020 E. P. F. Rose (ed.), German Military Geology and Fortification of the British Channel Islands During World War II, Advances in Military Geosciences, https://doi.org/10.1007/978-3-319-22768-9_1

1

2

E. P. F. Rose and J. T. Renouf

Fig. 1.1  Location of the largest Channel Islands relative to England and France

principal islands of Jersey (118  km2), Guernsey (65  km2), Alderney (8  km2) and Sark (5 km2), and the lesser islands of Herm (2 km2), Brecqhou (0.3 km2) and Jethou (0.2  km2). Total resident population is currently about 166,000, of whom most (about 105,000 people) live on the largest island, Jersey, and of these about 33,500 in its main town, St. Helier.1 In 1939, before the start of World War II, there were only about 50,000 people on Jersey, 40,000 on Guernsey, 1500 on Alderney and 600 on Sark (Cruickshank 1975). There are also adjacent islets, rocks and reefs that are normally uninhabited: the Minquiers, Écréhous, Dirouilles and the Paternosters near Jersey; Caquorobert, Crevichon, Grande Amfroque, Les Houmets and Lihou near Guernsey; and Burhou, the Casquets, Ortac and Renonquet near Alderney. Characteristic features of Channel Island geography, history and culture have been extensively described in many earlier books (e.g. Coysh 1977), so much infor-

 www.gov.je/Leisure/Jersey/pages/profile.aspx, last accessed 12 January 2019.

1

1 Introduction

3

mation is widely available. Notably, Channel Island laws and administrative systems are distinct from those of England, and therefore the United Kingdom as a whole, having been independently derived from the medieval Duchy of Normandy. Although English is the local language, a few older inhabitants still converse in a Norman French patois. The islands are divided into two almost wholly self-governing bailiwicks:2 Jersey and Guernsey, each including its own offshore reefs but with the bailiwick of Guernsey also embracing Alderney and Sark. The last two islands have a measure of self-government but ultimately defer to Guernsey. The main islands issue their own postage stamps, banknotes and coinage (both notes and coinage being interchangeable locally with British currency). The economy is traditionally agricultural, noted both for dairy cattle and products and for the growing of vegetables. Tourism has always been important and continues to be so but has been overtaken by banking and finance. The warmest and sunniest of the British Isles, the geography and history of the Channel Islands have given them a unique character. When, after the defeat of France and its allies early in World War II, German forces began their occupation of the islands in the summer of 1940, they found that they had arrived in a region very different to their own home country. Both the islands and their 1940–45 German occupation have given rise to an extensive literature: Gardiner (1998) provides a comprehensive list of publications up to almost the end of the twentieth century, and more have been published subsequently.

1.2  Geography: The Physical Background As will be described in Chap. 2, it has long been recognized that all of the islands are formed by bedrock that is very old and very strong, and more closely similar to the rocks that can be found nearby in northern France than across the Channel in England. The islands themselves contain no bedrock younger than the Cambro– Ordovician, although rocks formed in the vast 450-million-year time interval since then are well represented in adjacent France. The low dissected plateaux that characterize the islands’ surface are erosional outliers of an ancient peneplain that begins at the coast of Normandy (Jee 1982; Renouf 1993). Jersey’s main plateau forms a major feature of its land surface. It slopes gently down (from a height that averages about 100  m above sea level along the north coast) to a height of some 60 to 50 m along the south coast, where abrupt steep slopes descend to the sea (Fig. 1.2). A series of major, deeply incised valleys, principally eroded by more or less north to south flowing rivers, is now occupied by  Bailiwick: the district or jurisdiction of an official known as a bailie or bailiff. In the Channel Islands, each of the two bailiwicks has a lieutenant-governor appointed by and representing the British Crown, and a bailiff who is the bailiwick’s most important citizen: traditionally the presiding officer of its local government (the States, to which members are elected) and the head of its judiciary. 2

4

E. P. F. Rose and J. T. Renouf

Fig. 1.2  Map of Jersey showing major features of relief. Modified by J.T. Renouf from a hachured map created by N.L.V. Rybot in the 1930s

small streams. Etched into the steep coastal cliffs and the gently sloping plateau above are what may be termed a staircase of flatter, bench-like breaks of slope, all of which may represent erosion reflecting a complex story of interplay between climatic and tectonic events and associated sea-level changes over the last several million years (Jones et al. 1990; Renouf and James 2010). Jersey’s shape is strongly influenced by geology. The NW, SW and SE corners of the island are formed by granite masses that are highly resistant to erosion. In contrast, the NE corner is formed of conglomerate: a sedimentary rock that is more easily eroded. The major embayments that are a feature of the western and southern coasts have been formed by erosion of the slightly metamorphosed, fractured and cleaved (and so relatively weak) Precambrian sedimentary rocks that otherwise form much of the island (Fig. 1.2). The high and steep cliffs that form the northern coast are, apart from the NW granite, eroded into a great variety of variably resistant igneous rocks, many of volcanic origin. Guernsey is less regular in outline than Jersey and is overall triangular rather than rectangular in shape (Fig. 1.1). The main plateau area occupies the base of the ­triangle in the south (where the coastal cliffs are steepest) and, in contrast to Jersey, slopes from south to north. However, its average height is similar. It declines from a maximum of about 100 m above sea level in the SE to end at a steep inland scarp, trending approximately west-east, that descends to the narrowing northern sector of the island: a region of much lower and more varied relief. No geological formation

1 Introduction

5

has been dominant in controlling the pattern of erosion and coastal embayments are the result of complex processes. However, the low-lying northern part of Guernsey has been significantly shaped by the effects of climate and sea-level changes during Quaternary time. Indeed, a major part of northern Guernsey was for a time a separate island, until the intervening strait was reclaimed from the sea in the 1800s. The small island of Sark, to the east of Guernsey (Fig. 1.1), is the most plateau-­ like of all the Channel Islands. It has an essentially flat surface at about 100 m in height, and is surrounded on all sides by steep to vertical cliffs. Alderney is also an inclined cliff-bounded plateau, but facing east (Figs 1.1 and 1.3). Much of the eastern lowland area is underlain by sandstone similar in age of formation to the conglomerate on Jersey, and similar in its response to erosion. High tidal ranges, between 6 and 14  m, are a feature of the coasts within the Normanno-Breton Gulf area in which the Channel Islands are situated (cf. Fig. 1.1). In consequence, extensive rocky wave-cut platforms or wide sandy beaches are exposed in many areas at low tide, but submerged at high tide level.

Fig. 1.3  Aerial view of Alderney, from 5000 ft (about 1500 m), showing the steep cliffs fringing the SW coast to the right of the image; the main town (St. Anne) in the centre; and the gentle regional downward slope of the surface plateau towards the major harbour of Braye Bay (partly enclosed by a nineteenth century breakwater) and the NE lowland of the island (top left of image). From an original print by Richards of Alderney, and reproduced by kind permission

6

E. P. F. Rose and J. T. Renouf

Climate is temperate maritime, lacking marked variations or extremes.3 At latitude 50 deg north, the islands lie essentially within the mild westerly airflow that typically separates the sub-tropical high-pressure zone to the south and the more northerly course of Atlantic low pressure systems. Sheltered to some extent by the much greater landmasses of England to the north and France to the south and east, the islands escape the worst of the winds often associated with offshore locations. The climate is thus typified by mild damp cloudy winters and warm drier sunnier summers. Jersey averages a mean daily air temperature of 11.5  °C, experiences ground frosts on only 60 days per year and has snow or sleet on only 12 days per year. The average long-term annual rainfall is 877 mm on Jersey and 790 mm on Guernsey, but actual annual rainfall may vary considerably from the long-term mean (see Chap. 9). The bedrock of the largest islands is covered by a variable thickness of loess: a wind-blown silt deposited during late Quaternary time (and so comparable with the ‘brick earth’ known from southern England and the ‘limon des plâteaux’ of northern France). Climate and this geological stratum in particular have together generated fertile soils (Jee 1982), as described for Jersey in particular by Jones et al. (1990). The islands are still predominantly rural, as they were in 1940, with intensively worked arable land and grassland used for dairy farming, plus small areas of heathland, semi-natural mixed deciduous woodland, and wetland. Palaeobotanical records reveal that, from about 10,000 years ago, periglacial tundra was succeeded by a vegetation mosaic in which woodland was important, its composition varying with changes in climate and sea level until reaching a fully temperate vegetation cover by about 8000 years before present. Not surprisingly, given its close proximity, the present biota has closest affinity with neighbouring France. Apart from birds, the native fauna is relatively restricted in variety, with few mammals or larger reptiles, although the islands host the most northerly occurrence of some southern European plants and contain some unique subspecies of small mammals.

1.3  Archaeology and History Each of the islands would have been accessible to ancient humans at times during the last million years from what is now France without the use of boats: when sea level was low because of cyclic climate (Jones et  al. 1990; Sebire 2011; Renouf 2015). Oldest habitation is currently known from Jersey, from a Palaeolithic (Old Stone Age) cave shelter at La Cotte de St. Brelade. Fossil remains and human artefacts from this well documented archaeological site have been dated to between 250,000 and 50,000 years before present and ascribed to habitation by Neanderthals (Homo neanderthalensis).

 www.metoffice.gov.gg/climate.htm last accessed 11 January 2019.

3

1 Introduction

7

The first modern humans (Homo sapiens), associated with a Magdalenian culture, are also currently known only from Jersey. Evidence of their presence has been dated to about 14,000 years before present, a time following the end of the extremely cold climate of the Last Glacial Maximum. Following the end of the last Pleistocene glacial period (and so of the Quaternary ‘Ice Age’) at about 11,000 years before present, the subsequent post-glacial warming led to rising sea levels and so progressive flooding of the former cold climate landscape. Each of the islands in turn became isolated by the sea from the adjacent Normandy coast and nearby islands. Temperate climates prevailed over the whole Normanno-Breton Gulf area by about 9000 years before present and there is evidence for Mesolithic habitation on the islands then, followed by extensive occupation by Neolithic farming communities from about 8000 years ago. It was these Neolithic peoples who exploited the geological resources of the islands to a significantly greater extent, using stone for domestic purposes (fashioning it into axes, mullers for grinding cereals, and ornaments) and constructing passage graves and other massive stone (megalithic) monuments. Trade beyond the islands became increasingly important from the Neolithic onwards, but this was trade based initially within the prevailing societies of NW France—of which the island communities were an integral part—and only later to areas further away, such as southern England across the Channel. Towards the close of the Iron Age, during the first century before the Christian era, the Celtic tribes of France (then known as Gaul) became subject to Roman conquest. Thereafter, for almost four centuries, Gaul—including its NW area near the Channel Islands, given the Celtic-derived name of Armorica by the Romans—became increasingly Romanized, and its culture consequently Gallo-Roman. Cross Channel trading was important in the Roman empire, which included England (Britannia) from the first century of the Christian era. The final collapse of the Gallo-Roman realm in the fourth century was followed by several centuries of instability with much movement of peoples. This culminated in sea-borne raids by Norsemen (= ‘northmen’, from Scandinavia) into northern France and Armorica to the west early in the ninth century, followed later by settlement. In 911 Norsemen founded what became the Duchy of Normandy, named after them, and soon the Channel Islands were annexed to the Duchy and ruled by its Duke. From the conquest of England by Duke William of Normandy (who became King William I of England) in the year 1066 through to 1204, Normandy, as well as other areas of France, owed allegiance to the English monarch. However, in 1204— the most critical date in the islands’ history—King John of England was forced to cede all of Normandy except the Channel Islands to the King of France, Philippe Augustus. Thereafter the islands retained allegiance to the Crown in England and the scene was set for the islands to endure what would become six hundred years of intermittent strife with France. The military consequences of this were profound: because the islands always faced potential attack by the French, they were ­progressively fortified by the English. A few strong castles were developed on Jersey and Guernsey from the thirteenth to seventeenth centuries. More numerous and widespread redoubts and batteries appeared within the eighteenth century with

8

E. P. F. Rose and J. T. Renouf

a series of coastal towers towards its end. Further coastal towers were constructed to the middle of the nineteenth century along with major forts in Jersey and Guernsey. British fortification ended with a massive programme of perimeter forts constructed on Alderney, part of arguably the greatest programme of national fortification in the British Isles.

1.4  British Fortification of Jersey Jersey was to be defended by a series of British coastal fortifications constructed between the thirteenth and the nineteenth centuries (Fig. 1.4). The earliest of Jersey’s stone-built fortifications is Mont Orgueil Castle (Fig. 1.5), built on a granite promontory overlooking the wide sweep of the Royal Bay of Grouville and facing the coastline of the Cotentin, that part of Normandy from the northern tip of the Cherbourg peninsula to the port of Granville in the south (cf. Fig. 1.1). Erosion of the isolated boss of granite had produced steep slopes and high cliffs on three sides of the promontory that offered a naturally defended site. The castle was founded in the early thirteenth century and progressively strengthened thereafter to counter developments in weapon technology, and perceived threats from France. Its role as a fortress lapsed from the late sixteenth century but, between 1942 and 1944, German occupying forces developed it as a self-contained strong-

Fig. 1.4  Map of Jersey indicating sites of British-built thirteenth to nineteenth century fortifications. After Rose et al. (2002)

1 Introduction

9

Fig. 1.5  Mont Orgueil Castle (cf. Fig. 1.4), viewed from the south, sited on a granite promontory and overlooking the village of Gorey at the northern end of Grouville Bay. The flat tops on the three turrets of the mediaeval castle are German concrete additions to convert the turrets to artillery fire control and observation posts (M7 on Fig. 4.3). From Wikimedia Commons, file Gorey_and_ Mount_Orgeuil_from_the_south.jpg, reproduced under the terms of the GNU Free Documentation License (CC BY-SA 3.0)

point. It was adapted to provide coastal artillery fire control and observation towers, dugouts and trenches to provide shelter from bombardment of its infantry garrison, and positions for automatic and other small arms. Twenty-four ‘roll bombs’ were placed on the outer walls and a flame-thrower at the top of the steps leading to the keep. The NE outworks were heavily reinforced to house a battle headquarters. An electricity system was installed and the water supply improved. Large rooms in the keep were converted into fully furnished barracks (Rybot 1978; Rose et al. 2002; Ford 2007). Grosnez Castle (Fig. 1.4) was built in the early fourteenth century atop granite cliffs at the NW point of Jersey. Before the invention of firearms it provided a place of refuge for the population in the west of the island, but had no secure water supply to enable it to withstand siege, and no resident garrison. It was easily (if briefly) captured by the French during a raid in 1373, and had fallen into ruin by the sixteenth century (Balleine 1950; Syvret and Stevens 1981). From the late sixteenth century Mont Orgueil was superseded as the principal island fortress by Elizabeth Castle (Figs. 1.4 and 1.6), constructed on a granite-like (granophyre-diorite) tidal island facing Jersey’s principal port, St. Helier. With the development of cannon, Mont Orgueil was no longer a safe refuge in time of conflict, despite re-modelling of its defences. The first part of a new castle (the keep, known as the Upper Ward), adapted to contemporary fighting methods, was largely con-

10

E. P. F. Rose and J. T. Renouf

Fig. 1.6  Elizabeth Castle (cf. Fig. 1.4), sited on an igneous rock (granophyre-diorite) island facing Jersey’s main port, St. Helier. View south showing beneath the British flag the massive concrete fire control tower constructed by German forces on the castle’s keep. From Wikimedia Commons, file Elizabeth_Castle.jpg, reproduced under the terms of the Creative Commons Attribution 2.0 Generic License (Jon [CC BY 2.0 (https://creativecommons.org/licenses/by/2.0)], via Wikimedia Commons)

structed between 1594 and 1601, during the reign of the English Queen Elizabeth the First. By 1668 construction had progressively extended to form the Lower and Outer Wards. With some re-modelling, Elizabeth Castle retained a British garrison until 1923, when it was sold to the Jersey government for use as an historic monument. However, German forces returned it to its former use. In the Outer Ward, three light machine guns were installed, together with a 105  mm calibre gun protected by a bomb-proof casemate, plus other bomb-proof shelters and searchlight positions. Similar installations were constructed in the Lower Ward. The keep was in part adapted for the then modern weapons, modified to provide increased barrack accommodation, and surmounted by a massive concrete artillery fire control tower (Partridge 1976, p. 136; Rybot 1986; Ford 2008). In the eighteenth and nineteenth centuries coastal redoubts and batteries were constructed to defend those bays and beaches deemed vulnerable to sea-borne invasion and, although long disused, several were adapted and re-activated by German troops. Examples are the two redoubts of Grouville Common on Jersey’s east coast: Fort William and Fort Henry (Fig. 1.4). Fort William (built originally in 1760) was strengthened with concrete emplacements and heavy machine guns. Fort Henry (an eighteenth century infantry base: Fig. 1.7) was similarly converted to a ‘resistance nest’—with two 105 mm calibre guns in concrete emplacements, two 50 mm calibre mortar emplacements connected by trenches to the fort, two light and three heavy machine guns, ten flame-throwers, one anti-aircraft gun and four searchlights, manned by an officer and 35 troops of other ranks (Ginns 1973).

1 Introduction

11

Fig. 1.7  Fort Henry (cf. Fig. 1.4), viewed from inland. In 1781 known as Fort Conway and the barracks for the 93rd Regiment of Foot, the site was re-fortified and manned by German troops during the years of occupation. From Wikimedia Commons, file Fort Henry, Jersey.jpg, reproduced under the terms of the Creative Commons Attribution-Share Alike 3.0 Unported License (Danrok [CC BY-SA 3.0 (https://creativecommons.org/licenses/by-sa/3.0)], from Wikimedia Commons)

In 1778, when the French became the allies of the American Colonies in their war for independence from the United Kingdom, the governor of Jersey was certain that the French would seize the opportunity to attack the Channel Islands. He therefore proposed the construction of 30 coastal towers to resist enemy landings (Pocock 1971; Grimsley 1988; Davies 1991)—round towers, now known as Jersey Towers, about 10 m high and 500 m apart, the walls pierced with loopholes for musketry in two stages (Fig. 1.8). They were designed for musket fire because of the shortage of cannon at that time. Later, at the onset of the Napoleonic Wars, each tower roof was reinforced and received a 12-pounder carronade (a short, smoothbore, cast-iron cannon), on a pivot mount that allowed 360 deg of rotation. Twenty-two round towers and one square tower were eventually built by 1798; a further three round (Martello)4 towers between 1807 and 1814; and five more between 1834 and 1838 (e.g. Fig. 1.9)—31 towers in total. Jersey contains both the earliest and the latest exam-

 Inspired by a tower at Mortella Point on the Mediterranean island of Corsica that resisted considerable British naval bombardment in 1794, coastal towers of this type were built in England and across the British Empire during the nineteenth century. Round towers with thick masonry walls, Martello towers were resistant to cannon fire and their height (about 12 m) made them an ideal platform for a single heavy gun. 4

12

E. P. F. Rose and J. T. Renouf

Fig. 1.8  Le Hocq tower (number 18 on Fig. 1.4), a typical example of the Jersey round towers completed between 1780 and 1798 during the threat of invasion from revolutionary France, built initially to defend potential landing areas (at a time when there was a shortage of artillery) by musket fire from loopholes and the roof. Floor openings in the parts which jut out at the top of the tower enabled the defenders to fire straight down the walls whilst protected from enemy fire, and so prevent the enemy from undermining the tower. Photo by Paweł “pbm” Szubert, from Wikimedia Commons, file Jersey_Le_Hocq_Tower_01.jpg, reproduced under the terms of the Creative Commons Attribution-Share Alike 3.0 Unported License

ples of such British-built fortifications in Europe—which by 1838 fringed all but the steep and so amphibiously inaccessible northern coast of the island (Fig. 1.4). All but four of these granite-built towers were still standing at the time of the German occupation. Three were destroyed by the Germans to make way for other works, but many were modified for contemporary use (see Partridge 1976, p. 134).

1 Introduction

13

Fig. 1.9  Kempt tower (number 4 on Fig. 1.4), one of the five ‘Martello’ towers constructed on Jersey between 1834 and 1838. Elliptical rather than round in plan, with its thickest wall towards the sea, this contained a trefoil gun platform—and was designed for defence by artillery rather than by musketry. The door inserted at ground level was part of later adaptation for German use. The original design was for mid-level access by means of wooden steps, as also shown. These could be burnt if the tower was attacked, to deny an easy means of access to the attacker. Photo: E.P.F. Rose

The last British fortress to be constructed on Jersey, Fort Regent, was founded in 1806, when defensive works throughout the island were in hand to counter the threat from Napoleonic France, and completed in 1814 just as the threat came to an end (Davies 1971). By this time Elizabeth Castle was considered to be too isolated to be of much value, except as a refuge of last resort, and the extensive new fort was constructed on the granite-like rocky promontory overlooking St. Helier (Fig. 1.4). It was garrisoned by the British regular army until 1932, but from that date onward no longer regularly occupied. Ironically, the only shots ever fired in anger from the fort were by German occupying troops against Allied aircraft during World War II.

1.5  British Fortification of Guernsey On Guernsey, the first stone-built fortification was also founded in the thirteenth century: Castle Cornet, sited on a rocky granitic (granodiorite) islet facing the island’s capital town, St. Peter Port (Fig. 1.10), from which it was accessible on foot at low tide. Built initially between 1206 and 1256, Castle Cornet (Fig. 1.11) was developed through later centuries as the island’s principal fortress (Barton 2003; Guernsey

14

E. P. F. Rose and J. T. Renouf

Fig. 1.10  Map of Guernsey indicating sites of existing British-built thirteenth to nineteenth century fortifications. After Robins et al. (2012)

Museum Team 2008; Rose et al. 2012). The first structure comprised a keep, chapel, two courtyards and curtain walls. However, the island was captured by French troops in 1338 and the castle’s defences were improved prior to recapture by the English in 1345. Further and more major re-modelling took place between 1545 and 1548, and later that century, to counter improvements generally in artillery, but the keep and part of the living quarters were catastrophically destroyed in 1672 when a strike by lightning detonated the castle’s gunpowder magazine. Upgraded during the French Revolutionary and Napoleonic Wars of 1793–1815 with additional barracks, the castle later became a prison. During World War II it was occupied by a small garrison of German troops, and known as strongpoint ‘Hafenschloss’ (‘harbour fortress’). German modifications were made in concrete rather than stone as formerly, to adapt it to contemporary warfare.

1 Introduction

15

Fig. 1.11  Castle Cornet, built on a granodiorite island facing St. Peter Port, Guernsey’s principal town (cf. Fig. 1.10). From Wikimedia Commons, file Castle_Cornet_2009_b.jpg, released into the public domain

Fig. 1.12  Aerial view the Château des Marais, also known as Ivy Castle (cf. Fig.  1.10). From Green (2002), courtesy of Mark Lewis of the Guernsey Press and Star, per Richard Digard

The Château des Marais (Figs. 1.10 and 1.12) or the ‘castle of the marshes’, also known as Ivy Castle, was another stone-walled foundation of the early thirteenth century (Barton 1981). Constructed in the north of Guernsey, it provided a refuge for the local population potentially safe from attack in the marshland. It was built in classic ‘motte-and-bailey’ form, typical of the period, with the top of an enclosed mound flattened, and a surrounding ditch dug out with the soil used to level and raise a system of inner and outer defences. These structures were encircled by walls on new, solid ground, as the adjacent marsh had been drained by a channel cut towards the sea coast to the east. It was Guernsey’s principal fortified site for 20–30 years before the focal point shifted to Castle Cornet as the island’s main defensive

16

E. P. F. Rose and J. T. Renouf

Fig. 1.13  Entrance to Vale Castle, near Guernsey’s NE coast (cf. Fig. 1.10). Photo: E.P.F. Rose

structure about 1250. Late in the eighteenth century, at the time of wars with France, the castle was re-fortified and the magazine and much of the existing stonework belong to this period. It later fell into decay, becoming an ivy-covered ruin. However, during World War II, it was occupied by German troops who built a bunker within its inner walls, and machine-gun posts and communication trenches which destroyed much of the archaeological evidence of the earlier motte-and-bailey structure. Vale Castle (Figs. 1.10 and 1.13) overlooks and protects the port of St. Sampson towards the NE end of the island (Barton 1989) and has spectacular views of the nearby islands of Herm and Jethou. A stone castle was founded on the site of an Iron Age earthwork, curtain walls, a gatehouse and buttresses being constructed from granite in the fifteenth and sixteenth centuries, and barracks added in the eighteenth century (although the barracks were later demolished). During World War II, German forces fortified both the castle and its surrounding area, erecting a large gun position on top of an earlier battery and several machine-gun posts in and around the castle. On Guernsey, artillery batteries on stone-built foundations existed on several headlands (particularly on the east coast) during the seventeenth century, and 15 round towers of distinctive local style were built in 1778 or shortly thereafter (Fig. 1.14). Three Martello towers (each mounting at least one 24-pounder carronade, to counter naval attack) were built in 1804. These three towers formed the nucleus of more extensive fortifications and artillery batteries, developing into three of the island’s nine forts (Fig. 1.15). Fort George, completed in 1812 near Guernsey’s main town St. Peter Port (Fig. 1.10), was built to become the main British military headquarters for the island. Three of the round towers were later destroyed, as were four of the round towers on Jersey, but all these British fortifications were thick-­

1 Introduction

17

Fig. 1.14  One of Guernsey’s surviving 12 round towers, constructed in 1778–79 from local stone to deter possible French attacks. The towers are numbered sequentially, counter-­clockwise from St. Peter Port on the east coast. That shown is number 4, near Fort Le Marchant in the NE of the island (cf. Fig. 1.10). Built to one design, they allowed musket fire to cover all approaches to the tower. Similar to the earlier towers constructed on Jersey (see Fig. 1.8), they differed in several details, notably in possessing a slope at the base but in lacking projections for downward fire at the top. Photo: E.P.F. Rose

Fig. 1.15  Fort Hommet, on a headland overlooking Vazon Bay on Guernsey’s western coast (cf. Fig. 1.10). Built on the site of fortifications that date back to 1680, this consists of a central stone-­ built Martello tower (cf. Fig. 1.9), constructed in 1804, together with the mid nineteenth century barracks and batteries that later extended it into a fort. Concrete fortifications were added during the German occupation of the island, including a casemate for a 105 mm calibre gun, to transform the headland into Strongpoint ‘Rotenstein’. Photo: E.P.F. Rose

18

E. P. F. Rose and J. T. Renouf

walled structures strategically sited to provide defence against a sea-borne invasion, and so were adapted by German forces to form part of their own system of coastal defence.

1.6  British Fortification of Alderney Alderney boasts arguably the best-preserved small Roman fort in the British Isles (Monaghan 2011). Known locally as ‘the Nunnery’, it lies at the foot of the hill beneath Essex Castle (Fig. 1.16) near the western end of Longis Bay (whose anchorage it was possibly intended to protect) and is the earliest evidence for military construction on the island. Its shape (30 m by 30 m square, with rounded towers at the corners) much resembles that of five signal-station forts of Roman date on the NE coast of England (Fig. 1.17), so its original purpose may have been to help guide shipping rather than to protect the harbour. A building of stone and Roman concrete, dating from about AD 350, it was adapted by the British during the Napoleonic Wars, and later fortified by the Germans with two 47 mm calibre anti-tank guns and concrete bunkers to form a small but powerful ‘resistance nest’. The island generally was not fortified until much later than the two larger islands (Partridge and Davenport 1993). Work was begun on its Essex Castle (Figs. 1.16 and 1.17) in 1546, but soon abandoned, as English government policy changed. The present ‘castle’ is a barracks (later converted to use as a hospital) built in the 1840s, so is also and more appropriately known as Fort Essex. This too was fortified by the Germans during World War II, and known by them as ‘Burg Essex’. During the eighteenth century, the island was protected by over a dozen artillery batteries sited along its low-lying northern coast (Fig. 1.16), the region most vulnerable to amphibious assault. However, unlike Jersey and Guernsey, no towers were built to enhance coastal defence. In the 1840s the French created a strongly fortified naval base at Cherbourg (Fig. 1.1) on their Normandy coast and, some 40 km to the west, the British countered this by construction of a ‘harbour of refuge and observation’ at Braye Bay on Alderney (States of Alderney 1963; Partridge and Davenport 1993). This was complemented by a series of forts (e.g. Fig. 1.18) constructed between 1850 and 1858 on all but the southern, cliff-bounded and so inaccessible coast of the island. These provided a girdle of sites with overlapping arcs of fire to protect the remaining 65% of Alderney’s perimeter. Fort Albert, overlooking Braye Bay, was developed as the site of main defence. This series of fortifications was thus more substantial and more recent than military works on Guernsey and Jersey, where construction of new harbours was planned (on Guernsey) or begun but never completed (on Jersey at St. Catherine’s Bay, north of Mont Orgueil Castle). All the forts except the first, Grosnez, were designed by William Jervois, then a Royal Engineers captain but later to become Lieutenant-General Sir William Jervois (Kinross 2004). In 1860 (as a young major) he was to be appointed ‘design leader’ for a programme of fortifications on the United Kingdom mainland to defend against a potential French inva-

1 Introduction

19

Fig. 1.16  Map of Alderney showing sites of British-built thirteenth to nineteenth century fortifications. After Robins et al. (2012)

Fig. 1.17  View SW along part of Longis Bay, on Alderney’s NE coast (cf. Fig. 1.16). The Nunnery, originally a Roman fort but later much altered and partly re-fortified in concrete during the German occupation, occupies the foreground to the right. Essex Castle, founded in the mid sixteenth century but converted into a barracks in the nineteenth and also partly re-fortified in concrete during the German occupation, covers the hilltop in the distant centre. Photo: J.T. Renouf

20

E. P. F. Rose and J. T. Renouf

Fig. 1.18  Château à l’Étoc, the mid nineteenth century perimeter fort on the most northern point of Alderney (cf. Fig. 1.16), viewed from the west. During World War II German forces sited a battery of 20 mm anti-aircraft guns at this location, which they consequently re-named Flakbatterie ‘Einsiedlerschloss’ (Hermit’s Castle). Rooms are now in use as private apartments. Photo: J.T. Renouf

sion: arguably the largest system of fortifications that the British Isles had ever seen (Crick 2012). They were a major step in a career of fortress construction and public service that would see the works of Jervois spread across the world, from the United Kingdom to North America, Bermuda, India, Australia and New Zealand. Strongly built and strategically sited, the Alderney forts (e.g. Fort Tourgis: Fig. 1.19) proved amenable to adaptation and enhancement during the German occupation.

1.7  Conclusion: The Arrival of German Armed Forces Impressive as the British coastal fortifications on the larger islands were (and the smaller islands were not important enough to merit any fortification), they did not deter German forces in the summer of 1940. German forces seized the islands from the air rather than the sea, to begin a five-year occupation that has been described and/or illustrated in an extensive literature (e.g. Cruickshank 1975; Pantcheff 1981; Ramsey 1981; Forty 1999; Stephenson 2006). Major events in World War II that provide the background to this operation have also been extensively documented (e.g. by Dear and Foot 1995). The United Kingdom and France had declared war on Germany on 3 September 1939, when the German Führer Adolf Hitler had refused to withdraw his troops invading eastwards into Poland. However, there was no invasion westwards into France. The ‘Western Front’ remained relatively quiet until May 1940: a period that became known as the ‘Phoney War’. During this time major British–German hostilities took place at sea rather than on land. A British Expeditionary Force had begun moving to northern

1 Introduction

21

Fig. 1.19  Fort Tourgis on the NW coast of Alderney (cf. Fig. 1.16), completed in 1855 and the second strongest of the island’s British forts, was designed to accommodate 346 men, 33 heavy cannon and four 13-inch mortars. It commanded Platte Saline Bay to the east and Clonque Bay to the west. German forces sited a battery of 20 mm anti-aircraft guns in the highest part of the fort, which they adapted to form Strongpoint ‘Turkenburg’ (Turk’s Castle) armed with two 105 mm calibre guns (both in casemates) plus two 75 mm guns sited just above the beach in bunkers outside the fort, together with two or three searchlights and numerous machine-gun and other weapon positions (see Davenport 2003, pp 34 and 35, for plan and description). From Green and Digard (1985), courtesy of Mark Lewis of the Guernsey Press and Star, per Richard Digard

France from 4 September, but this was purely a defensive force. It lacked the resources to take offensive action. Initially formed by some 152,000 troops, by May 1940 the Force had grown to nearly 400,000. Yet of these, more than 150,000 were in rear areas and mostly without military training. Life in the Channel Islands was remote from the area of potential combat, and carried on very much as normal. However, on 10 May 1940 German forces launched an attack westward, across France and the Low Countries. The Allies were staggeringly defeated in less than six weeks, the British Expeditionary Force mostly evacuated via Dunkirk and nearby beaches in late May and early June, and the French government forced to sign an armistice with Germany on 22 June. As German troops swept across France towards Normandy and Brittany, the decision was taken in London that since the Channel Islands had no strategic importance and could not be defended from a power in control of the French mainland, they should be demilitarized and surrendered.

22

E. P. F. Rose and J. T. Renouf

With defeat and occupation looming, the British authorities made an offer on 19 June to evacuate the civilian population from the Channel Islands, although it was obvious that the shipping potentially available would be inadequate for all islanders. In Jersey, 23,000 registered for evacuation but only about 6000 people actually left (a mere 13% of the population). In Guernsey, about 17,000 people (40% of the population) chose to leave and were actually evacuated. In Alderney, all but seven people (so almost 100% of the population) had left the island by 22 June. British troops who had reached the islands from France were amongst the first to be withdrawn to England, on 20 June. On the next day the lieutenant-governors of the two bailiwicks, as representatives of the Crown, were also withdrawn, leaving the bailiffs to head the remaining administrations. Unfortunately, news of this demilitarization was not sent to the German forces now occupying France. On 28 June they duly launched air raids on the major towns of St. Helier in Jersey and St. Peter Port in Guernsey, killing 44 civilians. This prompted the British government to announce via the British Broadcasting Corporation that the islands had already been demilitarized. Two days later, the Foreign Office in London asked the Ambassador of the United States of America, as the representative of a major power at that time still neutral in the war, to inform the German government in Berlin officially that the British had indeed evacuated all military personnel and equipment from the Channel Islands some days previously. German forces arrived on Guernsey on 30 June 1940. Seeing its airport apparently deserted, the pilot of a reconnaissance aircraft made a brief test landing, leaving the other three aircraft in his flight to circle protectively overhead. Following his report that the airfield was undefended, a platoon of German Air Force (Luftwaffe) personnel was flown in by transport aircraft, to take possession. That evening, the senior German officer announced to the Bailiff that the island was now under German occupation. Jersey surrendered on 1 July, after a summons to do so dropped from the air was followed by another speculative Luftwaffe pilot landing on a seemingly peaceful island. More planes from the pilot’s squadron quickly followed, packed with men to take control. Alderney was occupied from 2 July, but the troops initially sent by air had to abort their intended landing because of obstacles placed on the airfield by the British prior to evacuation. Sark surrendered to a small force sent by sea from Guernsey on 4 July. The first German troops to arrive on Guernsey by sea did so on 14 July: two units of the Luftwaffe’s anti-aircraft artillery. However, as the Germans consolidated their position, reinforcements arrived from the German Army, from Infanterieregiment 396 of 216 Infanteriedivision, who were shuttled to Guernsey by air while other units arrived on the islands by sea. Units of the German Army and of the German Navy were soon in occupation as well as the Luftwaffe. The islands were to be held by part of an infantry division, later by a reinforced infantry division, and to endure German occupation until the war in Europe had officially ended, on 8 May 1945, and a British liberating force had arrived, on 9 May. From the summer of 1941, German military might became increasingly focused eastwards against the Union of Soviet Socialist Republics (‘Russia’), reducing manpower in western regions and so leaving the western boundary of German-occupied Europe potentially vulnerable to Allied raids or invasion. German forces therefore

1 Introduction

23

initiated a massive programme of coastal fortification—the so-called Atlantic Wall—from northern Norway south to the border of France with Spain, in order to deter potential attack. To deny any attempt to recapture the Channel Islands by the British, they took advantage not only of the pre-existing fortifications left on the islands but (as described next, in Chap. 2) a considerable legacy of geological knowledge. The Germans were to make use of considerable engineering resources for construction work (Chap. 3), and from effectively the start of their fortification programme, they appointed a military geologist to serve as such on the largest island, Jersey: Lieutenant (later Captain) Walther Klüpfel (Chap. 4). At least 13 other German geologists were to follow (Chaps. 5, 6, 7, 8, and 9), and to generate a record of work that has implications even for the present day (Chap. 9).

References Balleine GR (1950) A history of the Island of Jersey. Staples, London Barton KJ (1981) Excavations at the Château des Marais (Ivy Castle), Guernsey. In: Report and transactions of the Société Guernesiaise, vol 20. La Société Guernesiaise, Guernsey, pp 657–702 Barton KJ (1989) The principal fortifications of the Channel Islands before 1750. In: Fortress: the castles and fortifications quarterly, vol 1. Beaufort Publishing, England, pp 24–32 Barton KJ (2003) The archaeology of Castle Cornet. Guernsey Museum Monograph No. 7. Guernsey Museum, St. Peter Port Coysh V (ed) (1977) The Channel Islands: a new study. David & Charles, Newton Abbot Crick T (2012) Ramparts of Empire: the fortifications of Sir William Jervois, Royal Engineer 1821–1897. University of Exeter Press, Exeter Cruickshank CG (1975) The German occupation of the Channel Islands. Oxford University Press, London Davies W (1971) Fort Regent: a history. Private publication, St. Helier, Jersey Davies W (1991) The coastal towers of Jersey. Société Jersiaise, Jersey Dear ICB, Foot MRD (eds) (1995) The Oxford companion to the Second World War. Oxford University Press, Oxford Ford D (2007) Mont Orgueil castle. Jersey Heritage Trust, St. Helier, Jersey Ford D (2008) Elizabeth castle. Jersey Heritage Trust, St. Helier, Jersey Forty G (1999) Channel Islands at war: a German perspective. Allan Publishing, Shepperton, Surrey Gardiner V (1998) The Channel Islands. World bibliographical series, vol 209. Clio Press, Oxford Ginns M (1973) Grouville common during the German occupation. Annu Bull Soc Jersiaise 21:194–199 Green B (2002) The Bailiwick of Guernsey from the air. Guernsey Press, Guernsey Green B, Digard R (1985) Islands in focus: an aerial appreciation of the Bailiwick of Guernsey. Guernsey Press, Guernsey Grimsley EJ (1988) The historical development of the Martello Tower in the Channel Islands. Sarnian Publications, Guernsey Guernsey Museum Team (2008) Castle Cornet. Guernsey Museums & Galleries, Guernsey Jee N (1982) Landscape of the Channel Islands. Phillimore, London Jones RL, Keen DH, Birnie JF, Waton PV (1990) Past landscapes of Jersey: environmental changes during the last ten thousand years. Société Jersiaise, Jersey

24

E. P. F. Rose and J. T. Renouf

Kinross JS (2004) Jervois, Sir William Francis Drummond (1821–1897). In: Matthew HGC, Harrison B (eds) Oxford dictionary of national biography, vol 30. Oxford University Press, Oxford, pp 81–83 Monaghan J (2011) Alderney: a new Roman fort? Curr Archaeol 261:28–33 Pantcheff TXH (1981) Alderney, fortress island: the Germans in Alderney, 1940–1945. Phillimore, Chichester Partridge C (1976) Hitler’s Atlantic Wall. D.I. Publications, Guernsey Partridge C, Davenport T (1993) The fortifications of Alderney. Alderney Publishers, Alderney Pocock HRS (1971) Jersey’s Martello towers. Annu Bull Soc Jersiaise 20:289–298 Ramsey WG (1981) The war in the Channel Islands: then and now. Battle of Britain Prints International Limited, London Renouf JT (1993) Solid geology and tectonic background. In: Keen DH (ed) Quaternary of Jersey: field guide. Quaternary Research Association, London, pp 1–11 Renouf JT (2015) La Cotte de St. Brelade: a new survey of its importance and for its conservation. Annu Bull Soc Jersiaise 31:431–461 Renouf JT, James HCL (2010) High level shore features of Jersey (Channel Islands) and adjacent areas. Quat Int 231:62–77 Robins NS, Rose EPF, Cheney CS (2012) Basement hydrogeology and fortification of the Channel Islands: legacies of British and German military engineering. In: Rose EPF, Mather JD (eds) Military aspects of hydrogeology, vol 362. Geological Society, Special Publications, London, pp 203–222 Rose EPF, Ginns WM, Renouf JT (2002) Fortification of island terrain: Second World War German military engineering on the Channel island of Jersey, a classic area of British geology. In: Doyle P, Bennett MR (eds) Fields of battle: terrain in military history. Kluwer Academic Publishers, Dordrecht, pp 265–309 Rybot NLV (1978) Gorey castle, 2nd edn. States of Jersey, Jersey Rybot NLV (1986) The Islet of St. Helier and Elizabeth Castle, 9th edn. Société Jersiaise, Jersey Sebire H (2011) The archaeology and early history of the Channel Islands. The History Press, Stroud States of Alderney Publicity and Entertainments (ed) (1963) A short history of and guide to Alderney. States of Alderney, Alderney Stephenson C (2006) The Channel Islands 1941–45: Hitler’s impregnable fortress. Osprey, Oxford Syvret M, Stevens J (1981) Balleine’s history of Jersey. Phillimore, Chichester

Chapter 2

Geology

John T. Renouf and Edward P. F. Rose

Abstract Historically, the British Channel Islands lay outside the scope of the national geological surveys of both Great Britain and France but were studied by a succession of geologists, British, French, and local, from at least 1811 onwards. These established that the islands reveal a terrain different from other regions of the British Isles and form part of the Armorican Massif, of Precambrian to Paleozoic age, that forms much of Lower Normandy and Brittany in nearby France. This region comprises metamorphic and igneous rocks together with a range of Paleozoic strata, the bedrock having a thin and patchy superficial cover of Quaternary sediments. Rocks of Mesozoic and Tertiary age are largely absent. When they seized the islands in 1940, German forces had access locally and via university libraries in Germany and France to a wide range of published information that revealed a developing understanding of the kinds of rocks to be found on the islands, and the times and processes involved in their formation. The distribution of the major rock types had been mapped at different scales and detail in the different islands, but provided a basis of ‘pure’ geology that could be enhanced and adapted to meet German military engineering requirements.

2.1  Introduction When German forces arrived on the Channel Islands in 1940, they came to what has now come to be regarded as one of the ‘classical areas of British geology’ (Institute of Geological Sciences 1982; British Geological Survey 1986; Bishop and Bisson 1989): an area whose geology has stimulated research since the early nineteenth J. T. Renouf (*) Le Cotil des Pelles, La Route du Petit Port, St. Brelade, Jersey e-mail: [email protected] E. P. F. Rose Department of Earth Sciences, Royal Holloway, University of London, Egham, Surrey, UK e-mail: [email protected] © Springer Nature Switzerland AG 2020 E. P. F. Rose (ed.), German Military Geology and Fortification of the British Channel Islands During World War II, Advances in Military Geosciences, https://doi.org/10.1007/978-3-319-22768-9_2

25

26

J. T. Renouf and E. P. F. Rose

century. Although the main islands were politically British, geographically and geologically all the Islands formed part of adjacent France: the region known to geologists as Armorica.1 By 1940, research had established that the island’s magnificent coastal rock exposures preserve a record of sedimentary, metamorphic, tectonic, and igneous events, dating mostly from Precambrian and early Paleozoic times, that can be more closely related to the early geological history of nearby Brittany and Normandy than to England, north of the Channel. The islands, like the rest of Armorica, reveal evidence for a period of mountain building, the Cadomian Orogeny,2 much older than the Caledonian3 and later orogenic periods important in England. Moreover, because the islands lay south of the ice sheets that intermittently covered much of Britain during Quaternary (Pleistocene) times, their superficial deposits permit interesting comparisons and contrasts with deposits of similar age in southern England and NW France. Although the geology of the Channel Islands had excited interest amongst geologists, both British and French, for a remarkably long time, lying outside the United Kingdom as such, the main islands lay outside the scope of the Geological Survey of Great Britain. They were not covered by any of its surveys or maps prior to the war. In addition, lying close to France and so necessarily included with parts of the continent on some of its government-sponsored geological maps, as ‘foreign’ territory the islands were not accessible to official French geological surveys. Geologically, they thus occupied a ‘no-man’s land’ between the regions covered by the two long-established national geological surveys, those of Great Britain and France. In this void, geological research was initiated and published by: (1) individuals from Britain, either completely in their private capacity or resulting from their academic (e.g. MacCulloch 1811; Parkinson 1907) or industrial affiliations (e.g. Transon 1851); (2) those domiciled in the islands or closely linked to them (e.g. Noury 1886; Mourant 1933a), with either British or French geological backgrounds; or (3) those individuals from France with an interest in the geology of the islands as a natural extension of the geology of their neighbouring French region (e.g. Bigot 1888). This geological knowledge developed intermittently for over 120 years prior to the German invasion: a case history illustrative of changes in both interests and understanding in earth science over an unusually long term. These studies generated maps and literature relating to ‘pure’ geology that formed the basis for the ‘applied’

 Armorica was a part of ancient Gaul, including the region of modern Brittany, known by this name in Roman times. 2  The Cadomian Orogeny was named by Bertrand (1921) from Cadomus, the Roman name for the city of Caen in Normandy. 3  Named after Caledonia, the Roman name for Scotland, the Caledonian Orogeny was a period of mountain building—involving metamorphic, tectonic, and igneous events associated with ocean closure—that affected rocks now found in northern Britain and Ireland, Scandinavia and elsewhere, from Ordovician to Early Devonian times and so about 490–390 million years ago. 1

2 Geology

27

geological surveys and maps that were a distinctive achievement of German military geologists during the war.

2.2  Armorica The Channel Islands’ archipelago became recognized both geographically and geologically as part of the Armorican Massif: a part of western France where the Precambrian and Paleozoic basement crops out either because it has never been covered by Mesozoic–Cenozoic sediments or because these sediments have been removed by erosion, a region partly concealed beneath the sea of the Normanno-­ Breton Gulf (Figs. 2.1 and 2.2). Armorica is separated from England by the English Channel: it lies almost opposite the Cornubian Massif of Cornwall-Devon, a region essentially of Paleozoic rocks flanked to the east by the Mesozoic–Cenozoic sediments of the Wessex Basin. The Armorican Massif is likewise flanked to the east mostly by Mesozoic-Cenozoic sediments, in this case forming the Paris Basin. Similar colouring on early geological maps of SW England and NW France (e.g. Fig. 2.1) encouraged the belief that their rocks and geological histories were similar, but research in recent decades (e.g. Renouf 1974) has demonstrated that they are in fact very different. The Massif is easily recognizable to visitors because of its distinctive relief and vegetation, but these features essentially derive from its underlying geology. Some knowledge of that geology dates as far back as geological mapping in the early part of the nineteenth century. This pioneering work was later amplified, in particular by impressive studies by two French geologists: Charles Barrois (in Brittany) and Alexandre Bigot (in Normandy) in the late nineteenth and early twentieth centuries. Armorica was shown to be one of a number of Paleozoic massifs in central and western Europe (e.g. the Cornubian Massif of SW England, and the Massif Central in the middle of southern France). By the time German forces arrived in France in 1940, most Armorican rocks could be dated in relative terms if not yet in millions of years. It had been established that their formation could be interpreted principally in terms of nine successive geological events (Table 2.1): 1. In Guernsey, Sark, Alderney, and the northern Cotentin peninsula of adjacent Normandy, outcrops of high-grade metamorphic rocks (gneisses and schists) were agreed to represent the oldest rocks in the region and had been shown to be of Precambrian age from field relationships near Cap de la Hague in Normandy. 2. Rocks unaffected by this high-grade metamorphism had obviously formed later, and there must have been a significant time gap between formation of the crystalline basement and deposition of younger rocks upon it. 3. A thick sequence of weakly metamorphosed sedimentary rocks, named the Schistes or Phyllades de Saint-Lô from their early described outcrop near the town of St. Lô in Normandy, was known to succeed the crystalline basement in

28

J. T. Renouf and E. P. F. Rose

Fig. 2.1  NW sheet of the 1:1000,000 scale geological map of France, published by the War Office in 1943; part of GSGS No. 4452, reprinted from the Carte géologique de la France, third edition, of 1933. The legend (on the SW sheet only) shows that the rocks from the Cherbourg (Cotentin) peninsula westwards across Brittany are predominantly very old (Precambrian in age: coloured pink) and very strong (granites: coloured red), with a complex pattern of outcrop; Paleozoic rocks (brown) are subordinate but important. These constitute the ‘Armorican Massif’. In contrast, the rocks to the east are much younger and weaker, largely mid-Jurassic limestones (blue) to Cretaceous chalks (green). Their simple pattern of outcrop, concentric about Paris to the east of the map area, reflects their position on the western margin of the ‘Paris Basin’. Only the thickest deposits of more recent (Pleistocene) age can be shown on a map of this scale, notably those (coloured pale buff) flooring the major river valleys. Reproduced from the Shotton Archive, by permission of the Director of the Lapworth Museum of Geology, University of Birmingham. From Rose et al. (2006)

the Cotentin. Similar rocks occurred on Jersey. In Jersey and very locally in the Cotentin, a thick sequence of volcanic rocks was thought to be possibly near contemporaneous in time of formation with the Phyllades. 4. Metamorphism of the Phyllades and associated rocks, and intrusion of most of the volumetrically significant igneous rocks in the region, had been attributed by

2 Geology

29

Fig. 2.2  Main features of the geological structure of southern England and NW France as understood before World War II (cf. Fig. 2.1). Drawn by J.T. Renouf

Bertrand (1921) to a major period of earth movements (including folding, faulting, and uplift of the earlier rocks) associated with mountain building that he named the Cadomian Orogeny. Uplift led to erosion of the Cadomian mountain chain, eventually to generate a near horizontal surface: a peneplain of very wide extent. 5. Conglomerates and sandstones of a distinctive reddish colour were deposited on this eroded surface, but interpretation of the environment and time of their deposition proved controversial. Some authors inferred a marine rather than terrestrial environment of deposition and, although most authors favoured a late Precambrian or early Paleozoic (Cambrian or Ordovician) date for their deposition, some argued for a later Old Red Sandstone (Devonian) or New Red Sandstone (Permo– Triassic) age. 6. Sedimentary rocks, mostly sandstones and shales with a few limestones of shallow marine origin, of early to late Paleozoic (Ordovician, Silurian, Devonian,

30

J. T. Renouf and E. P. F. Rose

Table 2.1  Significant geological events and their relative ages in the Armorican Massif (including the Channel Islands) as understood in 1940—plus approximate dates in millions of years before present progressively determined postwar (after Cohen et al. 2013, updated) Geological events Raised beach deposits, loess, head, and alluvium Periods of submergence, some folding and vertical movements; Normanno-Breton Gulf mostly emergent or affected by shallow seas Variscan Orogeny ending with period of uplift, erosion, and peneplanation Sediments either eroded from or not deposited in the Channel Island region Deposition of red-beds (conglomerates and sandstones) Cadomian Orogeny with metamorphism and igneous intrusion ending with period of uplift, erosion, and peneplanation Deposition of Phyllades de St-Lô

Geological age Quaternary Mesozoic and Tertiary

Date >2.6 my c.250–2.6 my

Late Paleozoic

c.370– 250 my c.460– 250 my

Late Ordovician– Permian Cambro– Ordovician Late Precambrian and Cambrian Late Precambrian

Major Precambrian interval of uplift and erosion Formation of basement rocks with high-grade metamorphism to produce gneisses

Earlier Precambrian

c.540– 490 my c.650– 490 my c.600– 540 my c.2000– 900 my 2 m thick. Key: 2, gas lock; 3, entrance defence; 4, crew rooms; 6, ammunition room; 6a, ammunition storage; 9, gun emplacement; 10, ventilation; 13, communications room; 18, NCO’s room; 31, central heating; 46, crew shelter; a, ammunition hoist; b, access ladder to gun emplacement; c, entrance; d, generator. Reproduced from the Channel Islands Occupation Society (Jersey) leaflet ‘Batterie Moltke’ by kind permission of the Society

built to ‘fortress’ as contrasted with ‘field’ standard had concrete greater than 2 m rather than 1 m thick. A typical casemate during construction required excavation of 815 m3 of spoil and consumed 730 m3 of concrete, 40 tonnes of reinforcing steel, and 6.2 tonnes of steel girders. Other sites were equipped with 47 mm calibre anti-tank guns (Fig. 4.12), also usually within a standard casemate (e.g. Resistance Points La Carrière and Millbrook: Figs. 4.13 and 4.14). By 1944 Jersey’s armament included about 125 large-calibre weapons deployed to support infantry in a coastal defence role (Table 4.3). Structures associated with infantry defence areas commonly comprised not only casemates for weapons but also personnel and ammunition shelters. A second line of infantry defence installations was planned to support the coastal defences with fire from their main weapons, contain any Allied assaults that might have broken through, and repel them with counter attacks. However, building of this second line of fortification was given

128

E. P. F. Rose

Fig. 4.6  Naval observation tower MP3 (cf. Fig. 4.3) at Les Landes, NW Jersey, viewed from the south, sited at the edge of a granite cliff and with the Island of Sark visible in the distance. Each observation slit in the tower face was intended to permit control of fire from a separate artillery battery. Tower height 16 m. Photographed in 1998, by E.P.F. Rose

a lower priority than that of the coastal defences, and construction developed only in the hinterland of St Ouen’s Bay—the area deemed to be most vulnerable to amphibious attack. Klüpfel’s notebooks do identify the Igel and Jasmin strongpoints near Belle Hougue Point, on Jersey’s northern coast. Jasmin at least was manned from January 1943, but drainage was a problem. Accordingly, Klüpfel made several field inspections, in company with a captain named Mayer.

4.4.3  Anti-Tank Walls and Ditches Sea walls constructed from granite blocks during the latter part of the nineteenth century and the first two decades of the twentieth century to protect the coast from marine erosion were so massively constructed that they needed little if any modification to form a barrier to armoured assault (Ginns 1974). Gaps between the existing walls were, however, filled by nine lengths of concrete wall to form a barrier to tank movements landward of the beach areas lying to the west, south, and east of the island (Fig.  4.15). Altogether 8200  m of anti-tank wall were planned, of which 7397 m were completed. Seldom was any length built to a common plan, but these

4  Jersey and the German Army

129

Fig. 4.7  Plan of Noirmont Tower (MP 1 on Fig. 4.3), built between May and October 1943 by the Organisation Todt, with outer walls of concrete >2 m thick. Reproduced from the Channel Islands Occupation Society (Jersey) leaflet ‘Noirmont Observation Tower’, by kind permission of the Society

130

E. P. F. Rose

Fig. 4.8  Map of Jersey showing the main German infantry defence areas as at 1945. Symbols indicating ‘resistance nests’ grouped to form a ‘strongpoint’ relate to ‘resistance nests’ shown on land nearby: they are not additional defence areas. Modified from Ginns and Bryans (1978) and Baker (undated): reproduced from Rose et al. (2002)

Fig. 4.9  Map of Jersey showing deployment of 105 mm calibre K331(f) coastal defence guns, and casemate type, as at 1945. Modified from Channel Islands Occupation Society (Jersey) leaflet ‘Resistance Point La Carrière’: reproduced from Rose et al. (2002)

4  Jersey and the German Army

131

Fig. 4.10 105 mm K331(f) coastal defence gun and casemate at Strongpoint La Corbière. Photographed in 1998, by E.P.F. Rose

reinforced concrete structures remain impressively massive—rising in places to 6 m in height, with foundations 2  m thick (Fig.  4.16). Inland from St. Ouen’s Bay (Fig. 4.15), a nearly 2-km series of anti-tank ditches was constructed to form an additional barrier.

4.4.4  Minefields Over 100 minefields were laid in Jersey (Gander 1991), away from the populated area of the main town, St. Helier, on the south coast (Fig. 4.17). They supplemented the coastal defences of the western and eastern beach areas of the island, and provided an almost continuous obstacle along its less intensively fortified and garrisoned north coast. Most fields were of anti-personnel mines: over 20,000 Schrapnellminen 35 were emplaced, over 2000 of them tripwired, and over 20,000 Schützenminen 42, many of them improvised (Table  4.4). Anti-tank mines were much less numerous: only some 2000 Tellerminen of various types, plus an

132

E. P. F. Rose

Fig. 4.11  Plan of ‘Channel Islands’ standard casemate for 105 mm K331(f) coastal defence gun. The roof and outer walls are of reinforced concrete >2 m thick. Key: 1, Tobruk/observer’s open position; 2, gas lock; 3, entrance defence; 4, standby room; 5, escape hatch; 6, ammunition room; 7, empty shellcase room; 8, foul air extraction plant; 9, gun room. From Channel Islands Occupation Society (Jersey) leaflet ‘Resistance Point La Carrière’ by kind permission of the Society; reproduced from Rose et al. (2002)

additional 1000 Tellerminen 43 Pilz employed as beach obstacles (probably attached to anti-tank tetrahedra). ‘Roll bombs’ (made to roll down slopes towards an enemy) and other charges improvised from captured or obsolete ammunition were also deployed in a defensive role.

4.4.5  Command and Control Centres The Kernwerk, a complex of six bunkers dispersed to minimize vulnerability to aerial attack, was situated in the approximate geographical centre of Jersey—and close to the airport (Ginns and Bryans 1978; Fig. 4.18). There were three command bunkers (housing fortress, artillery, and infantry headquarters), two communications bunkers, and a bunker for pumping and storing potable water. The three command bunkers were identical in design: built on two levels they incorporated working and living quarters for the staff with wash rooms, flush toilets, and a central heating system. Bunker walls and ceilings were of reinforced concrete greater than

Fig. 4.12  Map of Jersey showing the deployment of 47 mm calibre Pak K36(t) anti-tank guns, and casement type, as at 1945. Modified from Channel Islands Occupation Society (Jersey) leaflet ‘Resistance Point Millbrook’: reproduced from Rose et al. (2002)

Fig. 4.13  Fortress-type casemate for 47 mm Pak K36(t) anti-tank gun, set in a rocky granite headland at l’Oeillere, at the southern end of Jersey’s western coast (see Fig. 4.12). The square metal embrasure that screened the gunners is typical for these installations. Photographed in 2016, by J.T. Renouf

134

E. P. F. Rose

Fig. 4.14  Plan of standard construction Type 631casemate for 47 mm Pak K36(t) ant-tank gun. The roof and outer walls are of reinforced concrete >2 m thick. Key as for Fig. 4.11. Reproduced from Channel Islands Occupation Society (Jersey) leaflet ‘Resistance Point La Carrière’ by kind permission of the Society

Table 4.3  German large-­ calibre weapons for infantry defence deployed on the coast of Jersey during World War II. From Rose et al. (2002), modified from Ginns and Bryans (1978)

Calibre 105 mm 37 mm 47 mm 75 mm 37 mm 50 mm 75 mm 80 mm

Weapon type K331(f) KwK 144(f) Pak 36(t) Pak 40 Pak 35/36 Pak 38 FK 231(f) FK 30(t)

Number deployed 30 30 23 12 14 8 4 4

2  m thick, and command/communications bunkers were disguised externally as dwelling houses with dummy windows, shutters and chimney pots (Ginns 1999). Outside the Kernwerk, command of the Navy’s coastal artillery was exercised from the Noirmont Point Command Bunker (Figs. 4.3, 4.19, and 4.20), and of the north, east, south, and west defence sectors by regional headquarters situated inland.

4  Jersey and the German Army

135

Fig. 4.15  Map of Jersey showing sites of pre-war granite-block sea walls and of concrete anti-tank walls of German construction (after Ginns 1974), plus anti-tank ditches inland from St. Ouen’s Bay (after Baker undated): reproduced from Rose et al. (2002)

Fig. 4.16  Concrete anti-tank wall constructed by the Germans at the south end of St. Ouen’s Bay, viewed from the south, with part of the British stone-built sea wall in the foreground. Note the wide flat beach, shown here at low tide, amenable to amphibious landing. Channel island beaches all have a low–high tidal range of at least eight metres. Photographed in 1998, by E.P.F. Rose

136

E. P. F. Rose

Fig. 4.17  Map of Jersey showing German minefields as at 1945, numbered from Grosnez in clockwise sequence as in Royal Engineer records. Modified from Gander (1991): reproduced from Rose et al. (2002)

Klüpfel is known, according to Bishop and Launert (1979, p. 36), to have been involved with groundwater problems encountered during construction of the ‘headquarters at l’Aleval’ (i.e. the Kernwerk). There is an entry in his notebooks dated 5 August 1942, although construction work had begun somewhat earlier. A boring was made in this area to a depth of 4.3 m, and elsewhere Klüpfel notes that water was found at a depth of 2 m. In February 1943 he observed, after rain, that there was ‘standing water’: presumably a rise in the water table causing flooding of the site.

4.4.6  Anti-Aircraft Batteries Once the decision to fortify the Channel Islands had been taken, anti-aircraft gun batteries multiplied prodigiously. However, these were emplaced and manned by the Air Force, the Luftwaffe, rather than the Army (see Chap. 5).

137

4  Jersey and the German Army

Table 4.4  Mine types and total numbers deployed by the German Army in the Channel Islands during World War II. From Rose et al. (2002), modified from Ginns and Bryans (1978) Mine type Tellerminen Schrapnellminen Schützenminen Improvised (locally manufactured) Stockminen Panzerabwehrminen 407f Panzerabwehrminen 408b 270 mm shells Concealed charges (unspecified) Beach obstacles: Tellerminen 43 Pilz ‘Rollbomben’ Air landing obstacles: Tellerminen 43 Pilz 270 mm shells Unspecified charges

Alderney 1657 6536b 10,611 8310 201 2206 554 43 –

Guernsey 5163a 23,912 16,352 13,851 800 – – 78 –

Jersey 2253 21,127c 20,597d 17,276 1781e 998 1987 – 62

Sark 29 3232 6872 2622 351 – – – –

– –

– –

1075 248

– –

51 176 –

– – –

– 19 460

– – –

Notes Of which 360 were tripwired; b Of which 2204 were fitted with anti-handling devices; c Of which 2319 were tripwired; d Of which 15,922 were improvised; e Of which 1026 were improvised; and f Total including all captured types and may include Panzerabwehrminen 408b a

4.5  Tunnels At least 16 relatively deep underground facilities, in German documents usually referred to as Hohlgangsanlagen (cave passage installations) rather than as Stollen (tunnels) but commonly known to the British on Jersey as ‘tunnels’, were planned to provide accommodation stores for rations, fuel, and ammunition, as well as troop shelters and electricity works (Fig. 4.21 and Table 4.5) (Ginns 1993, 2012). Tunnel excavation was facilitated by the island’s steep-sided narrow valleys that permitted vehicle access to sites with almost immediate potential protective rock cover of up to 36 m—proof against the most prolonged shelling or aerial bombardment. Only two tunnels seem to have been completed (Ho 5 and the defensive tunnel for Strongpoint Etaquerel), but four were in a sufficiently advanced state for partial use (Ho 1, 4, 8, and 19) and four more actively under construction (Ho 2, 10, 13, and 15) by the close of hostilities. Ho 8 (Figs. 4.22 and 4.23) was converted for use as an underground hospital, and was for some years preserved as such as a tourist

138

E. P. F. Rose

Fig. 4.18  Map of Jersey showing German command/control centres as at 1943: modified from Ginns and Bryans (1978) and Baker (undated). The fortress commander was sited adjacent to his artillery and infantry commanders within a command area known as the Kernwerk, NE of the airport. Reproduced from Rose et al. (2002)

Fig. 4.19  The Noirmont Point Command Bunker (cf. Fig. 4.3), mostly concealed beneath ground, viewed from the west. The camouflaged range-finder periscope is visible, and the twin armoured observation turrets. Photographed in 1998, by E.P.F. Rose

4  Jersey and the German Army

139

Fig. 4.20  Plan of the command bunker shown in Fig. 4.19, of the former Naval coastal artillery battery at Noirmont Point: one of four constructed in the Channel Islands (one each in Jersey and Guernsey, two in Alderney). Built between March 1943 and April 1944, the two-storey underground complex is reached by means of a long staircase. Outer walls are of concrete >2 m thick. Reproduced from the Channel Island Occupation Society (Jersey) leaflet ‘The Noirmont Command Bunker’ by kind permission of the Society

attraction and museum. It has more recently been refurbished as the ‘Jersey War Tunnels’.35 Klüpfel’s ‘short report’ of 4 February 1942 (listed on Table  4.1) summarizes geological conditions for four tunnel sites: Coin Varin, Bellozanne Valley, the St. Aubin Tunnel, and Grand Vaux Valley. ‘Coin Varin’ is seemingly Ho 1 (Munition Store I) constructed off La Route de l’Aleval, in the western side of the valley (see Figs. 4.21 and 4.24) in the Parish of St. Peter (rather than Ho 2: Ration Store I, in the east side, in the Parish of St. Lawrence) described and illustrated by Ginns (1993, 2012). Ginns records that this was under construction by 10 September 1941; that Klüpfel was still reporting on rock conditions there in August 1943 (as noted previously by Bishop and Launert 1979); and that although Ho 1 was never completed, construction was sufficiently advanced to permit the storage of ammunition from May 1944 onwards. Ginns has also described for the Bellozanne Valley, north of Jersey’s principal town, St. Helier, how Ho 3 (Munition Store II) was planned but never constructed 35

 https://www.jerseywartunnels.com/, last accessed 6 December 2019.

140

E. P. F. Rose

Fig. 4.21  Map of Jersey showing the positions of storage tunnels (numbered as for Table 4.5) and defensive tunnels (named) wholly or partly completed by 1945. Modified from Ginns (1993): reproduced from Rose et al. (2002)

on the west side of the valley, at the upper end, and Ho 9 (Electricity Generating Station I), at the east end. Ho 5 (Fuel Store I) was located inside the former Jersey Railways and Tramways Tunnel at the west side of St. Aubin’s Bay (Fig. 4.21), and converted into a munition store in 1944. Ho 4 (Munition Store III) at Grand Vaux was sufficiently complete to be used for its intended purpose from May 1944 onwards. However, Ho 12 (Fuel Store II) nearby at Grand Vaux, whilst under construction in mid-1943, had only been excavated for some 60 m when all work ceased. Klüpfel’s ‘short report’, which is not illustrated by any diagrams, notes that the tunnels at Coin Varin, Bellozanne Valley, and St. Aubin are all constructed within Präcambrisches Schiefergebirge, i.e. the rock commonly known as Jersey Shale Formation by British authors (see Sect. 2.3), whereas Grand Vaux Valley between ‘Pont Mill und Ponterrin Mill’ [sic] has bedrock within the Vulkanserie (i.e. the Jersey Volcanic Group), overlain by superficial deposits about 2–3 m thick. From Klüpfel’s notebooks, Bishop and Launert (1979, p. 35) note that two tunnels in particular merited his attention, Ho 1 (which he refers to as ‘Victoria Stollen’) and Ho 8 (‘Emphrie Stollen’), but that he visited other sites as well—none of them indicated by their official names or numbers. The Doktorhaus tunnel (Mont Matthieu), part of a ‘strongpoint’ (Fig. 4.21), was visited on 11 September 1941, Bellozanne Valley, and the St. Aubin’s tunnel (Ho 5) in early January 1942. A rock profile was recorded of 45 m of Ho 2 at this time, and reference made to a tunnel on Jubilee Hill. The Grand Vaux complex of tunnels is

4  Jersey and the German Army Table 4.5 Tunnels (Hohlgangsanlagen) scheduled for construction on Jersey by German forces during World War II, and their purpose. After Rose et al. (2002), modified from Ginns (1993, 2012)

141 Number 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25

Purpose Munition store I Ration store I Munition store II Munition store III Fuel store I Personnel shelter I Reserve artillery Artillery quarters Electricity works I Ration store II Personnel shelter II Fuel store II Munition store IV (Flak) Fuel store III (Flak) Stores Personnel shelter (?) Unknown Hospital (?) Harbour electricity generating station Mount Bingham connecting tunnel (?) Stores (?) Stores (?) Personnel shelter (?) Stores (?) Stores (?)

first mentioned in the spring of 1942: Klüpfel visited a tunnel (Ho 12) at Ponterrin Mill on 16 March which he recorded was being driven on the west side of the quarry. By June 1942 it is possible to identify Ho 10 as the tunnel being driven from Mal Assis Mill quarry, and on 13 August Klüpfel mentions ‘upper tunnel, Paul Mill’ which is probably Ho 4 or Ho 11. (‘Pont Mill’ in his ‘short report’, cited above, may be an error for ‘Paul Mill’.) Klüpfel made regular visits to tunnels during the rest of 1942, and especially in early 1943. Ho 13 in Beaumont Valley was under construction by 26 January 1943 and the Etaquerel tunnel by 3 March. During February and March Klüpfel made eight visits to Ho 8 alone, before drafting a ‘short report’ on 12 March. The report summarized the geological conditions in the Ho 8 tunnels and drew attention to an area of disturbed and faulted rock in the western part of the tunnel area. He indicated that the ground to the south of the tunnel would prove to be a crucial area and predicted that sound, unfaulted rock would be found there. Four projected chambers

142

E. P. F. Rose

Fig. 4.22  Sketch plan of Ho 8 (see Fig. 4.21), intended as an artillery headquarters but converted in 1944 for use as an underground hospital, and now operated as the Jersey War Tunnels. Bold stipple indicates the parts duly completed; fine stipple those planned but unfinished. Not to scale: dimensions for illustrative purposes only. For details of room use, and re-surveyed outline, see Ginns (2012, p. 114). From Rosenbaum and Rose (1992), courtesy of Blackwell/Wiley; © German Underground Hospital/Jersey War Tunnels, and reproduced by kind permission

were assessed as well sited because they would run near perpendicularly to the bedding of the rock. The month of August 1943 was taken up almost entirely by visits to Ho 1 and on 17 August Klüpfel drafted, for Fortress Engineer Staff 19 (as now the only Fortress Engineer Staff remaining in the Channel Islands: see Sect. 4.7), a ‘geological report on the storage tunnel system H 1, Jersey’. Three different rock groups were recognized, separated by faults and with displacements estimated as being 100  m or more. A major fault was met in the northern part of the tunnel system, the fault zone some 50 m wide, so Klüpfel recommended that the tunnel be diverted to the south and be driven through sound rock. He refers to a map showing these structures (cf. Fig. 4.24), but this does not seem to have survived the war. Although drafting of the report began on 17 August, a diary entry for 30 August refers to the need still to finish it. On 26 August Klüpfel recorded a visit to the Fort Regent tunnel (Ho 19: Fig. 4.21) and in later notes he refers to Verclut Point, the only indication that he might have visited the tunnels near St. Catherine’s Bay.

4  Jersey and the German Army

143

Fig. 4.23  Part of an unfinished tunnel in the German Underground Hospital, excavated through late Precambrian Jersey Shale Formation, with original timber roof supports. Dummy figures indicate working conditions. Reproduced from Rosenbaum and Rose (1992), courtesy of Blackwell/ Wiley; © German Underground Hospital/Jersey War Tunnels, and reproduced by kind permission

4.6  Quarrying for Raw Materials Construction of the fortifications on Jersey required very large quantities of building materials, especially concrete. It is therefore not surprising that Klüpfel completed a report on the local sources of naturally occurring materials as a priority task, some 4  months after his arrival on the island and as construction work intensified in response to the Führer’s instructions of 20 October and 15 December (see Sect. 3.2). Klüpfel was well qualified to do so. As noted above, he had been amongst the military geologists who pioneered compilation of thematic maps during World War I; he had wide experience of working in both ‘hard rock’ and ‘soft rock’ terrains; and he had long served as a consultant to a regional quarrying industry in Germany. Signed on Christmas Day, 25 December 1941, Klüpfel’s typewritten report contains six parts:

144

E. P. F. Rose

Fig. 4.24  German map showing ‘water supply for Coin Varin’ on Jersey; created by Fortress Engineer Staff 14 (Festungspionierstab 14), probably in 1942 (i.e. before the unit was withdrawn from the Channel Islands). The key (top right, top to bottom) refers to four symbols: (1) newly built roadworks; (2) tunnel system; (3) low-lying land; and (4) water conduits laid at a depth of 2.5 m. From Ginns (2012); reproduced by kind permission of the Channel Islands Occupation Society (Jersey)

4  Jersey and the German Army

145

1. Summary: 2 mm in diameter. 24  Wehrgeologenstellen. 25  Luftgaukommando Westfrankreich. 26  Militärbefehlshaber Frankreich. 27  Oberkommando der Wehrmacht. 22 23

5  Jersey and the German Air Force

169

Table 5.1  Membership of geologist appointment boards in Germany during World War II Berlin: Bonn: Halle: Munich: Stuttgart: Vienna:

(a) v. z. Mühlen (a) Rode (a) Weigelt (a) Beurlen (a) Wilser (a) Leuchs

(b) Cissarz (b) Päckelmann (b) Deubel (b) Schuster (b) Frank (b) Lotze

(c) Kraus (c) Schuh (c) Seidlitz (c) Kraus (c) K.G. Schmidt (c) Gallwitz

was assigned to Prague in Czechoslovakia. Schmidt’s assignment can now be amplified from 22 ‘expert opinions’28 preserved at the Bundeswehr Geoinformation Centre at Euskirchen in Germany. Written by Schmidt, these give his address as the military geology centre/team29 of the Air Command Region30 of western France from 1941 until 1943. Most of these documents relate to airstrip watering (to dampen dust) or water supply in general for German airfields (or other sites operated by the Luftwaffe) in the Normandy area (Rose and Willig 2013).

5.3  K.G. Schmidt on Jersey Little more than a week after he had visited Klüpfel on Jersey (on 3 May 1942: Table 4.1), Schmidt completed a geological report on reconnaissance for tunnel systems on the Channel Islands. It was a report only four pages in length, containing eight figures.31 The author32 of this report is given as K.G.  Schmidt, shown with his civilian academic status as a ‘professor’ who as a university graduate had achieved a diploma in mining engineering33 as well as a doctorate. His then wartime rank as a ‘Government Construction Administrator/Officer’34 was approximately equivalent to that of the head of a planning department or building control office in a British administration, and to the rank of a major in the Army (as explained in Sect. 3.6). The report is written as from the administrative centre35 of the Luftwaffe’s Air

 Gutachten.  Wehrgeologenstelle. 30  Luftgaukommando. 31  Geologischer Bericht Nr. 146. Grundsätze zur Geologischen Vorerkundung von Hohlgangsystemen auf dem Kanalinseln. Berichterstatter: Reg. Baurat Prof. Dr. K.G.  Schmidt, Dipl. Berging. Luftgaukommando Westfrankreich—Verwaltung—Az.: 63 c 26 A 87—Verw. III/7—Br. B.  Nr. 5122/42 geh. O.U., den 11.5.1942. [Bundesarchiv-Militärarchiv file RH32v.3041]. 32  Berichterstatter. 33  Dipl. Berging. 34  Regierungs-baurat. 35  Verwaltung. 28 29

170

E. P. F. Rose

Command Region36 for western France,37 as its geological report38 number 146. It was, however, prepared on 11 May 1942 on deployment39 rather than at the centre itself (known to have been based in Paris at this time). The document is stamped with the classification ‘secret’.40 The distribution list given at the end of the report indicates that eight copies were made, and sent to: 1 . Bautechn. Büro des RLM, Bln, Abt. 9 II F über Verw. III; 2. OKH In Fest Geol; 3. Geol. b. Stab d. Insp. D.  Landesbefestig. b. Ob. Befehlsh. West, Paris (Prof. Röhrer)—3 copies,41 4. Luftwaffenfeldbauamt Kanalinseln, 5. Verw. III/3, and 6. Verw. III/7 (Entwurf [top copy]). Thus three copies went to different units within the Luftwaffe’s department ‘III’; one to the geological staff of the Inspector of Fortresses at the High Command of the Army42 in Berlin; three to the military geologist (Professor Röhrer: see Sect. 6.3) on the staff of the Inspector of Land Fortification at the Commander-in-Chief West, based in Paris (cf. Sect. 3.4); and one to the Luftwaffe’s Field Works Office on the Channel Islands.43 The preface to the report states that Schmidt visited sites for tunnels on Guernsey between 19 and 21 April, accompanied by Reg. Baurat Theis, Reg. Bauass. a. K. Dr. Schneider, and Uffz. Dipl. Berging. Schulte. (For the Luftwaffe geologists Schneider and Shulte, see Sects 7.3 and 5.6, respectively.) On the afternoon of 23 April, Schmidt visited Army sites on Alderney, led by the Army geologist TKVR44 Dr. Hoenes (for whom see Sect. 6.4) but again accompanied by Dr. Schneider. Additionally, Schmidt visited two Jersey tunnel sites on 3 May, and an Army tunnel system there whose visit was led by an engineer, Ing. Meckel. As shown in Table 4.1, it is known that he visited the Army geologist Walther Klüpfel whilst on Jersey on 3 May. It is therefore evident that Schmidt had visited all three of the largest Channel Islands and consulted both Army and Air Force geologists already operational there before writing his report. After its preface, the report is divided into ten sections: principles; roofing and rock strength; weak zones; one direction of weakness (as at Trois Vaux Bay,

 Luftgaukommando.  Westfrankreich. 38  Geologischer Bericht. 39  O.U. = Ortsunterkunft: location undefined for military postal service. 40  Geheim! 41  Including that now preserved in the Bundesarchiv-Militärarchiv. 42  OKH In Fest. 43  Luftwaffenfeldbauamt Kanalinsen. 44  Technischer Kriegsverwaltungsrat. 36 37

5  Jersey and the German Air Force

171

Alderney); two directions of weakness (as near Fort Essex, Alderney); irregular patterns of weakness; munitions depots; principles of planning; and preliminary geological investigations (including preparation of ‘rose’ diagrams and stereographic plots for the joint and/or cleavage directions of bedrock, as illustrated below). The ‘principles’ section concludes with two key statements: that it is the task of a geologist to determine the direction for the major axis of a tunnel system; and that although the choice of site for a tunnel system should be determined by the tactical requirements, it should be guided by geologists. The report is illustrated by eight figures, giving examples of different situations in which tunnel alignment is related to the directions and type of rock weakness: 1. A rectangular tunnel system in rock with widely spaced zones of weakness, the zones at right angles to one another (Fig. 5.2). 2. A similarly rectangular tunnel system with widely spaced zones of weakness, but with main weakness zones oblique to one another, and tunnels appropriately off-set. 3. An obliquely orientated tunnel system aligned with oblique widely spaced weakness zones. 4. A ‘runic’ type of tunnel system, in rock with closely spaced but parallel zones of weakness (Fig. 5.3). 5. A rectangular tunnel system, all tunnels with the same profile, in rock with closely spaced zones of weakness (in two principal directions) with no region free of weakness; tunnels oblique to the directions of weakness. 6. A so-called fleur-de-lis pattern of tunnels (similar to Fig. 5.3 in that chambers radiate at an incline from a central axis, but with more—6 to 8—chambers each side, the chambers lacking interconnections, and at least two such chamber complexes lead from the side of a main access tunnel providing a through route); in badly weakened rock, with two main directions of weakness at right angles to one another, excavated mostly in unweakened stone. 7. A similar ‘fleur-de-lis’ in badly weakened rock, with two main weakness directions at right angles to one another; chambers radiate at right angles from the main axis, so that all excavation is at an angle to the weakness directions. 8. Three diagrams showing entry to the main access tunnel for a ‘fleur-de-lis’ system in rock with one main weakness direction, and two cleavage directions. Tunnel systems were to be excavated quite extensively on Guernsey and Jersey, and to lesser extent Alderney, as described by Ginns (1993, 2012). According to its preface, this report was written to provide a summary of principles to guide Luftwaffe geologists, and was written independently of guidelines issued earlier by the High Command of the Army.45 The principles were evidently soon put into practice: to guide site investigation and detailed planning for two locations on Jersey.

45

 Oberkommando der Heeres.

Fig. 5.2  Principles for planning the construction of underground accommodation. Sketch by Prof. Dr. K.G. Schmidt. Betonausbau = concrete finish; Fels = rock; Störungszone = zone of weakness. Example of a rectangular tunnel system in rock with widely spaced zones of weakness. Main weakness zones perpendicular to one another. Tunnels [should be] parallel to the weakness directions, [and be] constructed as much as possible in [those areas of] rock free of weakness. Appendix [Anlage] 1 (of 8) from Geological Report (here called ‘Expert Opinion’) 146; reproduced by permission of the Bundesarchiv-Militärachiv from file RH32/3041

5  Jersey and the German Air Force

173

Fig. 5.3  Example of a ‘runic’ type of tunnel system, in rock with closely spaced parallel zones of weakness or planar features. Tunnels [should be] perpendicular or at an angle to the weaknesses. Appendix 4 from Geological Report (here called ‘Expert Opinion’ 146) (cf. Fig. 5.2); reproduced by permission of the Bundesarchiv-Militärachiv from file RH32/3041

174

E. P. F. Rose

5.4  The Geologist Kurt Diebel The most tedious part of the site investigation was assigned to Schmidt’s assistant, listed as ‘Dr. Dübel of Berlin’ by Walther Klüpfel (cf. Table 4.1) but identified by Rose and Willig (2013) as Dr. Kurt Diebel. Schmidt himself credits ‘Hilfsgeologe Uffz. Dr. Diebel’, as his assistant on Jersey (in a report described in Sect. 5.5); the status, rank, title and name correspond precisely with credit on a subsequent report by Schmidt prepared in September 1942 for a site in France; and a Diebel but not a Dübel is known to have gained a geological doctorate and employment in Berlin. ‘Diebel’ thus appears to be the correct name spelling for the ‘Dr. Dübel’ recorded in Klüpfel’s diary of events (Table 4.1) as being present with Schmidt on Jersey on 3 May 1942. According to an obituary (Pietrzeniuk 1982), Kurt Diebel (Fig. 5.4) was born on 17 March 1915 in the Prussian province of Posen, but educated at school and university in Berlin. From the winter of 1933 he studied first zoology and chemistry, and then geology and palaeontology, obtaining the degree of ‘Dr. phil. Nat.’46 in 1939 for a thesis on an oil shale of Jurassic age in the Bielefeld region of northern Germany. Subsequently published (Diebel 1941) by the German national ‘Geological Survey’,47 the thesis is entirely stratigraphical and palaeontological in its focus: ‘pure’ rather than ‘applied’ geology. Pietrzeniuk (1982) records that on completion of his university studies, Diebel was called up for military service—but makes no mention of what that service entailed. Häusler (1995b, p. 15, 63) records that ‘Dübel’ served in October 1942 as a geologist with the Luftwaffe, but nothing more about him before that date. Since Diebel’s thesis was published by the national ‘Geological Survey’, and in 1941, he evidently had some association with that organization following his graduation. However, since he had achieved the rank of corporal by May 1942, he had equally clearly served long enough in the Army by that time to have achieved promotion through the more junior ranks.

5.5  Kurt Diebel on Jersey Schmidt as the senior member of the two-man ‘team’ was to complete a report concerning construction of underground facilities on the island of Jersey. This report contained only two pages of text, but was illustrated by 24 figures.48

 See footnote 12.  Reichstelle für Bodenforschung, re-named after 1941 the Reichsanstalt für Bodenforschung. 48  Geologisches Gutachten Nr. 161. Geologisches Gutachten betrifft Hohlraumbauten auf der Insel Jersey. Berichterstatter: Reg. Baurat Prof. Dr. K.G. Schmidt, Dipl. Berging. Luftgaukommando Westfrankreich—Verwaltung Az. A 88—Verw. III/7. Br. B. Nr. 1288/42 g. Kdos. O.U., den 30 Mai 1942. [Bundeswehr Geoinformation Centre, Euskirchen]. 46 47

5  Jersey and the German Air Force

175

Fig. 5.4  Kurt Diebel (1915–1981), from a group photograph taken in the summer of 1955 during a micropalaeontological field excursion in NW Germany; courtesy of the late Professor Eugen Seibold, Freiburg-im-Breisgau

The report was issued from the same address as the report by Schmidt for 11 May, also ‘on deployment’, but as ‘expert opinion’49 number 161. ‘Expert opinions’ were shorter documents than reports50 strictly so-called, but provided more technical detail than ‘short reports’51 (as described in Sect. 3.5). This ‘opinion’ was one that dealt with ‘command matter’52 and is stamped ‘secret command matter’.53 It was prepared with reference to a ‘planning report for Jersey’54 issued by the Luftwaffe Field Works Office for the Channel Islands.55 Although it does not refer to the ‘tunnel’ report of 11 May, it obviously applies some of the principles outlined in that document. The distribution list (for only five copies) that ends the ‘opinion’ is very similar to that for the 11 May report: 1. RLM, Bautech. Büro 9 I (Geol);

 Gutachten.  Berichte. 51  Kurzberichte. 52  Kdos = Kommandosache. 53  Geheime Kommandosachen! 54  Bericht zur Planung Jersey Az.63 c 26 Br. B. Nr. 114/42 g.Kdos. 55  Luftwaffenfeldbauamt Kanalinseln. 49 50

176

E. P. F. Rose

2 . OKH In Fest Geol;56 3. Lw. Feldbauamt Kanalinseln; 4. LGK Verw. III/3; and 5. Verw. III/7 (Entw.) [top copy]. This time the first copy on the list goes explicitly to the geologist at the Luftwaffe’s ‘Building Technology Office’, rather than just to the Office in general (evidence therefore that geologists had a chain of command within the Luftwaffe similar to the chain that operated within the German Army). No copies were sent to the staff of the Commander-in-Chief West. However, the top copy was again sent to the Luftwaffe’s department ‘III’, and single copies to the geological staff of the Inspector of Fortresses at the High Command of the Army (in Berlin) and to the Luftwaffe’s Field Works Office on the Channel Islands. According to the preamble opening the report, it was based on site investigation by the author (Schmidt) and measurement of rock fracture directions by Hilfsgeologe Uffz. (assistant geologist Corporal) Dr. Diebel. The two-page text is closely typed, and accompanied by two maps (appendices 1 and 22), 11 rose diagrams and 11 stereographic plots. It focuses on two projects: (1) a storage depot for aircraft munitions and fuel and (2) a depot for anti-aircraft gun ammunition—respectively the ‘St. Aubin Tunnel’ and ‘Vallée des Vaux’ (north of St. Helier) areas that Klüpfel and Diebel visited together on 4 May.57

5.5.1  The St. Aubin Tunnel The region surveyed for the munitions and fuel depot, inland from St. Aubin’s Bay on the south of the island (Fig. 5.5) en route to the airfield, is shown on an accompanying map (Fig. 5.6). The potential depot is drawn as a massive complex: two ‘fleur-de-lis’ structures, one of three axis routes and the other of two, aligned parallel to the strike of bedding, with nine northern and eight southern chambers aligned at right angles to each axis route. A southern entrance to the complex as a whole is shown leading off the Mont au Roux road, which leads down to a slipway in St. Aubin’s Bay. A northern entrance connects with the road to the airfield.58 A spur from the Jersey railway is shown entering from the NE via the region known to the Germans59 as ‘Beaumont Valley’, at the junction of the two fleur-de-lis structures. A major access route60 for the fuel depot is drawn to the north of area 3, extending NE from an entrance near the school, beside Mont les Vaux.

 The copy preserved in the Bundeswehr Geoinformation Centre.  Table 4.1. 58  Zum Flugplatz. 59  Not known to local people by this name. 60  Hauptachse. 56 57

5  Jersey and the German Air Force

177

Fig. 5.5  Map of Jersey, showing positions of the German underground facilities constructed or under construction as at May 1945. Defensive tunnels are named (Bty. = artillery battery, Stp. = heavily fortified infantry ‘strongpoint’; Wn. = infantry ‘resistance nest’, less fortified than a ‘strongpoint’), storage tunnels are numbered. From Rose and Willig (2013); after Ginns (1993), but with the position of the airfield added, by kind permission of the Channel Islands Occupation Society (Jersey)

Six nearby areas of rock outcrop (e.g. Fig. 5.7) in which joints and faults indicating directions of weakness in the bedrock could be measured are numbered 2–7 on the map, from south to north. These measurements generated a rose diagram61 showing the orientation of all faults in the region (Fig. 5.8) plus diagrams for the three separate areas with sufficient (15, 11 and 10) measurements to be significant, a rose diagram showing orientations of joints in the rock within each of the six areas (e.g. Fig. 5.9), and a stereographic plot62 (e.g. Fig. 5.10) of pole positions for joints or faults to accompany each of the rose diagrams— techniques still standard procedure in structural geology. Rose diagrams are plotted at 5° intervals. The depot is seemingly that designated Ho 14 (Fuel Store III—Flak: see Table 4.5), sited half way up Le Mont au Roux (La Haule Hill) at St. Brelade and intended for Lufwaffe use, a facility which Ginns (1993, 2012) states did not progress beyond the planning stage—although a start was made on the associated Ho 13 (Munition Store IV—Flak), following a survey led on 8 July 1942 by a young second lieuten-

 A circular histogram plot displaying directional data and the frequency of each data class.  Stereographic projection is a geometrical mapping technique used in many scientific disciplines that projects a sphere on to a plane; commonly used in structural geology to plot the orientation of linear and planar features. 61 62

178

E. P. F. Rose

Fig. 5.6  Map, original at scale of 1:5000, from ‘Expert Opinion’ 161, dated 30 May 1942, now held by the Bundeswehr Geoinformation Centre at Euskirchen, showing the position of planned underground munition and fuel depots inland from St. Aubin’s Bay (cf. Fig. 5.5) en route to the airfield on Jersey. Areas in which joints and other planar discontinuities in the bedrock were measured are shown with grey stipple and numbered 2–7. Points A and B are end points of a potential route alignment discussed in the report. From Rose and Willig (2013); reproduced by permission of the Bundeswehr Geoinformation Centre

ant63 of the Luftwaffe. Schmidt’s report concluded that the site ‘behind the school’ designated in an earlier ‘planning report’ for an access tunnel to the fuel depot was ‘good’ in terms of its geology. However, a proposal to align the main axis and access route for the munitions depot between points A and B shown on Fig. 5.6 could not be supported, because of poor rock and unsuitable slopes, and a new route was proposed. The rock itself is not described in the report, but all the outcrops lie within 63

 Leutnant.

5  Jersey and the German Air Force

179

Fig. 5.7  View of rock outcrop in the Beaumont Hill area shown in Fig. 5.6, illustrating rock typical of the Jersey Shale Formation (described in Chap. 2) through which tunnelling nearby was planned, and the jointing whose orientation was measured to create a ‘rose diagram’ (e.g. Fig. 5.9) for each of the areas shown in Fig. 5.6. Photo: J.T. Renouf

the region long mapped (and identified by Klüpfel in a ‘short report’ of 4 February 1942: see Table 4.1) as ‘Jersey Shales’: typically thin- to medium-bedded, fine- to medium-grained sandstones with subordinate siltstones and mudstones that have undergone low-grade regional metamorphism and been folded/faulted by tectonic processes (see Sect. 2.3).

5.5.2  The Vallée des Vaux The second site was an ammunition depot planned for construction north of St. Helier (Fig. 5.11). This is shown again as a ‘fleur-de-lis’, albeit more schematically, with three axis routes north of a main entrance route leading eastwards from the Vallée des Vaux. Joint measurements made in a nearby quarry, where bedrock was accessible on the surface, generated a rose diagram and accompanying stereographic plot in the same way as for those made in the St. Aubin’s Bay area.

180

E. P. F. Rose

Fig. 5.8  Rose diagram, from ‘Expert Opinion’ 161, showing the orientation of 41 faults measured in total within the six areas shown in grey and numbered 2 to 7 on Fig. 5.6, together with the alignment of the main axes [Richtung Hauptachse] of the munition and fuel storage depots (Munilager and Treibstofflager, respectively). Each concentric circle represents one unit of measurement within a 5° category. From Rose and Willig (2013); reproduced courtesy of the Bundeswehr Geoinformation Centre

5  Jersey and the German Air Force

181

Fig. 5.9  Rose diagram, from ‘Expert Opinion’ 161, showing joint orientations based on 194 measurements within area 2 (St. Aubin railway tunnel) shown on Fig. 5.6, together with direction of bedding [Richtung Schieferung] in the rock and alignment of the main axis of the fuel depot. Each concentric circle represents one unit of measurement within a 5° category. From Rose and Willig (2013); reproduced courtesy of the Bundeswehr Geoinformation Centre

182

E. P. F. Rose

Fig. 5.10  Stereographic plot, from ‘Expert Opinion’ 161, showing pole positions for the 194 joint inclinations and the regional dip [Schieferung] of bedrock illustrated on Fig. 5.9. From Rose and Willig (2013); reproduced courtesy of the Bundeswehr Geoinformation Centre

5  Jersey and the German Air Force

183

Fig. 5.11  Map, from ‘Expert Opinion’ 161, original at scale of 1:25,000, showing the position of a munitions depot planned for construction north of St. Helier on Jersey. From Rose and Willig (2013); reproduced courtesy of the Bundeswehr Geoinformation Centre

184

E. P. F. Rose

The report describes the quarry as excavated into a strong porphyry and amygdaloidal rock (thus part of the Jersey Volcanic Group of modern authors; see Sect. 2.3), with cleavage fractures in the rock trending NW–SE in strike direction. It confirms as sound a ‘planning report’ proposal to site the western entrance to the depot in the quarry, but recommends that the site for the eastern entrance be re-positioned slightly northwards, where the hill slope was steeper. A broad access route perpendicular to the cleavage strike was proposed, with narrower outlets for one-way traffic created oblique to the strike direction. The ‘stem’ of the fleur-de-lis could be aligned with the cleavage strike direction, so that the chambers might be arranged perpendicular to it. This facility is another one that seems not to have progressed beyond the planning stage. It is not amongst those known to exist in 1945 (Fig. 5.5).

5.6  The Geologist Franz Schulte The ‘Dr. Schulte’ of the Luftwaffe who visited Walther Klüpfel on Jersey (Table 4.1) has been identified as the mining engineer Corporal Schulte whom Professor Schmidt met on Guernsey in April 1942 (Rose and Willig 2014). As he seems not to have pursued an academic career, biographical details have proved elusive, and no photograph of him has yet been found. A search of published literature has revealed only one article for which Schulte was the author, a ‘Contribution to geological mining investigation with the aid of the geophysical measuring instrument “geoscope”’ (Schulte 1940). Based on a study carried out in 1938, on a coal mine in the Lower Rhine area, owned by the firm ‘Hessische Feinmechanik’ in Giessen, this was published on 19 October 1940. An American translation of the author’s abstract was published the following year (WA 1941), under ‘electrical methods’. Using the ‘geoscope’, Schulte was able to determine the depth to the base of soft rocks overlying coal-bearing deposits of Carboniferous age. He therefore inferred that the ‘geoscope’ had great potential for use in mining investigations. Since his paper is illustrated by 19 figures which illustrate ground conditions to a depth of about 500 m, and (beneath a thin superficial cover) a faulted sequence of Tertiary sedimentary rocks underlain by Triassic (Bundsandstein and Zechstein) rocks above the Carboniferous, his education is presumed to have included a training in geology as well as geophysics. Authorship of the German paper is printed as ‘Dipl.-Ing. Markscheider Fr. Schulte, Homberg (Niedrerrhein)’. Schulte had thus qualified as an engineer (Dipl.Ing.) by this time, and gained employment in the coal-mining industry as a mine surveyor (Markscheider) based on the town of Homberg (now within Duisburg, a large city in the western part of the Ruhr region of northern Germany, in the state of North Rhine - Westphalia). He must have been over 20 years of age in 1940 to have achieved his professional qualification and employment—perhaps older, since there

5  Jersey and the German Air Force

185

is no acknowledgement in his article to the supervision that would have been customary had this been a student project. In Germany, ‘Coal was of enormous importance to the economy, especially during war’ (Gillingham 1982, p. 637), and the coal and steel industries of the Ruhr were of such importance that they were intensively bombed by Allied air forces (Dear and Foot 1995). It seems unlikely, therefore, that a ‘mine surveyor’ of mature age and experience would have been released from duties of importance to the war effort for service in the junior ranks of the armed forces, and probable that Schulte was then a young man, in his 20s rather than 40s in years of age. It is assumed that 1940 marked the end of Schulte’s professional education, and that like other young men, he volunteered or was conscripted for military service soon after graduation. He would thus have had appropriate time to complete his basic military training and sufficient service to achieve the rank of corporal by April 1942, when he appears in the Channel Island documentary records. Less than another year’s experience evidently qualified him for promotion: he was an official equivalent in rank to a captain64 (as explained in Sect. 3.6) by the following April. To be commissioned after qualifying service in the ranks was normal practice in the German armed forces at that time: e.g. geologist Dieter Hoenes (described in Sect. 6.4) was promoted in this way, after service on the Channel Islands as a lance corporal. Häusler’s (1995b, p. 44, 82) first record for ‘Franz Schulte’ in his authoritative account of work by German military geologists as a whole during World War II, for April 1942, is probably in error, for the reasons given by Rose and Willig (2014). However, Häusler does note that by October 1942 Franz Schulte had been assigned to a geological centre/team65 of the Luftwaffe, and that from October 1943 he served within a Field Works Office66 that was part of the Senior Construction Management67 responsible to the Luftwaffe General of the Channel Islands, based in Guernsey. The record provided here and by Rose and Willig (2014) is consistent with, and amplifies, these brief service details.

5.7  Franz Schulte on Jersey Schulte wrote at least one report for Jersey that is known to have survived. It has a text of two pages, and five annexes.68  Regierungsbauassessor.  Geologenstelle. 66  Feldbauamt 1/3. 67  Oberbauleitung. 68  Gutachten über die Baugrundverhältnisse beim Dammdurchlass im Beaumont-Tal auf der Insel Jersey. Der Geologe beim Lw. Feldbauamt 14 im Luftgaukommando Westfrankreich. Anfordernde Dienststelle: Lw.-Feldbauamt 14. Sachbearbeiter: Reg. Bauassessor a.K.  Franz Schulte, Dipl.Bergingenieur. O.U., 9 April 1943. [Bundeswehr Geoinformation Centre, Euskirchen]. 64 65

Fig. 5.12  Site map for part of the Beaumont Valley on Jersey (viewed to SSW), about 150 km south-east from Ho 13 (see Fig. 5.5), contoured at 2 m intervals, and showing line of geological cross-section marked by vertical dashed line A-B, crossing sites of trial pits numbered I to IV, and outline of proposed embankment cutting (as central dashed outline). Part of an annex [Anlage] to a geological ‘expert opinion’ [Gehört zu geol. Gutachten] by F. Schulte, dated 9 April 1943. From Rose and Willig (2014); reproduced courtesy of the Bundeswehr Geoinformation Centre

5  Jersey and the German Air Force

187

Fig. 5.13  The culvert which was intended to cover the stream flowing down Beaumont Valley, as shown on Fig. 5.12. Had tunnelling not ceased in the autumn of 1943, the entire valley floor would have vanished beneath some 4 m of rubble. Photo: J.T. Renouf; cf. Ginns (2012, Fig. 12)

This ‘opinion’69 (a shorter document than one that would be titled a ‘report’) for a valley site in mid-Jersey (Figs. 5.5, 5.12, 5.13, and 5.14) is shown as a geological opinion on its annexes (e.g. Figs. 5.15 and 5.16) although the word ‘geological’ is missing from the title heading the typescript. It was written by the geologist70 at Air Force Field Works Office number 1471 in the Air Force Regional Command area72 of western France.73 The office requesting the opinion74 was Field Works Office 14 itself. The expert author75 was Franz Schulte, writing ‘on deployment’.76 According to the covering letter (dated 30 June 1943, signed by Schulte on behalf of the head77 of Field Works Office 14), five copies were made, and sent to:  Gutachten.  Der Geologe. 71  Luftwaffen Feldbauamt 14. 72  Luftgaukommando. 73  Westfrankreich. 74  Anfordernde Dienststelle. 75  Sachbearbeiter. 76  O.U. = Ortsunterkunft. 77  Leiter. 69 70

188

E. P. F. Rose

Fig. 5.14  Entrance to the culvert shown on Fig. 5.13. Photo: J.T. Renouf

1. Amt f. zentrale Bauaufgaben der Luftw. Bautechn. Abtlg. Arbeitsgeb. Wehrgeologie Berlin SW 68; 2. O.K.H. In Fest. Geol.; 3. LGK.-WFr. Verw. B III; 4. LGK.-WFr. Verw. B III 12/b; and 5. Lw. Feldbauamt 14 (Entwurf) [top copy]. Thus the first two copies went to geological staff in Berlin (of the Luftwaffe and the High Command of the Army, respectively); the next two to units within the

5  Jersey and the German Air Force

189

Fig. 5.15  Geological cross-section A-B across the Beaumont Valley, coloured (as explained in the text) to indicate the occurrence of five different soil/rock conditions, and with the position of the proposed embankment cutting outlined to the right; on original, horizontal scale 1:500, vertical scale horizontal ×5. Part of an annex [Anlage] belonging to a geological ‘expert opinion’ [Gehört zu geol. Gutachten] by F. Schulte, dated 9 April 1943. From Rose and Willig (2014); reproduced by permission of the Bundeswehr Geoinformation Centre

administrative headquarters of the Air Command Region of western France78 (known to be in Paris at this time); and the fifth, the top copy, went to the Luftwaffe’s Field Works Office 14 at its Channel Islands’ base. It is the copy sent to the geologist at the Inspectorate of Fortresses, at the High Command of the Army (in Berlin), which has been preserved. The two-page text opens with a brief preamble that sets the scene: the site is south of tunnel systems79 numbered 13 and 15 (Fig. 5.5), where an embankment is scheduled for construction through the ‘Beaumont Valley’, with a passage way through the embankment for a road and a railway (cf. Fig. 5.6). Next comes a section on geological conditions, describing the site as within an area of Precambrian slates with intercalated beds of greywacke (i.e. impure sandstone): the Jersey Shale Formation of British authors (as described in Sect. 2.3). A section on the surface weathered zone concludes the text, noting that the constituents of this zone comprise slope debris on slightly weathered greywacke with a loam covering. From the top down:

78 79

 Luftgaukommando Westfrankreich.  Hohlgangsanlagen.

190

E. P. F. Rose

Fig. 5.16  Log of trial pit [Schurf] number 1, dated 26 February 1943, for the embankment passage near Ho 13 and 15, dug about 150 m SE of ‘tunnel [Stollen] 1 (Ho 13) as shown on Fig. 5.12. Columns left to right give, for each of eight units distinguished within the pit, dug to 2.1 m (1) depth in metres to top and bottom; (2) description of soil/rock material; (3) geological interpretation; (4) groundwater level (‘moist’ at top, ‘water-bearing’ beneath and at one lower horizon); (5) bearing strength in kg/cm2 (unsurprisingly increasing with depth, from 0.5 to 4.0); and (6) permeability (low at the very top, bad at the very bottom, but good for the two water-bearing units). Source and acknowledgement as for Figs. 5.12 and 5.15

5  Jersey and the German Air Force

191

1. humic alluvium; 2. greywacke rubble (first water level); 3. weathered loam; 4. greywacke and slate rubble (second water level); and 5. bedrock. These five divisions are illustrated on a hand-coloured geological profile, adjacent to a large-scale map of the site, annexed to the text. Both map (Fig. 5.12) and profile (Fig. 5.15) indicate the position of the four trial pits that provided the evidence of subsurface conditions (e.g. Fig. 5.16). The logs of these pits (numbers 1–3 dug to some 2 m depth, number 4 to 4.6 m), also attached as annexes, describe the successive strata, with columns summarizing their geological identification, groundwater content, bearing strength and permeability. This study is therefore a sequel to that by Schmidt and Diebel in May 1942, as described in Sect. 5.5, for a Luftwaffe tunnel system to be partly accessed by rail via the ‘Beaumont Valley’. It adds little to knowledge of Jersey’s geology, but is of historical significance in that it illustrates the procedure for logging trial pits then presumed to be standard procedure for site investigation by the Luftwaffe.

5.8  Schmidt, Diebel and Schulte after Jersey Professor Schmidt completed a report for Guernsey (see Sect. 7.4), but then returned to work in France, together with his assistant Kurt Diebel. Franz Schulte continued to work in the Channel Islands, based on Guernsey, until at least October 1943 (see Sect. 7.6), but was then transferred to the Balkan peninsula, in the Mediterranean area of operations.

5.8.1  K.G. Schmidt Schmidt was evidently focused on work in Normandy by the autumn of 1941. Documents preserved at the Bundeswehr Geoinformation Centre at Euskirchen include a two-page ‘expert opinion’ signed on 12 September 1941 recommending that a well be drilled at le Bourg Chateau, some 7 km west of Caen on the road to Bayeux (Rose and Willig 2013);80 a three-page geological ‘opinion’ on water supply for the airfield at Carpiquet,81 signed on 19 September 1941; and a two-page ‘opinion’ of nearly a year later, signed on 29 September 1942, predicting well-drilling conditions to supply water to a Luftwaffe camp of 300 men near Douvres-la-­  Geologisches Gutachten Nr. 49. Gutachten über die Wasserversorgung der Unterkunft Bourg, bei Caen. 81  Geologisches Gutachten Nr. 60. Gutachten über die Entwässerung von Flugplatz Carpiquet. 80

192

E. P. F. Rose

Délivrande, some 13 km north of Caen.82 All of these were written as from the military geology centre/team83 for the Air [Force] Construction Command84 for western France. Since these ‘opinions’ are numbered 49, 60 and 182, respectively, it is evident that this particular military geology centre/team had generated at least 182 ‘expert opinions’ by the end of September 1942—a considerable amount of work; that Schmidt himself was responsible for many of these; and that although the documents were compiled for a wide range of local units, the copy of the report which has been preserved at the Bundeswehr Geoinformation Centre is that sent to the High Command of the Army (in Berlin). On these documents, Schmidt’s status is given more fully, as Regierungs Baurat a. K. (a status explained in Sect. 3.6). Schmidt continued to serve with the Luftwaffe in France until the country’s Allied liberation in the summer of 1944: Häusler (1995b) records him as the geologist for the central region85 at Field Works Office number 686 of the Air Command Region for western France87 in July 1944, tasked with work in the French Jura. However, as the Allies advanced and the Luftwaffe was progressively degraded by Allied attack, he was transferred in the winter of 1944 to a unit88 of the paramilitary construction agency Organisation Todt (cf. Sect. 3.7) based at Weimar in Germany. The February 1960 letter in his personal file notes that on 20 January 1945, Schmidt was promoted to serve as the head89 of the Hochschule at Karlsruhe.90 However, following the surrender of Nazi Germany on 8 May, he was dismissed from office on the orders of the new Allied military government. Four years of subsequent unemployment were attributed to injury and ill health. However, from September 1949 he gained employment in Bonn, in the Mineralogy-Chemistry Department of a Dust Research Institute operated by the federation of trade unions representative of the German ceramic and glass industries. There he developed research on industrial illnesses such as silicosis (a respiratory disease caused by breathing in silica dust), in time developing associations with academic institutions in Bonn, Aachen, Essen and Düsseldorf, and an appropriate publication list, e.g. Schmidt and Lüchtrath (1958). His Festsetzung (see footnote 11) records that he retired at the end of June 1967, and University of Karlsruhe archives note that he died on 30 October 1976, aged 74.

 Geologisches Gutachten Nr. 182. Gutachten über die Wasserversorgung der Ln—Anlage Douvres. 83  Wehrgeologenstelle. 84  Luftbaukommando. 85  Bezirksgeologe Mitte. 86  Feldbauamt 6. 87  Luftgaukommando Westfrankreich. 88  Einsatzgruppe IV (Kyffhäuser). 89  Rektor. 90  Noted also by Hoepke (2007, p. 125, 165). 82

5  Jersey and the German Air Force

193

5.8.2  Kurt Diebel Häusler (1995b, p. 14, 62) records that Corporal ‘Diebel’ was later promoted to a Beamter-grade appointment (with status equivalent to a captain: see Sect. 3.6), and followed a pattern of employment similar to that of Schmidt. He served from September 1943 as the geologist for the southern region91 at Field Works Office 1392 of the Air Command Region for western France, before transfer in the winter of 1944 to a unit93 of the Organisation Todt, based at Essen in Germany. He thus appears to have served as a military geologist until the end of the war. After the war, Diebel became an ‘assistant’ at the Geological Institute of the University of Jena in eastern Germany, moving in 1947 to become an ‘assistant’ at the Geological-Palaeontological Institute and Museum of the Humbolt University in the eastern sector of Berlin. Gross and Schultze (2004, Table 1) confirm that he had obtained his doctorate in Berlin in 1939, and record that he was appointed as an ‘assistant’ to the staff the University’s Natural History Museum94 in 1947, being promoted to the status of a curator in 1951 and remaining as such until retirement in 1980. Established in 1810, the museum had developed by World War II into the largest museum of natural history in Germany. Diebel became curator of microfossils and leader of the Museum’s working group on micropalaeontology. Consequently, his postwar publications (e.g. Diebel 1956, 1957, 1960) were focused on aspects of this particular discipline: he wrote at least 15 papers on Quaternary non-marine ostracods (Rose and Willig 2013). He did not, so far as is known, publish any geological work derived from his brief wartime service on the Channel Islands. He died on 3 November 1981, aged only 66.

5.8.3  Franz Schulte According to Häusler (1995b), from March 1944 Schulte was re-assigned to the Army, serving in Wehrgeologenstelle 35 with Fortress Engineer Staff 19, for whom he compiled a report that month on tunnel construction near the harbour of Bar in the Antivari region of Montenegro (Häusler 1995b, p. 149). In August 1944, still with Wehrgeologenstelle 35 but now attached to Engineer Regiment Staff 679, Schulte completed a report on tunnel construction near Rogotin, a seaside resort in southern Dalmatia, present-day Croatia, further to the north along the Adriatic Sea coast of the Balkan peninsula. Since this period follows the end of significant construction work on the Channel Islands, and the reports deal with tunnelling and so

 Bezirksgeologe Süd.  Feldbauamt 13. 93  Einsatzgruppe III (Hansa). 94  Museum für Naturkunde. 91 92

194

E. P. F. Rose

seemingly continue an aspect of that work, it seems likely that they were indeed written by the Franz Schulte whose work is described in this book. In the winter of 1944, according to Häusler (1995b, p.  44), like his former Channel Island Luftwaffe geologist colleague Hans Schneider (whose work on Guernsey is described in Sect. 7.2), Schulte was assigned as a military geologist to a unit95 of the Organisation Todt—and based in Villach, a town in southern Austria now popular as a ski resort. Presumably like Schneider he surrendered to the western Allies rather than forces of the Soviet Union, and pursued a career in the west thereafter, possibly even returning to his pre-war career in Germany’s coal-mining industry until that was brought almost to an end by the 1970s. The ‘geoscope’ with which he had demonstrated expertise pre-war was still in use postwar (see, e.g. Anon. 1950).

5.9  Conclusion Forty (1999, p. 110), in giving a German perspective of the German occupation of the Channel Islands, notes that Jersey’s newly constructed airport at the time of its capture boasted not only four grass runways, but a modern terminal building, complete with control tower. It had state-of-the-art direction-finding equipment, plus floodlights to assist night landing. It was ‘The most completely equipped of any [airport] in the British Isles apart from Croydon’. During the aerial Battle of Britain that raged through the summer of 1940, the airport was used primarily by the twin-­ engined aircraft Messerschmidt Bf110C of a fighter-bomber wing96 and the Dornier Do217s of a long-distance reconnaissance group.97 Dorniers (with operational strength of nine aircraft) of second Squadron, Long Range Reconnaissance Group 123 were transferred from France to Jersey during the period 15 to 17 July 1940. The airport was much used and enlarged during that summer (Ginns 2001). The small airfield rapidly became congested, and the German forces extended it by adding further grassed areas that allowed for a 930 m NW-SE runway, in addition to the others. However, by the end of the year fighter defences of the Royal Air Force made daylight reconnaissance missions over England difficult, and German military objectives had re-focused eastwards. Accordingly, the squadron (some 680 men in total) left Jersey on 18 February 1941, leaving only a maintenance unit at the airport to service the occasional visiting plane. Forty (1999, p. 47) further notes that during the early months of the Occupation, Luftwaffe personnel on the Channel Islands outnumbered those of the Army and Navy, principally because the Jersey and Guernsey airfields were used for staging and refuelling by German aircraft taking part in the Battle of Britain. An Air Force

 Einsatzgruppe VIII (Alpen u. Italien).  Zerstöresgeschwader 76. 97  Fernaufklärungsgruppe 123. 95 96

5  Jersey and the German Air Force

195

Works Company98 was located and busy on each airfield, carrying out enlargement work. After the Battle, Jersey airfield’s main use was for communication flights between the island’s garrison and France. With such low usage, there was no need for further alteration of the airfield as such, and thus apparently no need for site investigation by geologists. It is evident, however, from the studies described in this chapter and the comprehensive accounts of German tunnelling on Jersey by Ginns (1993, 2012) that underground facilities for the Luftwaffe to store fuel and ammunition formed part of the overall fortification plan for Jersey and that some tunnelling for the Luftwaffe was in progress if not completed by the end of hostilities. A geologist was resident on Guernsey with the Luftwaffe Works Office there for at least 18 months (initially both Schneider and Schulte, later Schulte alone). It therefore seems likely that advice would have been contributed to Luftwaffe construction units wherever needed much as Klüpfel provided advice to Army units tunnelling on Jersey. The two geological ‘expert opinions’ described in this chapter probably represent only a small part of the geological service actually rendered. Moreover, in the Luftwaffe generally and on the Channel Islands in particular, more men served in anti-aircraft gun batteries than on airfields. According to Forty (1999, p. 111), when war began, almost a million men (nearly two-thirds of the total Luftwaffe manpower) were serving generally with the Flak (anti-aircraft) arm. The number of men and women increased to a total of about 1.25 million by the autumn of 1944. Whilst the anti-aircraft units on the Channel Islands formed only a very small percentage of this total, they were nonetheless a significant proportion of the garrison. A Flak regiment was stationed on Jersey, another on Guernsey, together forming a complete Flakbrigade. By September 1944 Jersey had at least 165 guns in position—thirty-six 88 mm calibre Flak 36 and Flak 37 guns, fifteen 37 mm Flak 41 guns, about 104 light 20 mm Flak 30, Flak 38 or Flak Oerlikon guns, and ten of the four-barrelled 20 mm Flak vierling weapons—excluding mobile weapons and those mounted on ships in the harbour (Ginns and Bryans 1978). It has been noted above that Schulte was providing advice on water supply for an anti-aircraft battery at Mont Gavey in October 1942, and it is more than likely that he provided advice to batteries at other sites on the island. By 1942 there were some 1450 Luftwaffe personnel who remained on Jersey, mostly responsible for anti-aircraft defences. As the largest island, Jersey was deemed particularly vulnerable to Allied airborne assault in an effort to recapture it. It was therefore well defended with a large number of anti-aircraft batteries. The Luftwaffe headquarters, however, lay not in Jersey but on the outer island, closer to the Allied threat—Guernsey.

98

 Luftwaffe Baukompanie.

196

E. P. F. Rose

References Anon. (1941) Diplomgeologenhauptprüfung. Geologische Rundschau 32:400 Anon. (1950) Bericht zur “Geoskop-Untersuchung” für die Staatsmijnen Heerlen, Holland im April-Mai 1950. Hessische Feinmechanik, Giessen, 10 pp Dear ICB, Foot MRD (eds) (1995) The Oxford companion to the Second World War. Oxford University Press, Oxford Diebel K (1941) Ein Oelschiefer im Lias Alpha bei Bielefeld. Jahrbuch der Reichstelle für Bodenforschung für das Jahr 1939 (60):157–196 Diebel K (1956) Ueber Triasconodonten. Palaeontologische Zeitschrift 30:19. [abstract only] Diebel K (1957) Juengste mesozoische Conodonten. Palaeontologische Zeitschrift 31:6–7. [abstract only] Diebel K (1960) Die palaeontologischen Originale der Berliner Museen. Palaeontol Z 34:59–60 Forty G (1999) Channel Islands at war: a German perspective. Allan Publishing, Shepperton, Surrey Gillingham J (1982) Ruhr coal mines and Hitler’s war. J Soc Hist 15:637–653 Ginns M (1993) German tunnels in the Channel Islands. Archive Book No. 7. Channel Islands Occupation Society, Jersey Ginns M (2001) The Luftwaffe on Jersey. Channel Islands Occup Rev 29:5–19 Ginns M (2012) Jersey’s German tunnels. Channel Islands Occupation Society, Jersey Ginns M, Bryans P (1978) German fortifications in Jersey. Meadowbank, Jersey Gross W, Schultze H-P (2004) Zur Geschichte der Geowissenschaften im Museum für Naturkunde zu Berlin. Teil 6: Geschichte des Geologisch-Paläontologischen Instituts und Museums der Universität Berlin 1910–2004. Mitteilungen aus dem Museum für Naturkunde in Berlin, Geowissenschaftliche Reihe 7:5–43 Häusler H (1995a) Die Wehrgeologie im Rahmen der Deutschen Wehrmacht und Kriegswirtschaft. Teil 1: Entwicklung und Organisation. Informationen des Militärischen Geo-Dienstes, Vienna 47:1–155 Häusler H (1995b) Die Wehrgeologie im Rahmen der Deutschen Wehrmacht und Kriegswirtschaft. Teil 2: Verzeichnis der Wehrgeologen. Informationen des Militärischen Geo-Dienstes, Vienna 48:1–119 Häusler H (2000) Die Österreichische und Deutsche Kriegsgeologie 1914–1918. Informationen des Militärischen Geo-Dienstes, Vienna 75:1–161 Hoepke K-P (2007) Geschichte der Fridericiana. Stationen in der Geschichte der Universität Karlsruhe (TH) von der Gründung 1825 bis zum Jahr 2000. Universitätsverlag Karlsruhe, Karlsruhe Pietrzeniuk E (1982) Kurt Diebel zum Gedenken. Mitteilungen der Gesellschaft für Geologische Wissenschaften der DDR 10:26–29 Ramsey WG (1981) The war in the Channel Islands: then and now. Battle of Britain Prints International Limited, London Rose EPF, Willig D (2013) Work by German military geologists on the British Channel Islands during the Second World War. Part 5: work by Luftwaffe geologist professor K. G. Schmidt and Hilfsgeologe Dr. K. Diebel, for tunnelling (in general and on Jersey) and water supply (on Alderney). Channel Islands Occup Rev 41:78–101 Rose EPF, Willig D (2014) Work by German military geologists on the British Channel Islands during the Second World War. Part 6: work by the Luftwaffe geologist Franz Schulte on Jersey and Guernsey. Channel Islands Occup Rev 42:152–172 Schmidt KG (1925) Geologie von Neumarkt (Oberpfalz). Ein Beitrag zur geologischen Geschichte des Frankenjuras. Dr. phil. nat. Dissertation, Albert-Ludwigs-Universität Freiburg im Breisgau Schmidt KG (1926) Geologie von Neumarkt (Oberpfalz). Ein Beitrag zur geologischen Geschichte des Frankenjuras. Berichte der naturforschenden Gesellschaft zu Freiburg im Breisgau 26:1– 120, pls 1–3

5  Jersey and the German Air Force

197

Schmidt KG (1927) Ueber die Vererzungserscheinungen im Schauinsland (Schwartzwald) Neues Jahrbuch für Mineralogie. Geologie und Paläontologie, Beilagen-Band, Abteilung A 55:163–182 Schmidt KG (1941a) Wehrgeologische Arbeiten im französischen Juragebiet. In: Anon. (ed) 6 Wehrgeologischer Lehrgang in Heidelberg. Reichsdruckerei, Berlin, pp 35–38 Schmidt KG (1941b) Über bohnerzführendes Tertiär und Diluvium im Kraichgau. Jahresberichte und Mitteilungen des Oberrheinischen Geologischen Vereins 30:48–91 Schmidt KG, Lüchtrath H (1958) Die Wirkung von frischem und gebranntem Kaolin auf die Lunge und das Bauchfell von Ratten. Hauptverwaltung der Bergbau-Berufsgenossenschaft, Bochum Schulte F (1940) Beitrag zur geologisch-bergbaulichen Forschung mit Hilfe des geophysikalischen Messgerätes “Geoskop”. Glückauf: Berg- und Hüttenmännische Zeitschrift 76:565–571 WA (1941) 5946. Beitrag zur geologisch-bergbaulichen Forschung mit Hilfe des geophysikalischen Messgerätes “Geoskop” [Contribution to geologic mining investigation with the aid of the geophysical measuring instrument “geoscope” Glückauf, 76(42):565–571, Essen, 1940.] Geophysical Abstracts, 104 [US Geological Survey Bulletin 932-A], 22 Willig D (2008) Das Wirken von Professor Dr. Wilfried von Seidlitz in den beiden Weltkriegen. Nachrichtenblatt zur Geschichte der Geowissenschaften 18:64–91 Wilser J (1921) Grundriss der angewandten Geologie unter Berücksichtigung der Kriegserfahrung für Geologen und Techniker. Borntraeger, Berlin Wilser J (ed) (1923–1929). Die Kriegsschauplätze 1914–1918 geologisch dargestellt in 14 Heften. Borntraeger, Berlin Wilser J (1941) Die wehrgeologischer Karte 1:300,000. In: Anon. (ed) 6 Wehrgeologischer Lehrgang in Heidelberg. Reichsdruckerei, Berlin, pp 125–126

Chapter 6

Guernsey and the German Army Edward P. F. Rose

Abstract  Support for the fortification of Guernsey was provided by five geologists of the German Army’s military geological service, each with the status (equivalent to major or captain) of a Technischer Kriegsverwaltungsrat (TKVR): a uniformed ‘Technical War Administration Officer’. A report with military engineering geology and water-supply maps for the island was compiled in November 1941 by TKVR Walter Wetzel, a former professor leading a military geology team (Wehrgeologenstelle 9) based in Paris. Site investigations were subsequently initiated by visits from TKVR Friedrich Röhrer, a Heidelberg professor serving in Paris as the geologist with the Army’s Inspectorate of Land Fortification (West) who generated three reports, and by his assistant TKVR Scherer, who generated another two. From May to December 1942 a military geology team (Wehrgeologenstelle 4) was based on Guernsey, led by TKVRs Bernhard Beschoren and Dieter Hoenes and including Lance-Corporal Dr. Gottfried Reidl. Its geotechnical output included at least seven major reports and seven thematic maps for the island: its first geological maps at 1:25,000, plus innovative maps for military geology, raw materials, and groundwater—documents that facilitate a detailed case history illustrating the work of one of the 32 Wehrgeologenstellen then established within the German Army. Additionally, Lance-Corporal Dr. Rolf Thienhaus contributed geological expertise to tunnelling projects (whilst serving within a Mining Engineer Company), and a map of coastal features was generated in 1943 for the Navy’s headquarters (in Paris), apparently by Wehrgeologenstelle 6.

E. P. F. Rose (*) Department of Earth Sciences, Royal Holloway, University of London, Egham, Surrey, UK e-mail: [email protected] © Springer Nature Switzerland AG 2020 E. P. F. Rose (ed.), German Military Geology and Fortification of the British Channel Islands During World War II, Advances in Military Geosciences, https://doi.org/10.1007/978-3-319-22768-9_6

199

200

E. P. F. Rose

6.1  Introduction German forces established their overall command of the Channel Islands on Guernsey, the habitable island closest to Great Britain and so potentially the island most vulnerable to a British attempt at re-capture. Fortification was implemented at the same time and in the same way as on Jersey, but geological guidance for construction work for the German Army was provided differently. Rather than a single geologist appointed to serve long-term as an officer actually within a Fortress Engineer Staff, on Guernsey five geologists of officer status were used, for shorter assignments in support of the local Fortress Engineer Command or Staff. Each of these held appointment as a ‘Technical War Administration Officer’ (Technischer Kriegsverwaltungsrat) (TKVR in common abbreviation)—so rank equivalent to that of major or captain but authority to provide technical advice rather than to take command decisions (see Sect. 3.5). In total, the five TKVRs generated 13 geological reports that are currently preserved in the Bundesarchiv-Militärarchiv at Freiburg-­ im-­Breisgau (Table 6.1) and nine thematic maps (Table 6.2) that are also preserved in Germany or in the USA. From the summer of 1941 and so essentially the same time that Walther Klüpfel became active on Jersey, the German Army’s fortification work on Guernsey was assisted by geological expertise contributed from headquarters based in Paris. Initially this was done remotely, soon followed by short-term site investigations, by geologists like Klüpfel over 50 years in age and of established academic standing, but men who had contributed to Germany’s war effort from the start of hostilities. The careers and roles of the two geologists primarily involved, professor TKVRs Walter Wetzel and Friedrich Röhrer, are illustrative of geologists who prospered under Germany’s National Socialist regime. Like Klüpfel, Wetzel was a veteran military geologist of World War I. Röhrer too had served in that war, although not as a military geologist, but postwar in employment at the University of Heidelberg was subject to the influence of veteran military geologists and of leadership enthusiastic in its support for Germany’s Führer Adolf Hitler. From the spring of 1942, and for the rest of that year, one of the German Army’s then 32 military geological centres/teams (Wehrgeologenstelle 4) was deployed longer-term to the Channel Islands, for work on both Guernsey and Alderney. Led by TKVRs Bernhard Beschoren and Dieter Hoenes, the documentary output of this unit illustrates the value of such units to German troops occupying ‘enemy’ terrain as their use approached its peak: the number of Wehrgeologenstellen was increased to 35 in September 1943, and finally to 40 in November (see Sect. 3.5). Thereafter as German forces generally retreated rather than advanced, their need for geologists correspondingly declined. Tunnels were as much a feature of the fortification plan on Guernsey as they were on Jersey, and one of the Mining Engineer Companies at work tunnelling on Guernsey made use of a trained geologist, Dr. Rolf Thienhaus—albeit at that time only in the rank of lance-corporal: as a young man, enlisted on completion of his doctoral thesis.

6  Guernsey and the German Army

201

Table 6.1  ‘Reports’ generated by German Army geologists for Guernsey, numbered sequentially here for convenience of reference in the text, with translated title, author, affiliation, date, page length and current file number at the Bundesarchiv-Militärarchiv 1. Wehrgeologische Beschreibung der Insel Guernsey [Military geological description of the island of Guernsey]. Sachbearbeiter TKVR Prof. Dr. Wetzel. (Oberfestungsbaustab beim Komm. Adm. Frankreich. Wehrgeologenstelle 9. Paris.) 3 Nov 1941 [2 pp., 2 maps at 1:25,000]. [RH32v.3041, also v.3082] 2. Untersuchung der Inseln Guernsey und Alderney, die Trinkwasserversorgung der auf Landzungen vorgeschobenen Stellung betrefend [Investigation of the islands of Guernsey and Alderney: concerning the supply of drinking water to headland positions] by KVR Röhrer, for Insp. Land. O. West. 19 Dec 1941. [3 pp.] [RH32v.3041] 3. Wasserversorgung der Batterie Nina [Water supply to Battery Nina], by KVR Röhrer, for Inspekteur der Landesbefestigung beim Oberbefehlshaber West. 19 Dec 1941. [2 pp.] [RH32v.3041] 4. Wasserversorgung der Batterie ‘Mirus’ [Water supply to Battery Mirus] by TKVR Scherer. 17 Jan 1942. [2 pp.] [RH32v.3041] 5. Erkundung von Wasserversorgungsmöglichkeiten, unabhängig von vorhandenen Versorgungssystem [Exploration of possible water supplies additional to the existing supply system] by TKVR Scherer, for Fest. Pi. Stab 19. 26 Jan 1942. [RH32v.3041] 6. Versorgung von Gustav mit Bentonzuschlagstoffen [Supply of aggregate for concrete on ‘Gustav’, i.e. Guernsey] by KVR Röhrer. Insp. d. L. West – Wehrgeologe – Az. 39 Geol. 1o/e Br. B. Nr. Geh. 36/42. 4 Apr 1942. [2 pp.] [RH32v.3041] 7. Gutachten 2. Gutachten über die zusätzliche Wasserversorgung der Batterie Mirus. [Expert opinion on the supplementary water supply of Battery Mirus] 1oc Br. B. Nr. 57/42 geh, 17 Jun 1942, by Beschoren, 2 pp., 1 fig. [RH32v.3028, also v.3041] 8. Gutachten 3. Gutachten über den Denkeneinbruch in Verpflegungsstollen 1 (Gaswerkstollen). [Expert opinion on the roof collapse in Ration Store Tunnel 1 (Gasworks Tunnel)]. 1ob Br. B. Nr. 58/42 geh, 23 Jun 1942, by Beschoren, 2 pp., 1 fig. [RH32v.3029] 9. Gutachten 7. Gutachten über die Schaffung eines panzerischeren Geländes landeinwarts der Vazon-Bay. [Expert opinion on the creation of an anti-tank terrain inland of Vazon Bay] 1ob Br. B. Nr. 76/42 geh, 22 Aug 1942, by Beschoren, 3 pp., 2 figs, annex 4 pp. [RH32v.3029] 10. Geplante Wasserwerke der Luftwaffe. [Waterworks planned by the Air Force] 5 September, by Beschoren, 3 pp. [RH32v.3041] 11. Wasserversorgung im Gebiet des verstärkten Küstenausbaus. [Water supply in the region of coastal fortification] 23 September, by Beschoren, 2 pp. [RH32v.3041] 12. Gutachten 8. Erläuterungen zur Baustoffkarte der Kanalinsel Guernsey. [Notes for the raw materials map of the Channel Island of Guernsey] 1oe Br. B. Nr. 91/42 geh, 29 Sep 1942, by Beschoren and Hoenes, 18 pp., 1 map [RH32v.3027] 13. Wasserversorgungsmöglichkeiten der Kanalinseln Guernsey. [Water supply facilities on the Channel Island of Guernsey] Fest. Pi. Stab. 19, 14/21 Dec 1942, by Beschoren, 19 pp., 2 maps [RH32v.3041]

202

E. P. F. Rose

Table 6.2  Thematic maps prepared by German military geologists for Guernsey, with authorship where known and present location; all at scale of 1:25,000. Sequential numbering adopted here is for convenience of reference in the text 1. Wehrgeologische Karte, by Wetzel: Bundesarchiv-Militärarchiv, file RH32v.3082 2. Wasserversorgungs Karte, by Wetzel: Bundesarchiv-Militärarchiv, file RH32v.3082 3. [Untitled] geological map (solid), by Wehrgeologenstelle 4: US National Archives and Records Administration 4. [Untitled] geological map (drift), by Wehrgeologenstelle 4: US National Archives and Records Administration 5. Wehrgeologische Karte, by Wehrgeologenstelle 4: US National Archives and Records Administration 6. Baustoffkarte, by Wehrgeologenstelle 4: Bundesarchiv-Militärarchiv, file RH32v.3027 7. Grundwasserkarte, by Wehrgeologenstelle 4: Bundesarchiv-Militärarchiv, file RH32v.3041 (draft as Brunnentiefen innerhalb des Gebietes des verstärkten Küstenausbaues: US National Archives and Records Administration) 8. Grundwasserkarte der befestigen Küstenzone, by Wehrgeologenstelle 4: Bundesarchiv-­ Militärarchiv, file RH32v.3041 (draft as tracing overlay Grundwasserkarte im Gebiet des verstärkten Küstenausbaues: US National Archives and Records Administration) 9. Karte der Wasserversorgungsmöglichkeiten, by Wehrgeologenstelle 4: Bundesarchiv-­ Militärarchiv, file RH32v.3041

6.2  Walter Wetzel and Wehrgeologenstelle 9 at Paris The earliest military reference so far known to German geological work on the Channel Islands as a whole is a one-page unsigned typed memorandum1 preserved in Germany at the Bundesarchiv-Militärarchiv in Freiburg (Rose 2005a). Annotated with the date of August 1941 (by Dr. B. Beschoren in April 1942), this opens with the statement that solutions to geological problems on the Channel Islands may be sought from ‘Prof. Wetzel für Guernsey, Prof. Klüpfel für Jersey und Alderney’, and that Professor Wetzel had overall responsibility for geological matters on the islands.

6.2.1  The Geologist Walter Wetzel Rose (2005a) has identified ‘Prof. Wetzel’ as Walter Wetzel (1887–1978) (Fig. 6.1). He was born  and named Konrad Alois Siegmund Walter Wetzel on 27 February 1887 in his mother’s home city of Hannover,2 in northern Germany (Seibold 1978; Dietz et al. 1999). His father was a merchant, but came from an old mining family rooted in the Hartz Mountains, so the love of geology was planted in him from the

 Untitled five-point memorandum, dated August 1941 at top, annotated at bottom by Beschoren with date of April 1942; within file RH32v.3041. 2  Hanover in English spelling. 1

6  Guernsey and the German Army

203

Fig. 6.1  Walter Wetzel (1887–1978), a photograph taken late in life. From Rose (2005a): published by kind permission of the Bibliothek, Institut für Geowissenschaften der Universität Kiel

cradle—or so it has been claimed. He began formal study of science and mathematics in 1905, first at the technical university in Hannover, then at universities in Jena and Munich, and finally at Göttingen, where he received his doctorate in 1910—for a thesis on Jurassic strata of the Teutoburger Wald, a region near to the town of Bielefeld (Wetzel 1911). Wetzel began his geological career in 1912 as an assistant at the University of Kiel, also in northern Germany. He married in 1913 and the first of his three children was born in 1914. However, in the spring of 1915 he was conscripted for service in World War I, and in the summer of that year sent to the front as a member of an infantry regiment. After various promotions, in 1917 he became a military geologist, and served widely as such until demobilization in December 1918. Häusler (2000) credits him with being the deputy3 and subsequently the leading military geologist4 at the Army’s military geology centre/team number 6.5 On demobilization, and after the death of both parents, Wetzel returned to Kiel to find that his former university position was now occupied by someone else, and that his private property had been confiscated. With university and scientific appoint Beamtenstellvertreter.  Kriegsgeologe, literally ‘war geologist’. 5  Geologischen Beratungsstelle VI. 3 4

204

E. P. F. Rose

ments then very few, and needing to support a growing family, Wetzel thereafter taught at secondary schools in Kiel, but from 1922 also as a lecturer at the University of Kiel. He seized opportunities to travel widely: to Scandinavia, France, Spain, Morocco and even South America. In 1939 he was awarded the title of ‘professor’, having over 80 publications to his credit by that time, on topics ranging across many fields of geology, from palaeontology to mineralogy (as listed by Dietz et al. 1999). In World War II, Wetzel quickly became a military geologist again. From 1939 to 1940 he served on the construction staff6 at Rendsburg (a town approximately 30 km west of Kiel) in the Northwest Air Defence Zone7 (Häusler 1995b), so in support of the Luftwaffe. However, during the early stages of the German occupation of France, he served with one of the five military geology ‘groups’ that at that time supported the German Army on the Western Front (as illustrated in Sect. 3.5)—the ‘group’ based in Paris.8 He served first (from September 1940) actually at Paris in support of the headquarters staff of the military administration for France; then (from October) in the group’s outstation9 at Rouen as part of the Army’s ‘Engineer Reconnaissance of France’;10 and finally (from late 1940 into 1941) as a member of the military geology centre/team for the Channel coast.11 At the beginning of 1940 (30 January to 6 February) Wetzel was one of at least 19 participants in a course for military geologists held at Giessen, Walther Klüpfel’s home town—the second of a series of six such courses held at various centres during the year (Häusler 1995a). At the end of the year (14 to 20 December) he was one of at least 31 participants to attend the sixth course, at Heidelberg. There he was seemingly one of the most active contributors, giving three papers subsequently published in the post-conference book. These dealt critically with military geological work on the Channel coast (Wetzel 1941a), the geology of airfield construction (Wetzel 1941b), and aspects of trench construction and water supply in sandy upland areas (Wetzel 1941c). Wetzel thus seems to have been particularly active and successful in developing his role as a military geologist throughout the year. By April 1941 he was back in Paris, as leader of Wehrgeologenstelle 9: one of the German army’s 25 military geology centres/teams created at about that time by the re-organization of the former five larger ‘groups’ on the Western Front (as described in Sect. 3.5). This team was assigned to the headquarters of the Admiral of France,12 presumably to assist with aspects of coastal fortification. From October assignment was specifically via its senior fortress construction staff.13

 Ausbaustab.  LVZ (Luftwaffen Verteidungs Zone) Nordwest. 8  Wehrgeologengruppe Paris. 9  Ausenstelle 6. 10  PiErkStab Frankreich. 11  Wehrgeologenstelle Kanalküste. 12  Kommandierenden Admiral von Frankreich. 13  Oberfestungsbaustab. 6 7

6  Guernsey and the German Army

205

6.2.2  Wetzel and Guernsey It was from Wehrgeologenstelle 9 at the Admiral’s fortress construction staff14 in Paris that Wetzel sent two-page military geological descriptions first for Alderney (see Sect. 8.2) and then Guernsey (Table 6.1, item 1),15 the latter dated 3 November 1941. Authorship is shown as ‘TKVR Prof. Dr. Wetzel’: TKVR appointment when over 35 years of age gave him status equivalent to an Army major (see Sect. 3.5). The Guernsey description comprises two pages of typescript summarizing knowledge of the island’s general rock conditions (with regard both to ‘hard’ bedrock and to superficial deposits), supply of raw materials, water supply and ground excavation for fortifications. It is accompanied by two thematic maps. These have been drawn on a topographical base map at 1:25,000-scale produced by the German army map service16 from the three-inches-to-one-mile (1,21,120) map published in 1934 by the British Ordnance Survey. This edition was seemingly produced in haste: captioned ‘4  cm Karte’, it reproduced the Ordnance Survey map exactly, including the words ‘Ordnance Survey’, at the new scale. A military geological map17 (Table 6.2, item 1; and Fig. 6.2) distinguishes rock types with geological boundaries that closely match those depicted by Parkinson and Plymen (1929) (see Sect. 2.4 and Fig. 2.22). Wetzel thus seems to have transcribed them from their figure. However, the three-part legend (Table 6.3) inked on the map shows how the geological data have been interpreted for German military use. Essentially, six rock types are shown on the map, interpreted in terms of: 1. three categories for potential tunnelling conditions, by means of coloured ornament; and 2. four categories as potential sources of raw materials for construction purposes, by means of a superimposed symbol. The map therefore seems to be the product of a desk study rather than of a field-­ based survey, and not strictly to be a ‘military geology’ map but a simplified geological map which attempts to interpret two features of engineering significance (tunnelling conditions and potential sources of raw materials) on one sheet. It is, however, seemingly the first map of this type to be compiled for any part of the British Isles, certainly the first at 1:25,000 scale for widespread use. A water-supply map18 (Table 6.2, item 2; and Fig. 6.3) has a key (Table 6.4) that distinguishes seven categories of engineering works and two categories of river valleys. The map thus summarizes the water-supply system then current on the island, with little indication of groundwater conditions and none of geology. Its features are  Oberfestungsbaustab beim Komm. Adm. Frankreich.  Wehrgeologische Beschreibung der Insel Guernsey. Sachbearbeiter TKVR Prof. Dr. Wetzel. (Oberfestungsbaustab beim Komm. Adm. Frankreich. Wehrgeologenstelle 9. Paris.) 16  Druckereizug 14 Feldpostn. 06661. 17  Wehrgeologische Karte. 18  WasserversorgungsKarte.

14 15

206

E. P. F. Rose

Fig. 6.2  Wehrgeologische Karte. Military geological map of Guernsey, original at scale of 1:25,000, on ‘first edition’ of German base map, compiled by Walter Wetzel, November 1941. For key, see Table 6.3. From Rose (2005b): reproduced by permission of the Bundesarchiv-­Militärarchiv from file RH32/3082

likely to have been derived from study of topographical rather than geological maps. It is not a hydrogeological map of the kind being devised at about the same time by Walther Klüpfel for Jersey, based on data collected by fieldwork. The report and its accompanying maps were presumably of value for the early stages of planning, but were soon to be superseded by accounts based on ground investigation rather than desk study.

6.3  F  riedrich Röhrer and the Inspectorate of Land Fortification (West) As shown for Jersey (in Sects. 4.3 and 4.7), provision of an adequate water supply for military sites was a task that could involve geological expertise. On Guernsey, by far the most important site soon under construction was a massive coastal artillery battery, named ‘Nina’ at first but later re-named ‘Mirus’. Two water-supply reports were compiled in December 1941, one for water supply on Guernsey and Alderney in general, the other specifically for the Nina/Mirus battery site. Both were by TKVR Professor Dr. Röhrer, the military geologist at the Inspectorate of Land Fortification (West). They were quickly followed by two ­further water-supply

6  Guernsey and the German Army

207

Table 6.3  Translation of the three-part key to the map shown on Fig. 6.2, top to bottom in each case Rock types (top left of map): Brown diagonal lines Loam Orange Shoreline gravel Yellow with brown crosses Sand (dunes) Yellow Sand (intertidal) Yellow with diagonal lines Sand (subtidal) Brown Sandstone Thin blue diagonal lines Micaceous schist Dark red Massive stone (granite, diorite and gabbro) Horizontal red lining Massive stone (gneiss) Tunnelling conditions (right corner of map, top): Brown diagonal lines Area of firm loam covering hard rock Dark red Hard rock with greatest resistance to working Horizontal red lines Hard rock with some resistance to working Brown Hard rock with some resistance to working Thin blue diagonal lines Hard rock with some resistance to working Raw materials (right corner of map, bottom): Brick ornament within a circle Hard rock (building stone) Two ovals within a circle Aggregate production Three dots within a circle Building sand Circle divided by vertical line Brickworks

reports, in January 1942, one again for the battery, by TKVR Scherer: a military geologist of unspecified address but seemingly a junior colleague of Röhrer.

6.3.1  The Geologist Friedrich Röhrer Rose (2005a) has identified Röhrer (Fig. 6.4) as Friedrich Röhrer (1885–1945) who had been born on 8 September 1885 at Baden-Baden (Drüll 2013), a spa town NE of Strasbourg, in the Black Forest region bordering the River Rhine in SW Germany. A student at the University of Heidelberg (80 km to the north) from 1904, he passed the examination that qualified him for employment as a high school teacher19 in 1909 and taught first at Heidelberg, then from 1910 to 1916 at Pforzheim (about 60 km south of Heidelberg, and some 60 km to the NE of Baden-Baden). He was awarded a doctorate20 at Heidelberg on 2 May 1916 for a geological investigation of the relationships between rock fractures, tectonics and the hydrographic network of the northern Black Forest (Röhrer 1916, 1922a, b). However, on graduation, he was

19 20

 Oberrealschullehrer.  Dr. phil. Nat.

208

E. P. F. Rose

Fig. 6.3  Wasserversorgungs Karte. Water-supply map of Guernsey, original at scale of 1:25,000, on ‘first edition’ of German base map, compiled by Walter Wetzel, November 1941. For key, see Table 6.4. From Rose (2005b): reproduced by permission of the Bundesarchiv-Militärarchiv from file RH32/3082

Table 6.4  Translation of key to the map shown on Fig. 6.3, top to bottom

Blue spot

Pumping stations in pools (former quarries) Double blue circle Large pumping stations (associated with wells) Single blue circle Small pumping stations (associated with wells) Broken circle Pumping station not available (at the airfield) Thick blue line Major water distribution system Purple line Example of a subsidiary system Double red lines Dam under construction Green line Valleys with a surface water course Brown line Valleys with an underground water course

6  Guernsey and the German Army

209

Fig. 6.4  Friedrich Röhrer, with students. From Rose (2005a): photograph courtesy of the late Professor Eugen Seibold, Geoarchiv, Universität Freiburg

called up in 1916 for military service, and continued to serve until the close of hostilities in 1918. Not listed by Häusler (2000) amongst the military geologists of that period, he presumably made little if any use of his geological expertise during World War I. Röhrer returned to teaching after the war, from 1919 at Schwetzingen until appointed to a school21 in Mannheim, both schools  near Heidelberg, in 1920. However, on 10 May 1920 he was granted qualification for teaching at university level22 by the University of Heidelberg. In the next few years he published articles on water supply (Daur and Röhrer 1921, Röhrer 1923a, b) and on local geology and geomorphology (Röhrer 1924, 1925). These gained him appointment as a junior professor23 at Heidelberg in 1925. Later publications included another on geology/ geomorphology (Röhrer 1927); on water supply (Röhrer 1929); contributions to  Realgymnasium: a type of secondary school that prepares students for university entrance, typically with emphasis on the sciences and modern languages. 22  Habilitation: the qualification necessary to conduct self-contained university teaching, and so a preliminary step towards appointment as a professor. 23  Professor a. o. [ausserordentliche] = ‘extraordinary’ professor: without a ‘chair’ and so junior to an ‘ordinary’ professor with a ‘chair’. 21

210

E. P. F. Rose

local geological mapping (Brill 1932); a paper on nitrates in groundwater (Röhrer 1933); and one on groundwater and mineral springs (Röhrer 1934). He was appointed a ‘staff geologist’24 and so with military responsibilities on 17 July 1938, and as a supernumerary professor25 within the University on 3 November 1939— presumably so as to retain his academic status but free him for military use from routine University duties. At the start of World War II, Röhrer was thus a well-established geologist and hydrogeologist, in a professorial appointment at Heidelberg: a university notable for its intensive nazification during the 1930s (Remy 2002; Eckart et  al. 2006). Moreover, Heidelberg’s Geology Institute was one of the departments engaged on war-related research before the outbreak of war: its director, Julius Wilser, and three assistants were working as geologists for the armed forces. Its geologists taught courses that included ‘Geology of the western theatres of war’ (Ernst Becksmann) and ‘The geological structure of Europe, its theatres of war, and its war- and economic-­ related deposits’ (Julius Wilser) (Remy 2002, p  95, 96). Wilser (1888–1949), promoted to a tenured chair at Heidelberg after a more junior position at the University of Freiburg (see Sect. 5.2), had served as the senior German military geologist of World War I. Presumably through his influence, Heidelberg was the venue for a conference of military geologists in December 1940, and from 1940 the site (with Vienna in Austria) of one of the two ‘information centres’ established to support the German Army’s military geological teams by provision of geological maps and literature (Wilser and Becksmann 1941; Häusler 1995a). Presumably also, it was the veteran Wilser who helped to influence his colleagues Röhrer and Becksmann to become active in military geology. Like Wetzel, Röhrer’s war service began early, and as a geologist in support of the Luftwaffe. Indeed, in Röhrer’s case from 1938, as one of the five ‘sector geologists’ in the ‘Air Defence Zone26 West’, and a participant in their field trip held in 1939 (Häusler 1995b). Subsequently, again like Wetzel, Röhrer transferred into the Army: in 1940 he became the leader of a military geologist group within the Army’s ‘Engineer Reconnaissance of France’. Initially supporting the German First Army Headquarters,27 this transformed into one of the five military geologist groups28 on the Western Front as the occupation of France began (cf. Sect. 3.5). In December 1940 he presented a paper (Röhrer 1941) on military aspects of the Triassic rocks of Lorraine, at the Heidelberg conference: so at his home university, and at a conference to which his more senior Heidelberg colleague Julius Wilser also contributed (Wilser 1941). On transformation of the five ‘groups’ into 25 smaller centres or teams (Wehrgeologenstellen) in April 1941, Röhrer became leader of Wehrgeologenstelle 4, at First Army Headquarters. He was subsequently promoted to serve as the mili Stabsgeologe.  Professor apl. [ausserplanmässiger] = supernumerary (i.e. unpaid) professor. 26  LVZ [Luftwaffen Verteidigungs Zone]. 27  AOK1 [Armee Ober Kommando 1]. 28  Wehrgeologengruppe Nancy. 24 25

6  Guernsey and the German Army

211

tary geologist29 at the Inspectorate of Land Fortification of the Commander-in-Chief West, from December 1941. There one of his first tasks was to deal with water-­ supply problems on Guernsey.

6.3.2  Röhrer and Water Supply on Guernsey Röhrer generated two reports dated 19 December, both written from ‘St. Qu.’ (i.e. St. Ouen in Jersey, so whilst he was actually in the Channel Islands), and both at the order of and so addressed to the Inspector of Land Fortification at the Commander-­ in-­Chief West.30 1. A three-page typescript concerned the supply of drinking water to fortifications on headlands of the islands of Guernsey and Alderney (Table  6.1, item 2).31 Röhrer deduced that in places near the coast there was a chance of freshwater overlying sea water in crevices within the bedrock, but that the amounts available were likely to be small. In consequence, best options for water supply on headland sites were either (a) to link them with the main water distribution network for the island as a whole, or (b) to supply them by means of a bowser from new or existing standpipes or intakes in the near vicinity. On Guernsey he inferred that would not be a major problem, since there was already a major ring-main system. 2. A two-page typescript concerned water supply for the massive battery of coastal artillery ‘Nina’ then under construction on Guernsey (Table 6.1, item 3).32 From November 1941, this site had been given absolute priority in the proposed 14-month fortification programme (Partridge and Wallbridge 1983). First surveys to site a heavy naval battery had been conducted in the late summer of 1941, and from November work began in earnest on an extensive arc of land, to accommodate four 48-ton guns of 305 mm calibre (Fig. 6.5): the most formidable on the Channel Islands. Ultimately some 47,000  m3 of concrete were to be consumed in the construction of this battery position by mid 1942, an amount that alarmed the local German command and proved a significant drain on material and manpower. To provide water to the site, Röhrer suggested three possibilities: (a) Connection to the existing water-supply network, by about 100 m of pipework. This was certainly the most convenient method, but carried a high risk

 Wehrgeologe.  Inspekteur der Landesbefestigung beim Oberbefehlshaber West. 31  Untersuchung der Inseln Guernsey und Alderney, die Trinkwasser-versorgung der auf Landzungen vorgeschobenen Stellungen betreffend by KVR Röhrer, for Insp. Land. O. West. 32  Wasserversorgung der Batterie Nina by KVR Röhrer, for Inspekteur der Landesbefestigung beim Oberbefehlshaber West. 29 30

212

E. P. F. Rose

Fig. 6.5  One of the four gun emplacements for the Mirus (formerly Nina) Battery of NW Guernsey (cf. Fig. 6.8). Translation of the nine items listed top down in the German key shows that each site had its own shell store, cartridge store, ventilation system, engine room, fuel store, heating system, washroom/lavatory, accommodation (for 72 men) and site entrance. Outer walls were of reinforced concrete 2 m thick; the bunker complex measured some 31 by 31 m and the gun barrel was 15.8 m in length. From Robins et al. (2012); originally from Festung Guernsey, a record of fortifications prepared by the German armed forces during their period of occupation, by permission of the Priaulx Library and the Royal Court, Guernsey

that the supply would be put out of action if the battery came under bombardment. (b) Connection to a spring in a nearby valley. The implications of this possibility were discussed in greater detail. (c) Following the precedent of the numerous shallow wells in the vicinity that supplied groundwater to the local nurseries, it would be possible to adopt shallow wells for the battery as well. The water table was recorded at a depth of 2 to 3 m, so wells would need to be emplaced to a depth of 6 to 7 m. In conclusion, Röhrer recommended that to avoid the risk of interruption to supply by bombardment, each of the four guns in the battery should have its own water supply, from a shallow well to be emplaced to the depth recommended. However, as a backup facility, the nearest spring should also be tapped in the manner outlined as the second possibility, to allow water to be accessed from a standpipe in an emergency.

6  Guernsey and the German Army

213

6.3.3  TKVR Scherer and Water Supply on Guernsey It seems that Röhrer then delegated more detailed work to a junior colleague, Scherer, so far known only from his surname and TKVR status. Häusler (1995b) records that Scherer began service as a military geologist later than Röhrer, in 1940. His first recorded appointment was to the military geology ‘group’ supporting the ‘Engineer Reconnaissance of France’ (formerly the headquarters of the First Army),33 as a member of that group’s sub-unit at Montmedy, later that at Saarlautern.34 Subsequently in 1940, when the organization was re-structured, his unit became the Saarlautern outstation of the military geology group of the Engineer Reconnaissance Staff based at Nancy.35 Scherer had thus served within the two military geology ‘groups’ led during 1940/1941 by Röhrer, and from March 1942 he was assigned to a military geology centre/team (Wehrgeologenstelle 5) that was based, like Röhrer, at the Inspectorate of Land Fortification (West). Scherer completed two water-supply reports for Guernsey in the month following those completed by Röhrer: 1. On 17 January 1942, a two-page typescript on water supply for Battery ‘Mirus’ (Table  6.1, item 4),36 essentially developing Röhrer’s recommendation that drinking water be supplied by pumping from the nearest spring. 2. On 26 January, on Guernsey’s potential water supplies independent of the existing mains system (Table 6.1, item 5).37 Prepared for Fortress Engineer Staff 19, the equivalent unit on Guernsey to the Fortress Engineer Staff 14 on Jersey to which Klüpfel belonged, his report was essentially a catalogue of 39 springs and 24 wells, whose locations were plotted on an accompanying 1:25,000-scale topographical map. It was in no way a hydrogeological map of the kind devised by Klüpfel for Jersey that same month, and was to be superseded later in the year by the more extensive studies of Bernhard Beschoren and Dieter Hoenes (described in Sect. 6.8).

6.3.4  Röhrer and Raw Materials on Guernsey The pace of geological activity quickened in 1942, following a visit between 24 March and 2 April to the Channel Islands and adjacent areas of the French mainland led by a TKVR Lautmann. Like Röhrer, Lautmann was also on the staff of the  WG-Gruppe PiErkStab Frankreich (zuvor AOK1).  WG-ErkTp Montmedy, später WG-ErkTp Saarlautern.. 35  WG-Gruppe Nancy PiErkStab Frankreich, Aussenstelle Saarlautern, later WG-Gruppe Nancy, Aussenstelle 2, Saarlautern. 36  Wasserversorgung der Batterie ‘Mirus’ by TKVR Scherer. 37  Erkundung von Wasserversorgungsmöglichkeiten, unabhängig von vorhandenen Versorgungssystem by TKVR Scherer, for Fest. Pi. Stab 19. 33 34

214

E. P. F. Rose

Inspectorate of Land Fortification (West).38 Since he is not listed as a geologist by Häusler (1995a, b), presumably he was an engineer. However, he was accompanied by his geologist colleague, Röhrer, and for the Channel Islands themselves, by two additional military geologists—TKVRs Beschoren and Hönes [sic] of Wehrgeologenstelle 4—according to his four-page subsequent report.39 Members of the group visited St. Malo in France at the start of their reconnaissance, followed by Guernsey on 27 and 28 March, Alderney on 29 March, Cherbourg and Granville in France on 31 March and Jersey on 1 April. Topics of interest on Guernsey comprised water supply, drainage, tunnelling, anti-tank walls and the Battery Mirus. Klüpfel recorded that he took Röhrer and Beschoren on an excursion in Jersey on 1 April (see Table 4.1), so apparently during the Lautmann-led visit. Lautmann’s report indicates that Röhrer and Beschoren were absent from his group that day. It is evident, therefore, that by the start of April, communication had been established between all the senior military geologists then active on the Channel Islands. Röhrer generated a two-page report on the same day as Lautmann’s: 4 April. This dealt with potential supplies on Guernsey of aggregates suitable for use in concrete manufacture (Table 6.1, item 6).40 Like Scherer’s water supply report of 26 January, Röhrer’s raw materials account was also to be superseded later in the year, by the more extensive work of Beschoren and Hoenes (see Sect. 6.7).

6.4  W  ehrgeologenstelle 4: Bernhard Beschoren, Dieter Hoenes and Gottfried Reidl The centre/team Wehrgeologenstelle 4 had come into existence in April 1941, as part of the overall re-organization of German military geology units (see Sect. 3.5). Led initially by Friedrich Röhrer, it was active in central France, generating reports on water supply for a region near Besançon in August 1941, on water supply at Le Valdahon in October, and on raw materials for concrete along the River Seine in November (Häusler 1995a, p. 114). However, by late March 1942 the unit was evidently being led by Bernhard Beschoren, and was ready for assignment to support Fortress Engineer Command XIV on the Channel Islands. During the rest of that year members of Wehrgeologenstelle 4 compiled ten numbered ‘expert opinions’41 for Alderney and Guernsey and at least 11 unnumbered reports or memoranda. All were to be issued either in the name of TKVR Beschoren or of his deputy, TKVR Hoenes, seven of them for Guernsey (Table 6.1, items 7 to 13).

 Inspekteur der Landesbefestigung beim Oberbefehlshaber West.  Die Besichtigungsreise zu den Kanalinseln vom 24.3 bis 2.4.42, of 4 April 1942: now preserved at the Bundesarchiv-Militärarchiv in file RH32v.3041. 40  Versorgung von Gustav mit Bentonzuschlagstoffen, by KVR Röhrer. Insp. d. L.  West  – Wehrgeologe. (‘Gustav’ was the contemporary code name for Guernsey.) 41  Gutachten.

38

39

6  Guernsey and the German Army

215

6.4.1  The Geologist Bernhard Beschoren Bernhard Beschoren (Fig.  6.6) was born on 6 December 1898  in Altlüdersdorf / Kreis Ruppin, according to his daughter, Renate Axthammer (Rose 2007). He was thus a native of the province of Brandenburg that surrounds the city of Berlin in NE Germany, and over 40 years old at the start of the war. He had attended school in Potsdam and Berlin-Wilmersdorf before service in World War I from 1916 to 1918, from the age of 18.42 Thereafter he studied natural science, principally geology, at the universities of Berlin, Freiburg and again Berlin, where he gained his doctorate in August 1926 with a thesis43 on the Upper Cretaceous (Chalk) in northern Germany (Beschoren 1926, 1927). The following year both he and Walter Wetzel coincidentally were amongst the 20 former students to contribute to a publication honouring one of their former university teachers (Beschoren et al. 1927). From 1927 to 1938 he followed a career at the Berlin-based Prussian Geological Survey, finally as a regional geologist, where he apparently worked primarily on Quaternary geology. Apart from a compendium of petroleum geology literature (Beschoren 1934), his principal Survey publications dealt with alluvium near Hannover (Beschoren 1932); Recent deposition of marsh peat in the Elbe and Oder river regions of Germany Fig. 6.6 Bernhard Beschoren, in 1956. From Rose (2007), photograph published courtesy of his daughter, Renate Axthammer

42 43

 www.pgla.de accessed 17 February 2016.  Cenoman und Turon der Kreidemulde von Sack bei Alfeld.

216

E. P. F. Rose

(Beschoren 1935a); post-glacial deposition in the Havel river valley (Beschoren 1935b); correlation of late Quaternary varve clays in the Havel-Oder river region (Beschoren 1935c); Pleistocene drainage changes of the Leine and Aller rivers in the Neustadt-Celle region (Beschoren 1936); and eskers44 in a region east of Schwerin (Beschoren 1937). His daughter recalled (Rose 2007) that in 1938 he declined to take an oath of allegiance to the German Führer Adolf Hitler, so had to give up his Berlin appointment. In 1939 he joined an American petroleum company, based at its office in Copenhagen, Denmark. However, after the entry of German troops into Denmark on 9 April 1940, the family was required to move back to Berlin, and Beschoren was quickly conscripted for military service. Häusler (1995b, p. 10, 60) records that by September 1940 Beschoren was serving as a member of the geological team45 based at Pulawy in Poland. The team comprised one of the five teams within one of the three military geology groups at that time deployed on the Eastern Front. According to Häusler (1995a), by early 1942 Beschoren had been assigned to Wehrgeologenstelle 29, a unit raised in October 1941 which from early 1942 at least was responsible to the commander of Land Works Section number 146 but which developed later into the larger Military Geological Staff47 supporting the High Command of the Army48 (see Sect. 3.5). This was based in Berlin, like the Prussian Geological Survey (transformed in 1939 into the National Geological Service49), from whom its staff members were partly derived. However, Beschoren had apparently assumed responsibility for geological work in Guernsey and Alderney by April 1942, with a base in Guernsey. From 13 April at least, the Bundesarchiv-Militärarchiv file RH32v.3041 contains Guernsey documents stamped and dated as ‘received’ at the ‘Wehrgeologengruppe’ signed by Beschoren. The same file contains copies of outgoing documents bearing his signature until 21 December, when copies of Guernsey’s water-supply map were circulated. From the large number of documents he received or generated in the months between it seems that Beschoren was continuously in post in the Channel Islands, and at work either on Guernsey or in Alderney, for at least 8 months. It is also clear that throughout this period he functioned as the senior geologist, with Dieter Hoenes as his assistant, although the ‘hard rock’ expertise of the latter might seem to be of greater use on the islands than Beschoren’s ‘soft rock’ background. But Beschoren was the older man. At 43 and so above 35, his age entitled him to TKVR status equivalent to the rank of major (as noted in Sect. 3.5). Hoenes at 30 was below the critical age, and so entitled only to the equivalent rank of cap-

 Long, winding ridges of sand and gravel seemingly deposited by streams flowing within and under glaciers. 45  Geologenstelle 2: assigned to support Fortress Engineer Staff 16 (Festungspionierstab 16). 46  Kommandeur der Landesbautrupp 1. 47  Wehrgeologenstab. 48  Oberkommando des Heeres. 49  Reichsamt für Bodenforschung. 44

6  Guernsey and the German Army

217

tain. Beschoren’s name appears first of the two when they are listed in Lautmann’s report. His signature rather than that of Hoenes appears on almost all of the documents received by the Wehrgeologengruppe for the Channel Islands. His signature authorizes almost all the reports it generated, even when these were largely or entirely the work of Hoenes as the ‘technical expert’.50

6.4.2  The Geologist Dieter Hoenes Dieter Hoenes (Fig. 6.7) was born on 8 May 1912 in Frankfurt, now the largest city in the state of Hesse in SW Germany, but grew up in Saarbrucken some 150 km further to the SW (Tröger 1956). He studied in the universities of Bonn and of Freiburgim-Breisgau, obtaining his doctorate51 on 19 June 1936.52 After three semesters as an assistant in the Mineralogical Institute at the University of Freiburg, he continued research in Heidelberg and then Berlin, publishing accounts of the petrography, tectonics and ore deposits in the Rhine/Munster valley region of Germany (Hoenes 1937), of mineral deposits (Hoenes 1939) and of magmatic activity, metamorphism, and migmatization in the SW Black Forest region (Hoenes 1940), before war interrupted his studies. The last paper treats at length with the interrelations, modes of formation, and composition of various gneissic and granitic bodies and so provided a very appropriate background for similar studies on Guernsey and Alderney. Like Beschoren, Hoenes had also begun his military service in Poland. After some 6 months of military training, he was appointed as a military geologist (Tröger 1956). Häusler (1995b, p. 22, 68) records that in November 1940 he was serving as a member of Geologenstelle 4 at Jaroslaw. However, by July 1941 he too had been re-deployed westwards, for he was posted to a military geology team (Wehrgeologenstelle 7) in France. He was to serve thereafter in France for most of the war, according to Tröger (1956), first as an assistant geologist, later as the leader of a Wehrgeologenstelle. At first responsible to the military command staff for France as a whole,53 Wehrgeologenstelle 7 was re-assigned to the Inspectorate of Fortresses (West)54 by February 1942. Häusler (1995a, p.  117) notes that Wehrgeologenstelle 7, led by Walter E. Tröger, was based in France at Nancy from early 1941 until re-deployed to the Ukraine in August 1942. In July 1941 Hoenes reported with his team-leader Tröger (a distinguished mineralogist: see Mehnert 1964) and Günter Schulz on the phosphate resources of northern France. In February 1942 Hoenes reported alone, on tunnel sites on Alderney (Häusler 1995a, p. 117).

 Sachbearbeiter.  Dr. rer. Nat. 52  Thesis title: Gesteine und Erzlagerstätten im Schwarzwälder Grundgebirge zwischen Schauinsland, Untermünstertal und Belchen. 53  Kommandostab des Militärbefehlshabers in Frankreich. 54  Inspektion der Festungen West. 50 51

218

E. P. F. Rose

Fig. 6.7  Dieter Hoenes. From Rose (2007): photograph courtesy of the late Professor Eugen Seibold, Geoarchiv, Universität Freiburg

According to Häusler (1995a), in August 1942 Hoenes was posted briefly to the paramilitary construction agency Organisation Todt, to serve with its Oberbauleitung Normandie, Abschnitt Adolf (i.e. on Alderney: see Sect. 3.7). However, reports preserved at Freiburg in the Bundesarchiv-Militärarchiv (Rose 2005a) indicate that Hoenes began geological work on Alderney in January 1942, initially as a lance-­ corporal; that he was swiftly promoted to TKVR status; and that by 1 February he was serving as a member of Wehrgeologenstelle 7. The Lautmann report described above indicates that Hoenes had been re-assigned, to Wehrgeologenstelle 4, by March 1942. This was officially deployed in support of Fortress Engineer Command XIV55 from May to the end of September according to Häusler (1995a, p.  114, 1995b, p. 22, 68), but effectively from April to 19 November (when both Fortress Engineer Command XIV and Fortress Engineer Staff 14 left the Channel Islands: Ginns 1994) according to the records cited in this chapter. By December 1942 Wehrgeologenstelle 4 had been assigned to the remaining unit, Fortress Engineer

 Festungspionier-Kommandeur XIV.

55

6  Guernsey and the German Army

219

Staff 19,56 originally co-located with Fortress Engineer Command XIV on Guernsey, and Hoenes reported then on the improvement of water supply in Alderney (Häusler 1995a, p. 113–114). By March 1943 Wehrgeologenstelle 4 had been re-assigned to Fortress Engineer Command I, and therefore to duties outside the Channel Islands. It thus appears clear that Hoenes was active in studies of Channel Island geology as a member successively of Wehrgeologenstelle 7 and Wehrgeologenstelle 4 throughout almost all of 1942, and for some 8 months (April to December) that he was contemporary there with Beschoren.

6.4.3  The Geologist Gottfried Reidl A typical Wehrgeologenstelle comprised two officers or equivalent officials; usually three non-commissioned officers (NCOs) to provide technical assistance; and four soldiers for driving, typing and assistance with exploratory boring (as noted in Sect. 3.5). From marginal data on the thematic maps compiled by Wehrgeologenstelle 4, it is evident that Beshoren and Hoenes were assisted by four NCOs, although perhaps not simultaneously: Corporals Hopp and Ludwig, and Lance-Corporals Gehrt and Dr. Reidl. Hopp, Ludwig and Gehrt seemingly had no higher academic qualification additional to their rank, and were credited only as draughtsmen for particular maps. Reidl, however, was credited with a doctorate, and on the final water-supply map compiled by the Wehrgeologenstelle (see Sect. 6.8) credited as co-compiler with Beschoren rather than as the draughtsman (his superior, Corporal Ludwig). He was thus acknowledged as a geologist in his own right. Lance-Corporal Reidl can be identified with the G. Reidel whom Häusler (1995b) lists as a military geologist serving from 1943 to 1944 still with Wehrgeologenstelle 4 but in the Pyrenean region. He can be further identified as the Dr. Gottfried Reidl (Fig.  6.8) who was employed by Germany’s National Geological Survey57 at the start of the war in its branch58 at Vienna in Austria, and who was transferred to another Austrian branch, at Linz, on 3 December 1940, before assignment to military service on 15 March 1941 (Schadler 1942). Born in Vienna on 3 June 1912,59 Reidl had obtained his doctorate by studies completed in the Palaeontological and Palaeobiological Institute of the University of Vienna. His thesis, examined on 21 May 1937,60 described a Miocene limestone sequence in part of the Vienna Basin: the sediments and their fossil fauna, with only brief reference to economic use of the rock. He worked subsequently (see

 Festungspionierstab 19.  Reichstelle für Bodenforschung. 58  Zweigstelle. 59  Archives of the Landesmuseum Linz: http://www.zobodat.at accessed 14 March 2016. 60  Thesis title: Paläobiologische Untersuchungen im Leithakalkaufschluss am “Aeuszeren Berg” bei Müllendorf im Bergenland. 56 57

220

E. P. F. Rose

Fig. 6.8  Gottfried Reidl, in 1940. From the archives of the Geological Survey (Geologische Bundesanstalt) of Austria, Vienna, per Thomas Hoffman and Hermann Häusler; reproduced by kind permission

Reidl 1939, 1940) as a palaeontologist61 at the Geological Survey of Austria,62 which was soon to be incorporated as a branch within Germany’s National Geological Survey. He worked principally in the Survey’s Museum, his most notable publication describing a new fossil species of spatangoid echinoid (Reidl 1941) distinguished earlier in his thesis. The Survey’s records (Hermann Häusler, pers. com. 2016) credit him with assisting military geological mapping: in 1938 and 1939 for the 1:75,000-scale map sheets of Znaim, Nikolsburg, Hollabrunn and Mistelbach (in the fortified border area of Lower Austria with Czechoslovakia), and mapping in 1939 for the 1:75,000 sheets of Eisenstadt, Neusiedl am See and Pamhagen (Burgenland). He married on 12 October 1939:63 a son was born on 24 August 1943.64 At some time he became a member of the National Socialist party: Survey records indicate that he was removed from office at the end of the war because of this (in ignorance of his earlier death, as a prisoner-of-war, on 24 April 1945).

 Wissenschaftlicher Beamter.  Geologischen Bundesanstalt. 63  Johanna Franziska Vitovsky. 64  Günther Gottfried Franz. 61 62

6  Guernsey and the German Army

221

Reidl thus arrived on the Channel Islands in 1942 at about 30 years of age; with some three years of experience as a ‘Survey’ geologist, albeit in palaeontology and sedimentary petrography rather than ‘applied’ geology; and with about a year’s experience of life in the armed forces.

6.5  German Fortifications and Military Geological Mapping Guernsey was to be fortified in much the same way as Jersey: numerous coastal artillery batteries to command the adjacent seaways (Fig. 4.2), anti-aircraft batteries to defend from aerial attack, infantry strong points and resistance nests to defend the coast from amphibious assault, all commanded from headquarters constructed to be safe from bombardment by sea or by air (Fig. 6.9) (Gavey 2001). Guernsey’s airfield and anti-aircraft batteries were manned by the Luftwaffe, and planning for construction work was therefore assisted where necessary by Luftwaffe geologists, as described in Chap. 7. Wehrgeologenstelle 4 focused its attention on sites operated by the Navy and the Army. Beschoren and Hoenes generated four of their ten ‘expert opinions’ (Gutachten numbered 2, 3, 7 and 8) and three unnumbered ‘memoranda’ for Guernsey (Table 6.1, items 7 to 13), together with at least seven geotechnical maps (Table 6.2, items 3 to 9). Widest circulation was given to the Gutachten: typically two copies to the geologist at the Inspectorate of Land Fortification of the Commander-inChief West, and one each to Fortress Engineer Command XIV,65 Fortress Engineer Staff 1966 and Wehrgeologenstelle 4 itself. Gutachten 2 and 3 were also sent to the Air Force Works Office on the Channel Islands.67 Overall, the distribution indicated that no more than six copies in total were generated for any of these documents.

6.5.1  Artillery Batteries As shown on Fig. 6.9, 17 batteries of coastal artillery were sited on Guernsey: three to be operated by the Navy and 14 by the Army. Their total was thus significantly more than the nine batteries assigned to the larger island of Jersey, closer to the protection afforded from the French coast.

 Fest. Pi. Kdr. XIV.  Fest. Pi. Stab 19. 67  Luftwaffenfeldbauamt Kanalinseln 65 66

222

E. P. F. Rose

Fig. 6.9  Map of principal German fortifications on Guernsey as at June 1944. For coastal batteries, data are provided for number, calibre, type and range of guns in place. After Rose (2005b) and a larger-size figure within a poster illustrating Atlantic Wall defences 1940–1945: Guernsey, published in 1992 by Colin Partridge through Ampersand Press, Alderney. From Robins et al. (2012) courtesy of the authors and the Geological Society of London

The Navy’s Battery ‘Mirus’ (Figs. 6.5 and 6.9) was the most formidable not only on Guernsey but on the Channel Islands as a whole. Problems of water supply led to geological reports by Friedrich Röhrer and by TKVR Scherer (as described in Sects. 6.3.2 and 6.3.3 above). They led also to the first ‘expert opinion’ for Guernsey from Wehrgeologenstelle 4 (Gutachten 2) (Table 6.1, item 7).68 Dated 17 June 1942 and signed by Beschoren, this two-page document recommended emplacement to a depth of 3.5 m of a suitably positioned infiltration gallery to collect groundwater— the site indicated on an accompanying diagram—to develop Scherer’s proposal. Fire from the batteries was also to be controlled, as on Jersey, from a system of naval direction and range-finding towers planned to ring the island. Eight were planned: on the Chouet headland, and at Fort Saumarez, Pleinmont, L’Angle, La Prevote, Vale Mill, Icart Point, and Jerbourg Point (Forty 1999). Of these, five were completed: Chouet (although later demolished due to subsidence), Fort Saumarez, Pleinmont, L’Angle (Fig.  6.10), and Vale Mill (the last only as an Army coastal observation post, and now partly destroyed). This corresponds with nine planned for the much larger island of Jersey, of which only three were actually completed.

68

 Gutachten über die zusätzliche Wasserversorgung der Batterie Mirus.

6  Guernsey and the German Army

223

Fig. 6.10  Coastal artillery fire control tower MP4 at L’Angle, near Batterie Dollman (site 4 on Fig. 6.9) on Guernsey’s SW headland, perched high on the rocky cliffs that overlook the southern coast. Photo: E.P.F. Rose

6.5.2  Beach Defences The principal weapons associated with beach defence were, as on Jersey, 105 mm calibre guns, originally designed as World War I field guns and manufactured by Schneider of France. These had been captured by German troops by victory in the Battle of France in 1940, removed from their wheeled carriages, and mounted on turntables. Twenty-one of these K331(f) guns (the ‘f’ indicating their French origin) were emplaced in concrete casemates, and another 13 mounted in open field positions (Gavey 2001). Moreover, again as in Jersey, defence against armoured assault was provided by 47  mm calibre Skoda anti-tank guns of Czechoslovakian manufacture, seized on Czechoslovakia’s prewar annexation by Germany. One of the few strictly ‘fortress’ weapons used on the Channel Islands by the Germans, these Pak 36(t) guns (the ‘t’ indicating their ‘Tchechoslovak’ origin), each with a coaxial machine gun, were mounted in a ball mount in a square steel embrasure. Sixteen were mounted in casemates, and 12 emplaced in ‘field positions’. And yet again as in Jersey, MG34 machine guns were mounted in bunkers with thick armoured turrets, and French tank turrets were mounted in ‘Tobruk’ pits in large numbers, housing both machine guns and 37 mm anti-tank guns. Atypically, four bunkers were built for an advanced automatic-firing ‘fortress’ mortar, the 50 mm M19, which could fire up to 120 rounds a minute (Gavey 2001).

224

E. P. F. Rose

Fig. 6.11  Map of Vazon Bay area on the west coast of Guernsey (SW of 1  km scale bar on Fig. 6.9), with key (top left) showing (from the top down): lines A–B and C–D for excavation of an anti-tank ditch; low tide mark; high tide mark; sand; rock; made ground and a stream course. The straight lines A–B and C–D have been drawn on the printed base map slightly inland from the houses close to the high tide mark. From Gutachten 7 by Bernhard Beschoren, dated 22 August 1942; reproduced by permission of the Bundesarchiv-Militärarchiv from file RH32/3029

Apart from Gutachten 2, only one other geological ‘expert opinion’ for Guernsey dealt with aspects of fortification as such: Gutachten 7 (Table 6.1, item 9),69 on creating an anti-tank terrain inland of Vazon Bay, on the north-west coast of Guernsey (Fig. 6.11). It was thought that this bay, fringed by a beach partly of sand, partly rock, was the most likely to attract amphibious assault with armoured support, and that a barrier should be constructed to impede the movement of tanks inland from the potential landing zone. Boreholes were put down by auger to depths mostly between 1 and 2.5 m at 38 sites inland of the beach, to determine the nature of the ground and the depth to groundwater. In consequence, it was proposed to excavate an anti-tank ditch to a depth of 2 m along the lines A–B and C–D shown on Fig. 6.11, A–B being excavated by hand because of the low bearing strength of the peaty ground in this region and, with water table at 1–2 m depth, the need to protect from water ingress during construction. The length C–D, with firmer ground and deeper water table, could be created using mechanical excavators. The document comprises only 2.5 pages of text, dated 22 August 1942 and signed by Beschoren, but has two attached diagrams, one to show the location of borehole sites, the other

69

 Gutachten über die Schaffung eines panzerischeren Geländes landeinwarts der Vazon-Bay.

6  Guernsey and the German Army

225

Fig. 6.12  Vazon Bay, viewed northwards at low tide towards the headland whose summit is capped by Fort Hommet. The northern and western coasts of Guernsey lack the high cliffs that fringe much of the island’s southern coast, so are potentially more suitable for amphibious assault. Photo: E.P.F. Rose

(Fig. 6.11), drawn by Lance-Corporal Gehrt, to show the position of the proposed anti-tank ditch. Vazon Bay (Fig. 6.12), lacking in coastal cliffs, had been deemed vulnerable to amphibious assault even during the nineteenth century. The British had constructed Fort Hommet on the rocky promontory to the NE, Fort Le Croq on that to the SW, in order to subject a beach landing to cross (enfilade) fire. A Martello Tower guarded the sandy bay itself. For similar purposes, the Germans transformed the Fort Hommet area into a strongpoint70 by addition of bunkers and other fortifications that included two casemates with K331(f) guns (e.g. Fig.  6.13) plus one with a Pak K36(t). Fort Le Croq was transformed into a resistance nest71 that included one casemated K331(f), and three resistance nests were evenly spaced along the road that fringed the bay. The style of fortification was thus similar to that illustrated for Jersey in Sect. 4.4 (cf. Fig. 6.13 with Fig. 4.10).

6.5.3  Geological Maps To guide fortification in general, at some time during 1942 Wehrgeologenstelle 4 compiled a military geology map of Guernsey. Moreover, since published geological maps of the island existed only as small-scale figures (as described in Sect. 2.4), an early priority was to generate a geological map for Guernsey at an appropriately larger scale. The only known copies of this are two geological maps (plus the mili-

70 71

 Stützpunkt Rotenstein.  Wiederstandsnest Krossen.

226

E. P. F. Rose

Fig. 6.13  Casemate and renovated 105  mm calibre German gun at Fort Hommet. Image from Wikimedia Commons: file 10.5 cm Gun Casement.JPG, reproduced by licence CC BY-SA 3.0

tary geology map) preserved in the National Archives of the USA.72 Neither map bears a title, date, or indication either of the compiler or the sponsoring unit. Thus neither appears to be a finished document. They were presumably produced as a first stage in the preparation of more specialist geotechnical maps. Both have a German lithostratigraphical key, so were clearly prepared under German auspices, and since they are filed in the Archives with other maps generated by Wehrgeologenstelle 4, it seems likely that they were generated by this unit and therefore during 1942. One version of the map is of bedrock geology only (Table  6.2, item 3; and Fig. 6.14), comparable with the contemporary ‘solid’ geological maps then being published for the United Kingdom by the Geological Survey of Great Britain. In addition to dune sands and gravel beaches (which obscure the bedrock along parts of the coastline: coloured yellow), the map illustrates the island’s geology in terms of eight rock types. The key is neatly drawn, and this geological map is closer to being a ‘fair copy’ than the other version. The map confirms the outcrop pattern previously mapped (see Sect. 2.4) for granites bordering the northwestern coast and hornblende gabbro

72

 The National Archives and Records Administration, at College Park, Maryland.

6  Guernsey and the German Army

227

Fig. 6.14  Map of the bedrock (solid) geology of Guernsey, original at scale of 1:25,000 and on a ‘second edition’ topographical base map, apparently prepared by Wehrgeologenstelle 4 in 1942. Key in translation (from the top down): (1) dune sand, (2) beach sand, (3) quartzite, (4) hornblendite, (5) granite, (6) diorite plus syenite, (7) hornblende gabbro, etc., (8) augengneiss, (9) ‘injection’ gneiss and (10) gneiss. From Rose (2005b); reproduced courtesy of the US National Archives and Records Administration

along the northeastern coast, but presents an interpretation differing in detail for other features of the bedrock. The other version of the map (Table 6.2, item 4; and Fig. 6.15) plots the outline of a superficial cover within the island’s interior upon a partially completed map of the bedrock geology, so is more comparable with the Geological Survey’s former ‘drift’ maps. The area outlined approximates to the outcrop of loess as mapped postwar by British geologists (see Sect. 2.4, and references in Rose 2005b). The map was clearly a working document. The bedrock geology is not only less ­complete but slightly different from that plotted on the other map. The key does not distinguish coastal gravels from dune sands, excludes quartzite and transposes gneiss to before augengneiss in order of listing. Moreover, there are numerous annotations handwritten in German in the margins of the map sheet.

228

E. P. F. Rose

Fig. 6.15  Map of the bedrock plus superficial (solid and drift) geology of Guernsey, original at scale of 1:25,000, apparently prepared by Wehrgeologenstelle 4 in 1942. Key in translation (from the top down): (1) sand and loam, (2) hornblendite, (3) granite, (4) diorite plus syenite, (5) hornblende gabbro, (6) gneiss, (7) augengneiss, (8) ‘injection’ gneiss and (9) a blank box, perhaps intended for the ‘quartzite’ shown in the key to Fig. 6.14. From Rose (2005b): reproduced courtesy of the US National Archives and Records Administration

6.5.4  Military Geological Map of Guernsey The US National Archives also contain two identical copies of a ‘military geology map’ of Guernsey73 (Table 6.2, item 5; and Fig. 6.16). Although hand-coloured, the title and a key to symbols and colours used have been inked upon the base map by a cartographic draughtsman, presumably for multiple-copy reproduction. Responsibility for compilation74 is given as ‘Dr Beschoren’ and for draughtsmanship75 as ‘Lance-Corporal Gehrt’:76 the same draughtsman as for the map shown as Fig. 6.23.

 Wehrgeologische Karte von Guernsey.  Entwurf. 75  Gezeichnet. 76  Gehrt, Gefr. 73 74

6  Guernsey and the German Army

229

Fig. 6.16  Wehrgeologische Karte. Military geology map of Guernsey, original at scale of 1:25,000, on ‘second edition’ of German base map, prepared by Wehrgeologenstelle 4. The key (from the top down) distinguishes ten coastal features, associated with a tidal range of 8.5 m, and ten inland features. In translation, the coastal features comprise: (1) low tide level; (2) high tide level; (3) coastal cliffs; (4) valleys incised through the coastal cliffs; (5) sandy beaches; (6) areas with sand dunes; (7) positions of shoreline gravel banks; (8) crags and platforms of hard (igneous) rock; (9) crags and platforms of gneiss (metamorphic rock) and (10) areas of raised shorelines. Inland features comprise: (1) boundary of lowland areas; (2) steep cliffs; (3) blown sand; (4) alluvium; (5) peat; (6) loamy clay; (7) deeply-weathered bedrock; (8) loess; (9) gneiss and (10) igneous rock. From Rose (2005b); reproduced courtesy of the US National Archives and Records Administration

The map distinguishes features of coastal geology and geomorphology from those of inland areas. The ten coastal features were all those likely to influence choice by the Allies of particular areas as suitable for amphibious assault, and so determine the siting of appropriate countermeasures by German forces. The inland features distinguished are essentially those of geomorphology (the boundary of lowland areas contrasted with the steeper cliffs which border the high plateaux regions); superficial deposits (blown sand, alluvium, peat, loamy clay, deeply-weathered bedrock and loess) and bedrock (gneiss or igneous rock). The geomorphology was potentially important for siting lines of defence, cliffs being an obvious obstacle to vehicular movement and forming natural defensive features. That six categories of inland superficial deposits are differentiated is a

230

E. P. F. Rose

consequence of the importance of these near-surface materials for determining the diggability of the ground for infantry trenches, and its suitability for the siting of stores and vehicle parks, gun positions and other military engineering works. In contrast, the inland bedrocks are grouped into only two broad categories: gneiss and igneous rock. For military purposes, the type of igneous rock (acid, intermediate or basic) and time of relative intrusion are normally irrelevant, as are the different types of gneiss. Rocks of similar strength and so potential engineering use are normally grouped together on German military geology maps, as here. This is more truly a ‘military geology’ map than the so-called ‘military geology’ map compiled by Walter Wetzel (Fig. 6.2), which rather combines one type of ‘engineering geology’ map with a ‘resources’ map. Moreover, aspects of its basic geology can be matched more closely with the German maps illustrated here as Figs. 6.14 and 6.15 than map figures published earlier (those illustrated and described in Sect. 2.4). This map appears to be based on ground observations rather than desk study of published maps. It represents a type of mapping new to the British Isles.

6.6  Tunnels and Rolf Thienhaus Underground facilities77 were to be excavated on Guernsey as on Jersey. Ginns (1993) and Gavey and Powell (2012) record that building progress reports survive for May 1942, July and August 1943, and March 1944. In May 1942, 16 tunnels were planned, numbered Ho 1 to Ho 16. By July 1943 the numbering system had extended to Ho 41, but Ho 9, 13, 16, 17, 24 to 28, 38, and 39 were omitted from the accompanying plan of locations (Fig. 6.17). By March 1944, only 15 tunnels were actually under construction. Only Ho 8 and Ho 31 were eventually completed, whilst Ho 3, 4, 7, 12 and 40 were in a sufficiently advanced state, with sections concreted, to be put to use. However, the use planned for individual tunnels changed as the construction programme evolved.

6.6.1  Advice by Bernhard Beschoren Gutachten 3 (Table 6.1, item 8),78 signed by Beschoren and dated 23 June 1942, concerned a roof collapse during construction of ‘ration storage tunnel 1 (the gasworks tunnel)’. The report does not explicitly state whether this tunnel is on Guernsey or Alderney, but since the rocks involved are diorite and hornblende gabbro, the tunnel must be on Guernsey. The tunnel location map (Fig. 6.17), the geological map (Fig. 6.14) and the site description provided by Ginns (1993) and by

77 78

 Hohlgangsanlagen.  Gutachten über den Denkeneinbruch in Verpflegungsstollen 1 (Gaswerkstollen).

6  Guernsey and the German Army

231

Fig. 6.17  Map of Guernsey, showing positions of the German underground facilities planned or under construction as at 29 June 1943. From Rose and Willig (2013), after Ginns (1993) with enlargement of names and numbering, by kind permission of the Channel Islands Occupation Society (Jersey)

Gavey and Powell (2012) are consistent with identification as Ho 2. Ho 2, inland from Guernsey’s main town, St. Peter Port, was indeed listed as ‘Ration Store 1’ on German documents of 1942. The roof had collapsed over a distance of some 15 m at the end of May, at a point about 120 m south from the northern tunnel entrance (Fig. 6.18). Eleven borings had then been put down to determine rock conditions, and established that an oval-shaped weathering zone some 30 m long and 15 m wide extended down more than 12 m from the surface. The ‘opinion’ recommended that a cement grout be injected to stabilize the fractured rock in the tunnel roof, and that a vehicle exit branch to the tunnel be routed through the region identified as of fresh, unweathered and unfractured diorite. The requirement could probably be met by shifting the exit branch some 25–30 m to the south. It was important also that the ventilation shaft and the emergency exit should lie outside the weathered zone, in fresh rock. Consolidation of the friable parts within the zone of weathered rock by means of cement injection together with an increase in tunnel size were considered to be adequate to provide safety in case of fire.

232

E. P. F. Rose

Fig. 6.18  Diagram showing axis of Ho 2 (the ‘Gasworks’ tunnel), adjacent quarry (Steinbruch) faces and sites of 11 exploratory boreholes in the region of roof collapse. From Gutachten 3 by Bernhard Beschoren, dated 23 June 1942, with lettering enhanced by T. Davenport; reproduced by permission of the Bundesarchiv-Militärarchiv from file RH32/3029

Work on this complex had started early in 1942 and continued until the autumn of 1943, by which time much of it was concrete lined (Fig.  6.19). When work ceased, 11,783 m3 of rock had been extracted and 2841 m3 of concrete had been poured. After the war, the tunnel was prone to further roof collapse in unlined portions, and has normally been sealed to public access.

6  Guernsey and the German Army

233

Fig. 6.19  Plan of Ho 2 as at the end of hostilities, in May 1945. Reproduced from Gavey and Powell (2012), by kind permission of Steve Powell and Festung Guernsey

6.6.2  Advice by Rolf Thienhaus It seems likely that the geologists of Wehrgeologenstelle 4 visited other tunnel sites to provide advice, much as Klüpfel had done on Jersey. If so, such visits did not generate reports that are known to have survived the war. However, on Guernsey, the tunnellers had the benefit of advice from another geologist: the ‘Dr. Dienhaus’, supposedly from a geological team,79 who Klüpfel records (Table 4.1) as visiting him on Jersey on 14 May 1942, but who Rose and Willig (2009) have now identified as Rolf Thienhaus (Fig. 6.20). Rose (2005a) recorded the existence of a report by Second Engineer Mining Company,80 and later noted (Rose 2007, p. 12) that ‘the officer commanding the 2nd Mining Engineer Company, at work in the valley west of Fort George, commented on the nature of the geology’ both at the ground surface and underground ‘with reference to a geological map to hand provided by “Sonderführer Dr. Graupner”— and the expertise of Lance-Corporal81 Dr. Thienhaus of the First Mining Engineer Company.’

 ‘Wehrgeologenstelle’, although whether of the Army or the Luftwaffe is not specified.  [Hauptmann u. Komp. Chef] Die geologischen Verhältnisse im Einsatzgebiet der Pi. Minier Komp. auf der Insel Guernsey: Minierte Anlagen westl, Fort George – Tal “la Colombelle”. 2. Pi. Minier Komp., 15 Jun 1942, 2 pp. (In Bundesarchiv-Militärchiv file RH32v.3041.) 81  Gefreiter. 79 80

234

E. P. F. Rose

Fig. 6.20  Dr. Rolf Thienhaus (1913–1968), who served on Guernsey as a geologist Gefreiter (lance-corporal) with an engineer mining company in 1942. Photo taken in 1967. From Rose and Willig (2009), reproduced by permission of the Universitätsbibliothek Freiburg

‘Dr. Thienhaus’ is the correct spelling for the ‘Dr. Dienhaus’ recorded by Klüpfel. Presumably he is also the Rolf Thienhaus whom Häusler (1995b, p. 50, 85) says was in post as a military geologist in April/May 1942 with the Engineer branch of

6  Guernsey and the German Army

235

Seventh Army Headquarters,82 for Häusler records nothing more of any German military geologist of this surname during World War II. Häusler (1995b, p. 14, 63) does record that a ‘Dienhaus’ was an Air Force geologist83 in October 1942 but again nothing more. That Thienhaus [sic] did indeed provide geological information to support the work of the Second Engineer Mining Company on Guernsey is, however, substantiated by a report extract.84 This copy available is stamped ‘secret’,85 and as received by the military geologist at the Inspectorate of Land Fortification (West).86 It provides a brief summary of geological knowledge for Guernsey and Alderney prior to description of some features of tunnel boring on the two islands. A footnote to the section on the geology of Guernsey refers to an oral communication from ‘Dr. Thienhaus’ concerning the identification of a metamorphic rock—so evidently he was contributing to tunnelling projects on Guernsey in 1942, and regarded as a geologist whose identification of rock types could be treated as authoritative. It seems clear that ‘Dr. Thienhaus’ must have been Rolf Thienhaus (1913–1968), who postwar became a Professor of Geology at the Technical University of Clausthal (now Clausthal-Zellerfeld), near the edge of the Harz Mountains in central Germany. According to obituaries (Grabert 1969; Pilger 1969), Thienhaus was born on 26 October 1913 in Cologne,87 where he went to school and in 1932 began university studies—in geology and mineralogy, under Professor Philipp (coincidentally one of the pioneer German military geologists of World War I: see Sect. 4.2.2). In the summer of 1935 he transferred to the University of Göttingen, where he completed doctoral studies (Thienhaus 1940a) on rocks of Devonian age within part of southern Westphalia, under the supervision of professors Walter Schriel and Hermann Schmidt. Apart from his thesis, he published only one other geological article during the war: on ore deposits of the mineral barytes in the Richelsdorf Mountains of Hesse in central Germany (Thienhaus 1940b). Thienhaus was appointed an assistant (lecturer) in the Geological-Palaeontological Institute of the University of Göttingen with effect from 11 March 1940, but in February 1940 he began war service. Professor Pilger notes that this service ­continued to the end of the war, but says very little about it: only that Thienhaus served in the engineers, began his service in Höxter (a town in eastern North Rhine-­ Westphalia) where his former professor Walter Schriel was now a captain and company commander, and that he served in the engineers finally in the rank of second

 Armeepionierführer des AOK 7.  Luftwaffengeologe. 84  [Anon., report presumably written mid October 1942, extract without address or addressee] Auszug aus dem Bericht der 2. Pionier- Minier- Komp. über den Einsatz im Hohlgangsbau auf den Kanalinseln GUERNSEY- ALDERNEY, v. 16.3.-8.10.1942. 7  pp. (Filed in the archives of the Bundeswehr Geoinfomation Centre, Euskirchen.) 85  Geheim. 86  Insp. d. L. West. Wehrgeologe, for filing as ‘Anlage zu Az 39, Geol 10, Nr 56/43 geh’. 87  Köln. 82 83

236

E. P. F. Rose

lieutenant.88 Whilst on the Channel Islands, Thienhaus was thus a relatively young and newly qualified geologist, able to put his expertise to some relevant use despite his low military rank and engineering rather than geological appointment.

6.7  Quarrying for Raw Materials Gutachten 8 (Table 6.1, item 12),89 stamped ‘secret’,90 provided a two-part explanation to accompany a construction materials map (Table 6.2, item 6; and Fig. 6.21). Part one (16 pages) dealt with the ‘hard’ stone bedrock, and was written by Hoenes as ‘technical expert’.91 It summarized the properties of diorite, granite (distinguishing between granites of the Cobo Bay, Grand Havre and Lancresse Bay regions) and gneiss. Also, it listed 75 quarries with a brief description of each locality and its stone product. Part two (only two pages), by Beschoren, dealt with the superficial deposits: sand, gravel and loam. It ended with brief description of nine localities (all essentially coastal) from which beach sands and gravel could be obtained, and one locality (near St. Peter Port) for loam. According to its key (Table 6.5), the map itself classifies these sites potentially yielding building materials into ten categories. Evidently, qualities deemed militarily important are the nature of the product (stone, gravel, aggregate, sand or brickearth); for stone, its readiness of access (i.e. whether the quarry was working or disused, dry or flooded); and for gravel or aggregate, whether or not mechanical plant had been installed. To enable the vast quantities of building materials required to be transported from beaches and quarries, as well as imported stores and ammunition, a railway was built from the harbour at St. Peter Port north along the coast to St. Sampson, and thence inland to the NW and around the west coast of the island (Wilson 1970; Gavey 2001). Construction of the numerous, successive coastal defence bunkers was initiated as the head of the railway progressed along the west coast, to Vazon Bay and finally L’Erée. However, by the autumn of 1943 the principal fortifications had been completed, and most of the railway track was subsequently lifted. Only the line between St. Peter Port and St. Sampson remained in operation until the end of the war. Notably, construction of the original 22-km railway line itself demanded considerable quantities of aggregate: the crushed stone foundation on which the 0.9 m track was laid could rise to some 0.6 m above the surrounding terrain.

 Leutnant.  Erläuterungen zur Baustoffkarte der Kanalinsel Guernsey. 90  Geheim. 91  Sachbearbeiter. 88 89

6  Guernsey and the German Army

237

Fig. 6.21  Baustoffkarte. Map of Guernsey, original at scale of 1:25,000, issued by Wehrgeologenstelle 4 in September 1942, to show potential quarry sites for building construction materials. For key, see Table  6.5. From Rose (2005b); reproduced by permission of the Bundesarchiv-Militärarchiv from file RH32/3027

Table 6.5  Translation of key to the map shown on Fig. 6.21, top to bottom Large red circle Medium red circle Small red circle Blue W Yellow spot Yellow spot with black circles Orange spot Orange spot with black circles Brown spot Green spot

Quarry in operation. Mechanical plant present Excavation dry (red centre) or filled with water (blue centre) Disused quarry, worth bringing into operation Excavation dry (red centre) or filled with water (blue centre) Disused quarry, not worth bringing into operation. Excavation dry (red centre) or filled with water (blue centre) Quarry used as a reservoir for the waterworks Gravel pit Gravel pit with mechanical equipment Pit yielding aggregate for concrete Pit yielding aggregate for concrete, mechanically equipped Building sand Brickearth

238

E. P. F. Rose

6.8  Water Supply The first ‘opinion’ for Guernsey (Gutachten 2), by Beschoren, dealt with a supply of drinking water for the garrison of the island’s major battery, the formidable ‘Mirus’ (see Sect. 6.3.2 above). In September Beschoren sent a memorandum (Table 6.1, item 10)92 on waterworks planned for the Air Force to Fortress Engineer Command XIV,93 and later that month sent copies of two groundwater maps94 at 1:25,000 to the military geologist at the Inspectorate of Land Fortification West95 with a brief covering memorandum (Table 6.1, item 11).96 The memorandum concludes with reference to the two other large Channel Islands, which may be translated as: ‘The water supply of Alderney comes from collection galleries and improvement of springs. Lieutenant Prof. Klüpfel was asked on 18.9.42 for reference material for Jersey.’ A single original manuscript for each of the two maps is preserved in the US National Archives. Data shown on these two draft maps, slightly simplified, were apparently copied on to second edition topographical base maps as hand-coloured annotation. Single copies of the finished maps were sent to the military geologist at the Inspectorate of Land Fortification West with a covering letter, but no accompanying report, on 23 September 1942. These final copies of both maps and the letter are preserved, with a third map, in Germany in the Bundesarchiv-Militärarchiv, in file RH32v.3041. A map in the US National Archives entitled ‘well depth within the region of projected coastal fortifications’97 was seemingly the draft for a ‘groundwater map’98 (Table 6.2, item 7; and Fig. 6.22) now in the Bundesarchiv-Militärarchiv. On the draft, sites of wells in four depth categories are plotted by colour-infilled circle upon the 1:25,000-scale topographical base map: (1) wells to depths of 2 m (green); (2) to 7 m (blue); (3) to 15 m (orange); and (4) over 15 m (purple). The highest concentration of wells is plotted along the western rather than the eastern coastal regions. A caption indicates preparation by Wehrgeologenstelle 4, and is signed by ‘Dr Beschoren’ as compiler, ‘Hopp, Uffz.’ (Corporal Hopp) as draughtsman. The ‘groundwater map’ itself (Fig. 6.22) bears a cartographically drawn title and key to its colours and symbols, but no indication of authorship. It adopts different categories for well sites, combining the small number of wells with potable groundwater over 7 m deep into a single category and adding locations of springs, saline wells and dry wells, thus amplifying data provided by the draft map.

 Geplante Wasserwerke der Luftwaffe.  Fest. Pi. Kdr. XIV, copied only to Insp. d. Landesbefest. West and Wehrgeol. Stelle 4. 94  Grundwasserkarte and Karte der Brunnentiefen. 95  Inspekteur der Landesbefestigung West. 96  Wasserversorgung im Gebiet des verstärkten Küstenausbaus. 97  Brunnentiefen innerhalb des Gebietes des verstärkten Küstenausbaues. 98  Grundwasser Karte. 92 93

6  Guernsey and the German Army

239

Fig. 6.22  Grundwasser Karte. Map of Guernsey, original at scale of 1:25,000, on ‘second edition’ of German base map, prepared by Wehrgeologenstelle 4 in September 1942, to show well depths and salinity in the region of projected coastal fortifications. Key (translated, from the top down) showing wells with depth to groundwater: (1) 0–2 m (green spot); (2) 2–7 m (blue); (3) >7 m deep (red); plus (4) springs (dark blue); (5) saline wells (larger red-infilled circles) and (6) dry wells (red unfilled circles). From Rose (2005b); reproduced by permission of the Bundesarchiv-Militärarchiv from file RH32/3041

A tracing overlay entitled ‘groundwater map of the region of projected coastal fortifications’99 (Table 6.2, item 8; and Fig. 6.23) attached to the draft well-depth map in the US National Archives was seemingly the precursor of a ‘groundwater map of the fortified coastal zone’100 (Fig. 6.24) now in the ­Bundesarchiv-­Militärarchiv. Both the tracing overlay and the final map bear the source designation Wehrgeologenstelle 4, and show the name of the compiler as ‘Dr Beschoren’. The overlay, which additionally credits ‘Gehrt, Gefr.’ (Lance-Corporal Gehrt) as the draughtsman, has a cartographically drawn key to seven features. It shows very clearly that regions for potential groundwater extraction lie adjacent to the north and western coasts rather than to the SE of the island. The map actually circulated, in September 1942 (Fig. 6.24), shows the compiler as ‘Dr Beschoren’ with TKVR sta-

99

 Grundwasserkarte im Gebiet des verstärkten Küstenausbaues.  Grundwasserkarte der befestigten Küstenzone.

100

240

E. P. F. Rose

Fig. 6.23  Grundwasserkarte im Gebiet des verstärkten Kustenausbaues. Tracing overlay, original at scale of 1:25,000, prepared by Wehrgeologenstelle 4. Key to symbols and colours (translated, from the top down): (1) areas of saline groundwater; (2) water courses; (3) freshwater ponds; (4) springs; and areas of potable groundwater at depths below the ground surface of (5) 0–2 m, (6) 2–7 m and (7) >7 m. From Rose (2005b); reproduced courtesy of the US National Archives and Records Administration

tus. It omits plotting of freshwater ponds, and the key thus shows six rather than seven features. The definitive report on Guernsey’s water supply (Table  6.1, item 13)101 was issued in December to accompany a third type of map: potential water supplies.102 The report comprised a catalogue of 279 operational wells and springs with ­summary details of well type, depth to water table, well depth and remarks (e.g. as to whether water was raised by a windpump or motorpump), plus a list of 20 civilian waterwork installations (three wells, three waterworks, 12 reservoirs, two pumping stations) with potential yield or capacity and an indication of water quality: potable or merely utility water.103

 Wasserversorgungsmöglichkeiten der Kanalinseln Guernsey.  Wasserversorgungsmöglichkeiten. 103  Brauchwasser. 101 102

6  Guernsey and the German Army

241

Fig. 6.24  Grundwasserkarte der befestigen Küstenzone. Map of Guernsey, original at scale of 1:25,000, prepared by Wehrgeologenstelle 4 in September 1942, to show groundwater conditions within the region of projected coastal fortifications. Key (translated, from the top down): 1) groundwater at depths of 0–2 m (green diagonal shading); 2) 2–7 m (blue); 3) >7 m (orange); 4) saline groundwater (brown); 5) springs (blue spot) and 6) water courses (blue line). From Rose (2005b); reproduced by permission of the Bundesarchiv-Militärarchiv from file RH32/3041

The map of potential water supply104 (Table 6.2, item 9; and Fig. 6.25), preserved in the Bundesarchiv-Militärarchiv, has no counterpart in the US National Archives. This map was apparently produced later than the other two, and not circulated until December 1942 (Rose 2005b). Preparation is shown by two technical specialists105 TKVR Dr. Beschoren and Lance-Corporal106 Dr. Reidl, drawing by Corporal107 Ludwig, of Wehrgeologenstelle 4. The cartographically drawn key (Table 6.6) distinguishes 12 features. Preparation of these water-supply maps and reports in 1942 was, as in Jersey, to anticipate compilation of British hydrogeological data by over 50 years (see Chap.

 Karte der Wasserversorgungs-möglichkeiten.  Sachbearbeiter. 106  Gefreiter. 107  Uffz. 104 105

242

E. P. F. Rose

Fig. 6.25  Karte der Wasserversorgungsmöglichkeiten. Map of Guernsey, original at scale of 1:25,000, prepared by Wehrgeologenstelle 4 in September 1942, to show potential water supply. For key, see Table  6.6. From Rose (2005b); reproduced by permission of the Bundesarchiv-­ Militärarchiv from file RH32/3041

9). The current population on Guernsey is about 60,000. In 1939 it was about 42,000, but some 17,000 people were evacuated immediately prior to the German occupation, and about 1000 deported during it. Offsetting this reduction in numbers, the German forces had built up to a strength of about 13,000 by 1942/1943, supplemented at this time by a construction workforce of up to 7000 men (Robins et al. 2012). German demand for water was a consequence of this increased population; military demand for increased agricultural productivity; extensive concrete ­production for the massive fortifications and requirement to supply water to the newly garrisoned strongpoints. All this occurred at a time of reduced annual rainfall. The German groundwater maps almost certainly represent one of the most detailed pieces of hydrogeological mapping of this type conducted in the British Isles or indeed elsewhere by 1942 at a scale of 1:25,000. The detailed depth to water survey carried out along the coastal margins of Guernsey (Figs. 6.22 and 6.24) is unsurpassed by later British work on the island. However, postwar investigation of Guernsey (Robins et al. 2002) has concentrated, as with Jersey, on three-­dimensional portrayal of the groundwater flow regime and construction of a conceptual ground-

6  Guernsey and the German Army

243

Table 6.6  Translation of key to the map shown on Fig. 6.25, top to bottom Dashed blue line Red spot Blue spot Pale blue spot Red spot with blue outline Thin red ‘v’ Double red ‘v’ Large blue spot Red horizontal lines in blue box Large green spot Red horizontal lines in green box Area thickly outlined in green

Watercourses Springs of potential use Well shafts with hand pump or electrically driven pump Sites suitable for Abyssinian wells, to 8 m depth Sites suitable for well shafts fitted with ‘Kolben’ pumps, to 8 m or greater depth Site where installation of an infiltration gallery is possible Site where an infiltration gallery is already present Reservoir for drinking water Pumping station for drinking water Reservoirs for utility water Pumping stations for utility water The airfield and catchment area for planned Air Force water works

water flow scheme for the island, making use of groundwater chemistry as an indicator of provenance and aid to delineate areas of brackish groundwater—and with focus primarily on data from relatively few recently drilled deep boreholes rather than the many shallow wells in use during the German occupation.

6.9  The Island of Sark A geological map of the island of Sark, to the east of Guernsey, was compiled during the German occupation (Fig. 6.26). However, this small island merited only a small garrison. A few tunnels were excavated to provide shelter, but fortification as such was relatively light. None of the tunnels was given a number; none was completed; no concrete linings were installed; all are small excavations; and no German map showing their location is known to have survived the war. Gavey and Powell (2012) describe four tunnels, and indicate the site of a fifth that may be German in origin, but there are no known geotechnical reports, or thematic maps prepared by geologists, for this or any of the smaller islands in the Channel Islands archipelago.

6.10  Geological Work for the Navy Pierre Renier of Guernsey has recently bought an untitled 1:25,000 map (Fig. 6.27) that, from the logo stamped upon it, at some time formed part of the collection of the German Hydrographical Institute: an organization founded postwar in 1945 to take over the nautical and meteorological observations made since 1867 by the

244

E. P. F. Rose

Fig. 6.26  German geological map of Sark and Brecqhou, original at scale of 1:25,000. The key (bottom right) distinguishes four rock types: hornblendite, hornblende schist, diorite gneiss and mica schist; and (bottom left) faults by a solid line, other rock unit boundaries by a dashed line. From Robins et al. (2012); reproduced by permission of the Bundesarchiv-Militärarchiv from file RW35/72

North German Naval Observatory. After reunification of Germany in 1990, the Institute transformed into the present Federal Maritime and Hydrographic Agency. The map has no title, but the key (Fig. 6.28) indicates that the purpose of the map was to show ground features associated with Guernsey’s coast. A reference number and date pencilled on the back of the map indicates that it was compiled in 1943— and so after Wehrgeologenstelle 4 had left the island on assignment to other duties. Notes pencilled in the bottom right margin give (in translation): subject: geology;

6  Guernsey and the German Army

245

Fig. 6.27  Untitled German ‘geological’ map of Guernsey, dated 1943 in pencil on the back, showing coastal features (Fig. 6.28). Reproduced by kind permission of Pierre Renier, Guernsey, from his private collection

compilation: Wehrgeologenstelle 6 at Naval Headquarters Commander-in-Chief West; comparison: [topographic map of] France, at 1:50,000. This was therefore perceived as a geological map, and one made by a geological centre/team. It is known that the German Navy made use of a few geologists in its Marine Geography agency,108 for which the German Army had an equivalent,109 as noted by Häusler (1995a) and Häusler and Willig (2000). However, there is no known record that the Navy had its own organization of numbered Wehrgeologenstellen. Wehrgeologenstelle 6 is therefore presumed to have been one of the 40 Army units established by November 1943, attached to the Naval Headquarters (in Paris) much as Wehrgeologenstelle 9 had earlier been attached to the headquarters of the Admiral of France (Sect. 6.2.1). However, from March 1942 until May 1944 Wehrgeologenstelle 9 was active on the Eastern Front (Häusler 1995a), so far away from Paris in 1943.

108 109

 MarGeo.  MilGeo.

246

E. P. F. Rose

Fig. 6.28  Enlargement of key to map shown as Fig. 6.27. Translation from the top down: (1) depth contour in the shallow sea; (2) rocky beach with cliff; (3) sandy beach; (4) pebble beach with boulders (symbols (2), (3) and (4) bracketed as ‘within and beneath the high tide range of the sea’); (5) coastal cliff in hard rock: field positions [i.e. trenches] impossible; (6) sand above high water boundary; (7) sand blown into dunes (symbols (6) and (7) bracketed as ‘ [construction of] field positions [i.e. trenches] easily possible); (8) zone of large trees; (9) protected water area inland; (10) sand; (11) gravel; (12) hard rock (building stone); (13) brick clay (brickworks); (14) water (symbols (10) to (13) bracketed as ‘extraction points’); (15) waterworks; (16) main water conduit; (17) bank, mole; (18) underground facility (tunnel); (19) valley opening through the cliff zone, sandy, route passable at low tide. Reproduced by kind permission of Pierre Renier

6  Guernsey and the German Army

247

Wehrgeologenstelle 6 had been stationed at the Military Geological Training and Equipment Centre at Sternberg/Neumark in Germany (now in western Poland) in May 1942, when it issued a report on raw materials in England, but it too had deployed fully to the Eastern Front by July 1943. It might possibly have supported the Naval Headquarters in Paris at some time between these dates, but Häusler (1995a, b) provides no record of this—or of any other of the Army’s Wehrgeologenstellen as specifically deployed in support of a Naval Headquarters. Presumably any such assignments were exceptional, or brief, or their documentary record was confined to the Naval chain of command and not copied to the Army headquarters in Berlin. None of the geologists described in this book is recorded as having served with Wehrgeologenstelle 6. However, Walter Wetzel remained in Paris until France was liberated by the Allies, being much ‘involved with the coastal fortifications’ of the Atlantic shore (Dietz et al. 1999, p. 11), although Wehrgeologenstelle 9 to which he was initially appointed served from at least April 1942 on the Eastern Front. It is therefore possible that the map illustrated as Fig. 6.27 is a further example of his handiwork. If not, it would be evidence that even more military geologists applied their expertise to the Channel Islands than the 14 whose names are known for certain (Table 10.2).

6.11  The Geologists After Guernsey Walter Wetzel returned to Kiel at the end of the war, bringing a substantial collection of aerial photographs of the French coast with him. The Geological Institute of the university had suffered bomb damage, but Wetzel resumed teaching there and as a schoolmaster for 7 years, until retirement in 1952 at the age of 65. He travelled to South America many times, both before and after retirement, and maintained a prodigious output of research publications until 2 years before his death, at the age of 91. He had generated 83 publications by the end of 1938; 14 in the war years 1939 to 1945; and 94 between 1947 and 1976 (Dietz et al. 1999). Many were palaeontological in focus, some pioneering aspects of micropaleontology (notably with respect to fossil dinoflagellates), others ranging more widely across disciplines within geology. He developed innovative techniques, and his many contributions to the study of geology earned him international recognition by the time of his death, on 17 April 1978. Friedrich Röhrer served at the Inspectorate of Land Fortification (West) from January 1942 to May 1943 (Häusler 1995a, b), finally returning to his professorial appointment at the University of Heidelberg. However, the university was subject to de-nazification by the Allies at the end of the war (Remy 2002), and in 1945 Röhrer became a schoolmaster110 once more, at the Robert-Bunsen School in Konstanz (Drüll 2013). The appointment was very brief: he died in Konstanz on 14 July 1945.

110

 Oberstudienrat = senior teacher.

248

E. P. F. Rose

TKVR Scherer had been re-assigned from France by March 1944 (Häusler 1995a, b), still to water-supply tasks, but to Wehrgeologenstelle 29: the unit providing geological support to the High Command of the Army,111 in Berlin. What became of him thereafter is not known. Bernhard Beschoren returned to mainland Europe in time to attend a conference for military geologists held at the Geological Institute of the University of Bonn, from 9 to 11 January 1943. The other 17 registrants included Friedrich Röhrer and Walter Wetzel (Häusler 1995a, p. 71). Beschoren was serving with Wehrgeologenstelle 17 by December 1943 (Häusler 1995b, p.  10, 60), a unit reporting to the liaison staff of the Inspectorate of Fortresses (West).112 Wehrgeologenstelle 17 had been operating in the Netherlands from formation in about March 1941, from at least October 1942 until September 1943 based primarily on Utrecht, but reporting for part of that time to Fortress Engineer Command XIV.113 From October to December 1943 it was briefly re-assigned to the liaison staff, and then from the end of December 1943 to July 1944 to Fortress Engineer Staff 15114 (Häusler 1995a, pp. 131–132). In 1944 Beschoren became a British prisoner-of-war. After the war, he returned to Survey work from 1946 to retirement in 1954, in the newly constituted Bavarian Geological Survey, although to areas composed mainly of Triassic sedimentary rocks (Beschoren 1955) or of molasse115 (Abele et al. 1955). He died on 24 April 1982, in Bavaria at Werneck/Unterfranken, at the age of 83 (Rose 2007). Dieter Hoenes, according to Häusler (1995a, p. 114), had by February 1943 been re-assigned to the Inspectorate of Land Fortification (West),116 but by April he was back with Wehrgeologenstelle 4, now responsible to Fortress Engineer Staff 24. In August he reported on studies in the Pyrenean region of southern France, in association with colleagues named Keilbach, Kobold and Reidl (the latter presumably the draughtsman of that name to have served with the team on Guernsey). By March 1944 he had been assigned to Wehrgeologenstelle 19, a unit led since 1943 by his colleague of 1941 Walter Tröger, and in 1944 reporting to Fortress Engineer Staff 14. However, his work was still based in southern France, for in March he reported on studies in the Cannes region (Häusler 1995a, pp. 133–134). By July he had been posted once more, to Wehrgeologenstelle 26, which then reported to Fortress Engineer Command IV. There he helped to compile a report on water barriers relating to the 158th Infantry Division, in the Biarritz region of SW France (Häusler 1995a, pp. 139–140). However, as the Allies advanced swiftly across France in the summer of 1944, German military geological work in that country came speedily to an end, and Hoenes too became a British prisoner-of-war.

 Oberkommando des Heeres.  Verbindungsstab Inspektion der Festungen West. 113  Festungspionier-Kommandeur XIV. 114  Festungspionierstab 15. 115  Thick deposits of continental and marine clastic sedimentary rocks, mainly sandstones and shales, that formed in near-shore environments in front of rising mountain chains. 116  Inspektion der Landesbefestigung West. 111 112

6  Guernsey and the German Army

249

During the war Hoenes had, in 1943, obtained appointment to the University of Berlin, and later promotion as lecturer117 in mineralogy. However, 2 months as a prisoner-of-war lost him this position postwar, and he began research again at Freiburg, on the Black Forest region of Germany. He obtained appointment to the mineralogy institute at Kiel in 1947, moving to Freiburg as a lecturer in 1948, before promotion as Professor for Geology and Mineralogy at the ‘Polytechnic’ (now Fridericiana University) of Karlsruhe in 1953. His wartime work generated only a single publication directly, on oolitic iron ores in Normandy (Hoenes and Tröger 1945). His research interests otherwise continued to focus on the Black Forest region postwar as prewar. He published accounts of Variscan granitic rocks intruding pre-Variscan gneisses (Hoenes 1947a, b); basement rocks of the southern Black Forest (Hoenes 1948a); petrogenetic relations of metabasites and mixed gneisses (Hoenes 1948b); granodioritic mixed rocks plus hybrid and normal granites around the gneiss core of the central southern Black Forest (Hoenes 1949a); magmatic differentiation (Hoenes 1949b); Black Forest excursion guides (Hoenes et  al. 1949; Hoenes and Schneiderhöhn 1957); relations between magmatic and metamorphic rocks (Hoenes 1950); structure of the pre-Variscan gneiss (Hoenes 1952, 1956); an obituary for the German mineralogist and petrographer Otto Heinrich Erdmannsdörffer (Hoenes 1955a); part of a textbook on petrographic microscopy (Hoenes 1955b); and accounts of cordierite-bearing gneiss (Hoenes 1955c), plus chromite (Hoenes and Volkert 1954) and kaolinite (Hoenes and Behne 1955) ore deposits. A rising star in academic geology, he died at only 44 years of age, on 10 August 1955, whilst on a field trip to Norway (Rose 2007). Gottfried Reidl died during the last few weeks of war,118 on 24 April 1945 at Merseburg: a town in the south of the east German state of Saxony-Anhalt that was subject to intensive Allied bombing during World War II and finally occupied by Soviet forces. His death, according to a note as yet unsubstantiated by official documents, occurred whilst he was a prisoner-of-war. Rolf Thienhaus served with Wehrgeologenstelle 1 later in the war (Häusler 1995a, p. 110) and so came to be a true military geologist rather than a mining engineer. He was in the Harz Mountains of central Germany as the war came to an end (Pilger 1969; Grabert 1969) and hid himself there in the wild, watching American troops pass by, before returning to his home in Oldenburg, nearby in Lower Saxony. Three years in private practice as a consultant providing geological advice on building construction materials and water-supply problems ended with his appointment in May 1948 as the geologist for a steel company based in the region of Sieg, some 50 km SW of Cologne in North Rhine-Westphalia. In 1953, however, he moved to another (mining) company, and geological fieldwork prospecting for the iron and steel industry took him to many countries worldwide in the search for workable iron ore. From 1953, he became the author or co-author of some 26 academic articles, mainly on ore deposits of the countries he had visited. Based on this experience he was appointed in 1964 to the teaching staff of the Technical University (formerly 117 118

 Dozent.  Archives of the Landesmuseum Linz: http://www.zobodat.at accessed 14 March 2016.

250

E. P. F. Rose

mining academy) of Clausthal, Lower Saxony, in the SW part of the Harz Mountains. He was promoted as Professor for the Geology of Non-European Countries at Clausthal from May 1967. He was thus recognized as an applied geologist of some distinction, but he died the next year, after a long illness: during the night of 1 June 1968.

6.12  Conclusion It is therefore clear that five military geologists were used as such to compile at least 13 technical reports and nine thematic maps for Guernsey: TKVRs Wetzel, Röhrer, Scherer, Beschoren and Hoenes. Beschoren and Hoenes were respectively the leader and deputy leader of Wehrgeologenstelle 4, and assisted by Corporals Hopp and Ludwig, and by Lance-Corporals Gehrt and Dr. Reidl. Reidl was a graduate geologist and contributed significantly to at least one (water supply) map and its corresponding report. Lance-Corporal Dr. Rolf Thienhaus also contributed geological expertise, but through an Engineer Mining Company rather than the military geological organization as such (although he was later in the war to be appointed to a military geological unit). It seems, therefore, that at least seven geologists can be credited with wartime work on Guernsey. Of these, Wetzel, Beschoren, Hoenes and Thienhaus returned to careers of some distinction as professional geologists postwar; Reidl died during the war; and Röhrer survived the war by only a few months. What became of Scherer is not known. The thematic maps were all at a scale of 1:25,000, corresponding closely in scale to the topographical base maps prepared prewar by the British Ordnance Survey but differing significantly from the 1:63,360 and 1:10,560 topographical maps in more general British use. They included the first geological maps for Guernsey as a whole to be compiled at this scale; the first truly military geological maps for any part of the British Isles; a map indicating the occurrence of varied raw materials and maps indicating both depth to groundwater and potential sources for potable water supplies, which were also innovative in terms of scale and content with respect to coeval or earlier British maps. The German Army’s geotechnical achievement on Guernsey was therefore considerable. It was enhanced even further by geologists of the Air Force, as described next, in Chap. 7.

References Abele G, Beschoren B, Dehm R, Erb L, Fuchs B, Ganss O, Kiderlen H, Nathan H, Neumaier F, Schmidt-Thome P, Stephan W (1955) Erläuterungen zur Geologischen Übersichtskarte der Süddeutschen Molasse 1:300,000. Bayerisches Geologisches Landesamt, Munich Beschoren B (1926) Cenoman und Turon der Kreidemulde von Sack bei Alfeld. Neues Jahrbuch für Mineralogie, Geologie und Paläontologie. Beilage-Band. Abteilung B 55:87–132

6  Guernsey and the German Army

251

Beschoren B (1927) Cenoman und Turon der Gengend von Unna und Werl in Westfalen. Neues Jahrbuch für Mineralogie, Geologie und Paläontologie. Beilage-Band. Abteilung B 58:1–49 Beschoren B (1932) Über einheimisches Diluvium in der Umgebung von Burgsdorf in Hannover. Jahrbuch der preussischen geologischen Landesanstalt 52:79–85 Beschoren B (1934) Die erdölgeologische Literatur Deutschlands bis 1933. Archiv für Lagerstättenforschung, Berlin 60:91 Beschoren B (1935a) Über alluviale Neubildungen in historischer Zeit in Gebiet von Elbe und Oder. Jahrbuch der preussischen geologischen Landesanstalt 55:292–304 Beschoren B (1935b) Zur Geschichte des Havellandes und der Havel während des Alluviums. Jahrbuch der preussischen geologischen Landesanstalt 55:305–311 Beschoren B (1935c) Über jungdiluviale Staubeckentone zwischen Havel und Oder. Jahrbuch der preussischen geologischen Landesanstalt 55:395–428 Beschoren B (1936) Über das Alluvium im Leinetal bei Neustadt a. Rbg und im Allertal bei Celle. Jahrbuch der preussischen geologischen Landesanstalt 56:196–204 Beschoren B (1937) Über eine Oslandschaft in der Grenzmark ostlich Schwerin a W. Jahrbuch der preussichen geologischen Landesanstalt 57:65 Beschoren B (1955) Erläuterungen zur geologischen Karte von Bayern 1:100,000 Blatt Nr 510, Schweinfurt. Bayerisches Geologisches Landesamt, Munich Beschoren B, Blese W, Born A, Ebert A, Ernst W, Haupt O, Hilzheimer M, Jungst H, Kahrs E, Kauenhowen W, Leuchs K, Löwe F, Ortie GF, Potonié R, Quenstedt W, Rettschlag W, Stach E, Stappenbeck R, Wetzel W, Woldstedt P (1927) Festband J. F. Pompeckj zum 60. Geburtstage gewidmet von seinen Schülern und dem Verlag. Neues Jahrbuch für Mineralogie, Geologie und Paläontologie, Beilage-Band 58, Abteilung B, Geologie und Palaeontologie 1927:1–687 Brill R (1932) Pforzheim sheet. Baden Blatt 64 Pforzheim, Württemburg Blatt 54 Wurmberg. Geologische Specialkarte des Grossherzogtums Baden. Badische Geologische Landesanstalt, Karlsruhe Daur E, Röhrer F (1921) Die Pforzheimer Typhusepidemie vom Jahre 1919, ihre Entstehung und die zur Vermeidung einer Wiederholung getroffenen Massnahmen. Das Gas- und Wasserfach: Journal für Gasbeleutung und Wasserversorgung 61:227 Dietz LF, Sarjeant WAS, Mitchell TA (1999) The dreamer and the pragmatist: a joint biography of Walter Wetzel and Otto Wetzel, with a survey of their contributions to geology and micropalaeontology. Earth Sciences History 18:4–50 Drüll D (2013) Heidelberger Gelehrtenlexicon 1803–1932. Springer Verlag, Berlin Eckart WU, Sellin V, Wolgast E (eds) (2006) Die Universität Heidelberg im Nationalsozialismus. Springer Medizin Verlag, Heidelberg Gavey E (2001) A guide to German fortifications in Guernsey. Guernsey Armouries, Guernsey Gavey E, Powell S (2012) German tunnels in Guernsey, Alderney & Sark. Festung Guernsey, Guernsey Ginns M (1993) German tunnels in the Channel Islands. Archive Book No. 7. Channel Islands Occupation Society, Jersey Ginns M (1994) The Organisation Todt and the Fortress Engineers in the Channel Islands. Archive Book No. 8. Channel Islands Occupation Society, Jersey Grabert H (1969) [Obituary] Rolf Thienhaus. Decheniana 122:11–13, 1 plate Häusler H (1995a) Die Wehrgeologie im Rahmen der Deutschen Wehrmacht und Kriegswirtschaft. Teil 1: Entwicklung und Organisation. Informationen des Militärischen Geo-Dienstes, Vienna 47:1–155 Häusler H (1995b) Die Wehrgeologie im Rahmen der Deutschen Wehrmacht und Kriegswirtschaft. Teil 2: Verzeichnis der Wehrgeologen. Informationen des Militärischen Geo-Dienstes, Vienna 48:1–119 Häusler H (2000) Die Österreichische und Deutsche Kriegsgeologie 1914–1918. Informationen des Militärischen Geo-Dienstes, Vienna 75:1–161

252

E. P. F. Rose

Häusler H, Willig D (2000) Development of military geology in the German Wehrmacht 1939-45. In: Rose EPF, Nathanail CP (eds) Geology and warfare: examples of the influence of terrain and geologists on military operations. Geological Society, London, pp 141–158 Hoenes D (1937) Gesteine und Erzlagerstätten im Schwarzwälder Grundebirge zwischen Schauinsland, Untermünstertal und Belchen. Neues Jahrbuch für Mineralogie, Geologie und Paläontologie. Beilage-Band Abteilung A 72:265–346 Hoenes D (1939) Über den Mineralbestand der sauren Hochofenschlacken. Zentralblatt für Mineralogie, Geologie und Paläontologie. Beilage-Band Abteilung A 1939:257–271 Hoenes D (1940) Magmatische Tätigkeit, Metamorphose und Migmatitbildung in Grundgebige des südwestlichen Schwarzwaldes. Neues Jahrbuch für Mineralogie, Geologie und Paläontologie. Beilage-Band Abteilung A 76:153–256 Hoenes D (1947a) Zur genetischen Gliederung des variskischen Magmatismus im südlichen Scharzwald. Mitteilungsblatt der Badischen Geologischen Landesanstalt 1947:15–17 Hoenes D (1947b) Über die Beziehangen zwischen Granit und Gneiss und die magmatische Entwicklung in Grundgebirge des südlichen Schwarzwaldes. Fortschritte der Mineralogie, Kristallographie und Petrographie 26:65–66 Hoenes D (1948a) Petrogenese im Grundgebirge des Südschwarzwaldes. Heidelberger Beiträger zur Mineralogie und Petrographie 1:121–202 Hoenes D (1948b) Zur Frage der petrogenetischen Stellung der Metabasite und Mishgneiss des Südlichen Schwarzwaldes. Mitteilungsblatt der Badischen Geologischen Landesanstalt 1948:19–22 Hoenes D (1949a) Gesetzmässigkeiten in der Verteilung der Gneise, Granite und Mischgesteine des zentralen Sudschwarzwaldes und ihre Bedeutung für die Genese des Grundgebirges. Mitteilungsblatt der Badischen Geologischen Landesanstalt 1949:8–12 Hoenes D (1949b) Kriterien für die genetischen Typen der variskischen Magmatite des südlichen Schwarzwaldes. Fortschritte der Mineralogie, Kristallographie und Petrographie 28:29–32 Hoenes D (1950) Magmatische Entwicklung und Tiefenstufen im Grundgebirge der Vogesen und des Schwarzwaldes. Bericht der Naturforschungen Gesellschaft zu Freiburg im Breisgau 39:197–223 Hoenes D (1952) Über den Mechanismus der Gesteinsverformung in den Granuliten des prävariskischen Gneisgebirges im südlichen Schwarzwald. International Geological Congress, 19th Session, Algiers, Résumé des Communications, 18 Hoenes D (1955a) Otto Heinrich Erdmannsdörffer, in memoriam. Heidelberger Beiträger zur Mineralogie und Petrographie 4:I–XIV Hoenes D (1955b) Mikroskopische Grundlagen der technischen Gesteinskunde. In: Freund H (ed) Handbuch der Mikroskopie in der Technik, Band 4, Part 1. Umschau Verlag, Frankfurt, pp 323–695 Hoenes D (1955c) Der Para-Cordieritgneis der Bohrung Scherstetten 1 und seine Beziehungen zu den cordieritführenden Metatexiten des Schwarzwaldes. Geologica Bavaria 24:102–136 Hoenes D (1956) Der prägranitische Bau des Gneisgebirges im südlichen Schwarzwald und seine Abteilung aus den Fremdgesteinsinhalt der hybriden Granite. Heidelberger Beiträger zur Mineralogie und Petrographie 5:272–288 Hoenes D, Behne W (1955) Die Kaolinlagerstätte von Geisenheim (Rheingau). Heidelberger Beiträger zur Mineralogie und Petrographie 4:412–433 Hoenes D, Schneiderhöhn H (1957) Erläuterungen zur geologisch-lagerstättlichen Excursion von Baden-Baden über der Schwarzwaldhochstrasse zu den Erzgängen des mittleren Kinzigtales an 13 Oktober 1957. Zeitschrift für Erzbergbau und Metallhüttenwesen 5:344–350, 390–395, 443 Hoenes D, Tröger E (1945) Lagerstätten oolithischer Ersenerze in Nordwestfrankreich. Neues Jahrbuch für Mineralogie, Geologie und Paläontologie. Beilage-Band. Abteilung A 79:192–255 Hoenes D, Volkert G (1954) Das metallurgische Verhalten der Chromerze. Archiv für Metallkunde 25:1–10 Hoenes D, Mehnert KR, Schneiderhöhn H (1949) Führer zu petrographisch-geologischen Excursionen im Schwarzwald und Kaiserstuhl. Schweizerbart, Stuttgart

6  Guernsey and the German Army

253

Mehnert KR (1964) Walter Ehrenreich Tröger. Fortschr Mineral 41:8–10 Parkinson J, Plymen GH (1929) The Channel Islands. In: Evans JW, Stubblefield CJ (eds) Handbook of the geology of Great Britain. Allen, London, pp 514–528 Partridge C, Wallbridge J (1983) ‘Mirus’: the making of a battery. Ampersand Press, Alderney Pilger A (1969) Gedenkansprache anlässlich der Trauerfeier der Technischen Universität Clausthal für Professor Dr. Rolf Thienhaus am 23 November 1968. Mitteilungsblatt der Technischen Universität Clausthal 18:13–20 Reidl G (1939) Bericht über geologische Feldarbeiten 1938. Verhandlungen der Zweigstelle Wien der Reichsstelle für Bodenforschung (früher Geologische Bundesanstalt) 1938:64–65 Reidl G (1940) Ein Knochenfund im Keller der Geologischen Landesanstalt Wien. Verhandlungen der Zweigstelle Wien der Reichsstelle für Bodenforschung (früher Geologische Bundesanstalt) 1939:109–111 Reidl G (1941) Über eine neue Spatangidenart Plagiobrissus abeli nov. spec. aus dem Torton von Müllendorf (ehem. Burgenland). Berichte der Reichsstelle für Bodenforschung (Zweigstelle Wien) 1941:24–29, figures 1, 2 Remy SP (2002) The Heidelberg myth: the nazification and denazification of a German university. Harvard University Press, Cambridge MA Robins NS, Griffiths KJ, Merrin PD, Darling WG (2002) Sustainable groundwater resources in a hard-rock island aquifer—the Channel Island of Guernsey. Geological Society, London, Special Publications 193(1):121–132 Robins NS, Rose EPF, Cheney CS (2012) Basement hydrogeology and fortification of the Channel Islands: legacies of British and German military engineering. In: Rose EPF, Mather JD (eds) Military aspects of hydrogeology. Geological Society, London, Special Publications, vol 362, pp 203–222 Röhrer F (1916) Geologische Untersuchungen der Beziehungen zwischen den Gesteinsspaltern, der Tektonik und dem hydrographischen Netz im nördlichen Schwarzwald und südlichen Kraichgau. Jahresberichte und Mitteilungen des Oberrheinischen Geologischen Vereins, NF Bd 6(1):8–85 Röhrer F (1922a) Geologische Untersuchungen der Beziehungen zwischen den Gesteinsspaltern, der Tektonik und dem hydrographischen Netz im nördlichen Schwarzwald und südlichen Kraichgau: 1. Teil Die gemeinen Klüfte und die Harnische, Tübingen, Laupp Röhrer F (1922b) Geologische Untersuchungen der Beziehungen zwischen den Gesteinsspaltern, der Tektonik und dem hydrographischen Netz im nördlichen Schwarzwald und südlichen Kraichgau: 2. Teil Bemerkungen zur Tektonik Suudwestdeutschlands, Tübingen, Laupp Röhrer F (1923a) Beiträge zur Hydro-Geologie des nordöstlichen Schwarzwaldes: die Grösseltalquellen und die Reutbachquellen. Verhandlungen des Naturwissenschaftlichen Vereins, Karlsruhe i B 29:75–96 Röhrer F (1923b) Über Quellenuntersuchungsmethoden. Das Gas- und Wasserfach: Journal für Gasbeleuchtung und Wasserversorgung 66:81–82, 99–101, 117–120, 131–133 Röhrer F (1924) Das Ganggebeit von Neuenbürg und Pforzheim. In: Henglein M (ed) Erz- und Minerallagerstätten des Schwarzwaldes. Schweizerbart, Stuttgart, pp 121–139 Röhrer F (1925) Von der Geologie und Oberflächengestaltung des Enz-Pfinzgaus. In: Busse HE (ed) Badische Heimat 12. Braun, Karlsruhe, pp 10–19 Röhrer F (1927) Die Gegend von Baden-Baden. Geogr Z 33:204–209 Röhrer F (1929) Das Untergrundwasser, seine Bildungsweise und seine Erscheinungsformen. Das Gas- und Wasserfach: Journal für Gasbeleuchtung und Wasserversorgung 72:174–180, 199–205 Röhrer F (1933) Über den Nitratgehalt der Tiefenwässer. Geol Rundsch 23A(special issue):315–331 Röhrer F (1934) Über die Durchlässigkeit von Gesteinen. Zeitschrift der Deutschen Geologischen Gesellschaft 85 (for 1933):477–478 Röhrer F (1941) Wehrgeologische Überblick über die Trias in Lothringen. In: Anon. (ed) 6 Wehrgeologischer Lehrgang in Heidelberg. Reichsdruckerei, Berlin, pp 31–34

254

E. P. F. Rose

Rose EPF (2005a) Work by German military geologists on the British Channel Islands during the Second World War. Part 1: pioneering studies by Walther Klüpfel (Jersey and Alderney), Walter Wetzel (Guernsey and Alderney), and Friedrich Röhrer (Guernsey). Channel Islands Occup Rev 33:93–120 Rose EPF (2005b) Specialist maps of the Channel Islands prepared by German military geologists during the Second World War: German expertise deployed on British terrain. Cartogr J 42:111–136 Rose EPF (2007) Work by German military geologists on the British Channel Islands during the Second World War. Part 2: Bernhard Beschoren, Dieter Hoenes, and the role of Wehrgeologenstelle 4 on Guernsey and Alderney. Channel Islands Occup Rev 35:93–114 Rose EPF, Willig D (2009) Work by German military geologists on the British Channel Islands during the Second World War. Part 3. Reports with contributions by Walther Klüpfel and Rolf Thienhaus now preserved at the Training and Education Centre of the Bundeswehr Geoinformation Office, Fürstenfeldbruck. Channel Islands Occup Rev 37:105–118 Rose EPF, Willig D (2013) Work by German military geologists on the British Channel Islands during the Second World War. Part 5: work by Luftwaffe geologist Professor K. G. Schmidt and Hilfsgeologe Dr. K. Diebel, for tunnelling (in general and on Jersey) and water supply (on Alderney). Channel Islands Occup Rev 41:78–101 Schadler J (1942) Berichte über wissenschaftliche Tätigkeit im Gau (1940 und 1941). Landesmuseum 2. Bodenforschung. Jahrbuch des Vereines für Landeskunte und Heimatpflege im Gau Oberdonau (früher Jahrbuch des Oberösterreichischen Musealvereins). 90. Band. Verlag J. Wimmer, Linz, pp 323–338 Seibold E (1978) Walter Wetzel 27.2.1887–17.4.1978. Meyniana 30:1–6 Thienhaus R (1940a) Die Faziesverhältnisse im Südwestteil der Attendorfer Mulde und ihre Bedeutung für die Stratigraphie des Bergish-Sauerlandischen Mitteldevons. Abhandlung der Reichsstelle für Bodenforschung, Berlin, 199, 77 pp, 11 figs, 5 tables, 1 map Thienhaus R (1940b) Die Schwerspatgänge des Richelsdorfer Gebirges. Zeitschrift für angewandte Mineralogie 3:21–52 Tröger WE (1956) Dieter Hoenes, in memoriam. Heidelberger Beiträger zur Mineralogie und Petrographie 5:171–176 Wetzel W (1911) Faunistische und stratigraphische Untersuchung der Parkinsonienschichten des Teutoburger Waldes bei Bielefeld. Palaeontographica 58:139–278 Wetzel W (1941a) Wehrgeologische Arbeiten an der Kanalküste. In: Anon. (ed) 6 Wehrgeologischer Lehrgang in Heidelberg. Reichsdruckerei, Berlin, pp 65–68 Wetzel W (1941b) Geologie beim Flugplatzbau. In: Anon. (ed) 6 Wehrgeologischer Lehrgang in Heidelberg. Reichsdruckerei, Berlin, pp 69–70 Wetzel W (1941c) Stellungsbau und Wasserversorgung auf Geestböden. In: Anon. (ed) 6 Wehrgeologischer Lehrgang in Heidelberg. Reichsdruckerei, Berlin, p 155 Wilser JL (1941) Die wehrgeologischer Karte 1:300,000. In: Anon. (ed) 6 Wehrgeologischer Lehrgang in Heidelberg. Reichsdruckerei, Berlin, pp 125–126 Wilser JL, Becksmann E (1941) Die GAuSt Heidelberg. In: Anon. (ed) 6 Wehrgeologischer Lehrgang in Heidelberg. Reichsdruckerei, Berlin, pp 127–128 Wilson FE (1970) Railways in Guernsey, with special reference to the German steam railways. Paramount-Lithoprint, Guernsey

Chapter 7

Guernsey and the German Air Force Edward P. F. Rose

Abstract  A Field Works Office (Feldbauamt) was established by the German Air Force (Luftwaffe) on Guernsey by April 1942, containing at least one geologist appointed to serve as such, as a uniformed official equivalent in rank to an Air Force captain. Dr. Hans Schneider held the appointment until at least late July 1943, guided initially by a visit from his geological superior in France, Professor K.G.  Schmidt, and assisted by Corporal Franz Schulte. Subsequently in 1943, Schulte was promoted to succeed Schneider as the Luftwaffe’s geologist resident on Guernsey, following Schneider’s transfer to duties in Southern Italy. The geologists are known to have contributed to a variety of ground investigations: for potential extension of Guernsey’s airfield; for building construction associated with the airfield’s fortification against aerial attack; for site investigation and water supply associated with emplacement of batteries of heavy anti-aircraft guns; and for tunnelling to create underground storage facilities safe from potential Allied bombing—contributions to works that in total were intended to make Guernsey one of the most heavily defended territories in Europe. After the war, Schneider progressively achieved distinction in his home province of Westphalia in Germany as a hydrogeologist and university professor.

7.1  Introduction As on Jersey, the airport was the first part of Guernsey to be seized by German forces. This followed the landing on 30 June 1940 of an Air Force (Luftwaffe) pilot on his own initiative whilst flying a routine reconnaissance in the region (Forty 1999). Arrival was thus a day earlier than on Jersey (a landing described in Sect. 5.1).

E. P. F. Rose (*) Department of Earth Sciences, Royal Holloway, University of London, Egham, Surrey, UK e-mail: [email protected] © Springer Nature Switzerland AG 2020 E. P. F. Rose (ed.), German Military Geology and Fortification of the British Channel Islands During World War II, Advances in Military Geosciences, https://doi.org/10.1007/978-3-319-22768-9_7

255

256

E. P. F. Rose

In the months and years of occupation that followed, Luftwaffe personnel became a major component of the island’s garrison. The largest concentration of Luftwaffe troops in the Channel Islands as a whole, and aircrew in particular, developed on Guernsey. Luftwaffe troops operated and developed the airfield; introduced and operated batteries of anti-aircraft guns with their associated searchlights to provide protection from attack from the air; and tunnelled underground to create storage facilities proof against bombing. Guernsey’s airfield had been officially opened only on 5 May 1939, so was very recent in construction at the time of its capture. From near-contemporary maps its runway was as now approximately east–west in orientation (Rose and Willig 2012), but then with a grass rather than a hard surface (Forty 1999). Early in the German occupation of the Channel Islands but during the aerial Battle of Britain, squadrons1 of Messerschmidt Bf109 single-seater fighter aircraft operated out of Guernsey. They formed part of an elite fighter wing (Jagdgeschwader 53), known from its distinctive emblem as the ‘Pik As’ (Ace of Spades), based primarily in France at this time. However, elements from two of its three component ‘Gruppe’2 reportedly made use also of Guernsey (Prien 1997; Weale 2007). During the Battle of Britain, the Jagdgeschwader as a whole was credited as one of the most effective Luftwaffe fighter units, claiming 258 kills for 51 pilots killed or captured. Forty (1999, p. 109) records that in addition to Jagdgeschwader 53, elements of Jagdgeschwader 27, also flying Messerschmidt Bf109s, operated from Guernsey when it served as a forward fighter base during the Battle of Britain, but that after the Battle, ‘air activity at Guernsey decreased’. Quoting a Mr. Le Page, he notes that soon ‘the only aircraft left were a Henschel Hs126 [a two-seat reconnaissance aircraft] and a few Dornier Do17s [operated with a four-man crew as a reconnaissance aircraft or as a bomber, notably against Channel convoys]. A Fieseler Storch [a versatile two-seat army co-operation/reconnaissance aircraft with remarkable short take off and landing capabilities] replaced the Hs126 in 1941’. By 1943 the only aircraft to be permanently based in the Channel Islands was a Focke-Wulf Fw189 [a two-seat short-range reconnaissance aircraft], used by the paramilitary construction agency Organisation Todt, then at work on Channel Island fortifications, as a courier plane shuttling between Guernsey and Dinard in France. German troop numbers for the Channel Islands as a whole seemingly peaked in August 1942, with an approved establishment of 36,960 (18,460 army, 9500 Luftwaffe, 4100 navy, 3500 supply, and 1400 construction personnel), although doubt has been expressed that all these established posts were actually filled (Cruickshank 1975). Of the Luftwaffe troops, most (7000) were engaged in anti-­ aircraft defence (guns and searchlights) and relatively few (2500) were aircrew with their ground support. By the beginning of May 1943 the actual total troop strength on the islands had fallen to 26,800: 6 months later it had fallen further, to 23,700,

 II/Jagdgeschwader 53: Gruppe II of fighter wing number 53, each ‘Gruppe’ comprising a headquarters section and three squadrons [Staffel] of aircraft. 2  II/JG-53 and possibly also I/JG-53. 1

7  Guernsey and the German Air Force

257

and was much the same in July 1944 once the Allied liberation of France had begun. However, Guernsey’s defences came to include 33 anti-aircraft artillery sites (Ramsey 1981, p. 18). Tunnelling to create facilities underground was planned to be even more intensive on Guernsey than on Jersey: 16 sites were proposed or under construction by May 1942, increased to 41 by June 1943. The intended use changed with time, but these ‘tunnels’ included a Luftwaffe fuel store (Ho 4), munition store (Ho 10), other facility (Ho 29), and a complex for aircraft and other supplies (Ho 41) (Gavey and Powell 2012). To facilitate construction work on behalf of the Luftwaffe, a Field Works Office (Feldbauamt 14) was established on Guernsey by April 1942, and this included a geologist as such at this time. Lieutenant (later Captain) Walther Klüpfel of the German Army was the first military geologist to be stationed on the Channel Islands, serving in Jersey from July 1941 (as described in Chap. 4). From a diary of events annexed to one of his reports (Table 4.1), it is known that he was visited by a ‘Dr. Schneider’, a geologist supposedly from a Luftwaffe geological team,3 on 10 April 1942. Klüpfel records that he gave Schneider a copy of a geological map and took him on a geological excursion on Jersey on 11 April. No further visits by Dr. Schneider are recorded, so presumably this was a liaison visit arranged soon after Schneider’s arrival in the Channel Islands, to establish the geological basis for future work.

7.2  The Geologist Hans Schneider Rose and Willig (2012) have identified this ‘Dr. Schneider’ as Hans Schneider (Fig. 7.1), who served as a Luftwaffe geologist throughout the war and afterwards became a distinguished hydrogeologist and university professor. Records at the University of Münster show that Hans [August] Schneider was born at Bielefeld, now one of the 20 major cities in Germany and the cultural centre of eastern Westphalia, on 17 July 1914, and that he studied physics, mathematics, and natural science at the universities of Munich (in Bavaria) and Münster. He graduated from Münster in 1938 with a doctoral thesis on geological conditions in the Baumberg region, to the west of Münster, on the eastern bank of the River Rhine (Schneider 1940c). The thesis was supervised by Professor Friedrich Schuh, who was himself to serve as a senior military geologist with the German Army from 1940 to 1942, first in Belgium and finally in Ukraine (Häusler 1995a, b). Schneider apparently began his association with the Luftwaffe in the year of his graduation, at the age of 24, by service with his local unit. In an authoritative account of German military geologists of World War II, the Austrian professor Hermann Häusler (1995b, p. 81) lists only one with the surname ‘Schneider’, an ‘H. Schneider’

 Geologenstelle.

3

258

E. P. F. Rose

Fig. 7.1  Professor Dr. Hans Schneider, from the prologue (by Friedrich Nöring) of a book published to celebrate his 65th birthday (Schneider 1981); courtesy of Professor Eckehard Löhnert. From Rose and Willig (2012)

who served with the Air District Command based at Münster4 from December 1938 to December 1941, and with the Air District Command based in France at Paris5 from December 1941 to November 1943. Initially therefore Schneider served in the NW German province in which he was born and in which he made his life-long civilian home. Thereafter, at some time between December 1941 and early April 1942, he was re-assigned, to Lufgaukommando Paris, and to its outstation on the Channel Islands. It is evident from Schneider’s publications that by the time he arrived in the Channel Islands he was already an experienced hydrogeologist. Including his doctoral thesis, he had published 12 scientific papers by that time, most of them relating to aspects of groundwater occurrence (Schneider 1938, 1939, 1940a, b, c, 1941a, b, c, d, e, 1942; Schneider and Wehrli 1939). His only subsequent wartime paper (Schneider and Wehrli 1943), on the geology of a district some 40  km north of Münster, extended the published account of earlier work by the two authors. Schneider was thus a relatively young man, some 28 years old, when he arrived on Guernsey, but already a graduate geologist with a significant publication record in applied geology. He was also a veteran of 4 years of military service.

 Luftgaukommando Münster (Westfalen).  Luftgaukommando Paris.

4 5

7  Guernsey and the German Air Force

259

7.3  Hans Schneider on Guernsey From his visit to the Army’s geologist Walther Klüpfel on Jersey in April 1942 (Table 4.1), it is evident that Schneider had been assigned to Luftwaffe geological work on the Channel Islands by that time. Assignment was presumably recent, or there would have been little purpose in such a liaison visit: there is no record of any further such visit. April 1942 was the month in which the German Army initiated its own most intensive geological activity (as described in Chaps. 4 and 6). Presumably this was no coincidence. The Luftwaffe and the army apparently coordinated effort that provided a speedy response to the Führer Adolf Hitler’s directive initiating the main construction phase of the coastal fortification system that became known as the Atlantic Wall (Sect. 3.3). Although Schneider seemingly began work by his introduction to Channel Islands geology on Jersey, the only known case histories of the work that followed are three ‘expert opinions’ (Gutachten) documenting work in Guernsey. These three are now amongst documents preserved in the archives of the Bundeswehr Geoinformation Centre, transferred early in 2010 from Fürstenfeldbruck (Rose and Willig 2009, 2012) to Euskirchen in Germany. ‘Expert opinions’ were shorter documents than reports6 strictly so-called, but provided more technical detail than ‘short reports’7 (as described earlier: Sect. 6.4). Schneider’s were dated January, June, and July 1943, and all generated for the Luftwaffe. Following each title, the reference information given on Schneider’s three ‘expert opinions’ shows that all were written when he was serving as ‘the geologist’8 at ‘Air Force Field Works Office 14’9 when this was based in the Channel Islands10 within the ‘Air District’11 for ‘Western France’.12 The documents were prepared on deployment13 rather than at the District headquarters, in Paris. All were originally classified as ‘secret’,14 the last two also as dealing with ‘command matter’,15 and authorship is credited to Dr. Schneider as the ‘authorized expert’.16 Schneider’s status of ‘government construction officer’17 is of an intermediate-level professional grade long (and still) in use in the German civil service, but the qualification ‘d.B’ signifies that he was a Wehrmachtsbeamter des Beurlaubtenstandes: on leave from a civilian  Berichte.  Kurzberichte. 8  Der Geologe. 9  Luftwaffen Feldbauamt 14. 10  K.-I. = Kanalinseln. 11  Luftgau. 12  Westfrankreich. 13  O.U. = Ortsunterkunft: military postal service location undefined. 14  geh. = geheim. 15  Kdos = Kommandosache. 16  Sachbearbeiter. 17  Regierungs Bauassessor. 6 7

260

E. P. F. Rose

appointment as a Bauassessor whilst enlisted for service in uniform with the armed forces (so with age-related status equivalent to that of an Army or Air Force captain: see Sect. 3.6). The first of Schneider’s ‘opinions’ provides an assessment of ground conditions with regard to a planned extension of Guernsey’s airfield; the second conditions with regard to building work in the vicinity; and the third conditions at sites for the brigade of heavy anti-aircraft guns scheduled to provide the island’s main defence from aerial assault or bombardment.

7.3.1  Ground Conditions and Potential Airfield Extension The opening paragraphs of this ‘opinion’18 define its purpose: to provide an assessment, according to a specified code of practice, of ground conditions (bearing capacity, permeability, and workability) within boundaries defined for potential extension of the airfield (Fig. 7.2). Three main runways had been proposed: 1 . SE–NW in orientation, with length 1200 m and breadth 600 m; 2. NE–SW in orientation, with length 1500 m and breadth 550 m; 3. east–west in orientation, with length 2000 m and breadth 600 m; with the possibility that an additional north–south runway, 1200 m long by 300 m broad, might be included. A ground investigation was conducted by means of samples taken during creation of boreholes 80 mm in diameter, by means of a hand-operated shell-and-auger rig (Fig. 7.3). In total 196 boreholes (Table 7.1) were put down to depths typically of about 5 m, but ranging at rare extremes from 8 to 2 m (cf. Fig. 7.4), the depth of boring being limited by the depth to weathered/fractured bedrock and so the ease of ground penetration by the auger. The boreholes cover the areas of the four planned runways according to a series of line transects on which holes are typically spaced 100 m apart (Figs. 7.5 and 7.8). Holes 1–22 (completed by 10 June 1942) were put down to form a grid extending a short distance NNE from the SE corner of the area (essentially the southern part of the potential north–south runway); holes 23–43 (mostly completed by 19 June) continue north but on a new grid orientation to the NW (and so cover ground mostly common to the proposed north–south, SE–NW, and NE–SW runways); holes 44–86 (mostly emplaced by 8 July) complete the eastern area of the airfield (and so the area proposed for the new north–south and SE– NW runways); holes 87–196 cover the southern area (and so complete the areas for the NE–SW and east–west runways).

 Geologisches Gutachten über den Untergrund der Landebahnen des Flughafens-Guernsey. Der Geologe beim Lw. Feldbauamt 14 (K.-I.) im Luftgau Westfrankreich. Br.B.Nr. 17/43 geh. O.U. Jan. 1943. Sachbearbeiter: Reg. Bauassessor d. B. Dr. Schneider. [2 pp., 1 table, 7 maps plus key on separate sheet, 196 borehole logs.] 18

7  Guernsey and the German Air Force

261

Fig. 7.2  Map of Guernsey, showing position of the airfield

All of the borehole logs are included as appendices to the report. From these, it is clear that the first log (for borehole 3) was completed on or about 29 May 1942, and the second log (for borehole 2) on or about 30 May (the dates of Schneider’s endorsing signature). Logs from 1 to 13 June inclusive bear the signatures of both the borer (Bohrmeister Robert Penz) and the geologist (Schneider), signed on the same dates and so indicating that the two worked closely together at this time (e.g. Fig. 7.4). Logs for 16 and 19 June bear only Schneider’s signature: the space for borer’s signature is left blank. Logs from 20 June to 20 July are signed by the borer for the date indicated, but all these logs were countersigned by Schneider only on 9, 10, 12, or 14 October—presumably indicating that although boring continued daily or near daily from 20 June to 20 July, Schneider was otherwise involved (and possibly absent from Guernsey) until at least 9 October. From 20 July there is then a gap in borehole logs for 2 months, until 20 September, when 21 logs are dated by the borer (Penz) over 5 days, but not countersigned by Schneider until 14 and 15 October. Since boreholes were being logged and apparently constructed at the average rate of about four per day in June and July, this gap in drilling seems to mark a real cessation of activity at this time. There is then a further gap until 10 November, when 26 logs were signed by the borer (and countersigned by Schneider on 28 November), followed by 15 more logs

262

E. P. F. Rose

Fig. 7.3 Hand-operated boring rig of the type operated by or for German armed forces during World War II. Winde = winch; Dreibock = tripod; Krückel = tiller (for rotating the drilling rod); Rohrbündel = drill guide; Lehm = loam; Verrohrung = casing; Sand und Kies = sand and gravel; Drehgestänge = drilling rod; Schappe = auger; Ton = clay. From Geologen-Gruppe (1918)

on 20 November (countersigned by Schneider on 26 November). After a further gap, the final 19 logs bear the borer’s date and signature of 15 December, but Schneider’s endorsement dated 15 January 1943. In these cases it seems that gaps in completion of the paperwork (i.e. logging of the borehole samples) are more evident than gaps in the programme of boring as such. Certainly it seems that a programme of boring continued from the end of May through June and July to a later date (perhaps as late as mid-December since Robert Penz was still available to sign borehole

7  Guernsey and the German Air Force

263

Table 7.1  Dates of completion by the borer of 196 borehole logs (cf. Fig. 7.4) for the Guernsey airport region (Figs. 7.5 and 7.8) in 1942 Date 1st 2nd 3rd

May – − −

4th 5th 6th

− − −

June 1, 4 − 5–7, 9, 10, 11, 16 − 17, 21, 22 −

7th





8th



9th



12, 15, 18, 20 −

10th −

8, 13, 14, 19

11th − 12th − 13th −

− − 23, 24, 36, 37

14th − 15th −

− −

16th − 17th −

30, 33–35, 38–40 −

18th −



19th −

25–29, 31, 32, 41, 42 63, 64, 70, 71

20th −

July − − 50–52

August − − −

September − − −

October − − −

November − − −

December − − −

87 − 48, 49, 53 46, 47, 88, 89, 138 43, 44, 45 122, 146– 149 144, 145

− − −

− − −

− − −

− − −

− − −







































− − 110, 111, 124 − 115, 119

− − −

− − −

− − −

91–98, 112, 117, 123, 133, 134, 155–161, 165–170 − − −

− −

− −

− −

− −

118, 143









− 150–154, 162–164, 186–196 −

116, 120, 141, 142 113, 114, 121, 140 −































137, 139



126–128, 135, 136



171–185



− − −

(continued)

264

E. P. F. Rose

Table 7.1 (continued) Date May June 21st − − 22nd 23rd 24th 25th

− − − −

26th 27th 28th 29th 30th 31st

− − − 3 2 −

60–62 58, 59 55–57 54, 84, 85, 86 82, 83 77–81 72–74 65, 75, 76 66–69 xxxxxxx

July −

October November − −

December −

− − − −

August September − 107–109, 125 − 103–106 − 99–102 − 129–132 − −

− − − −

− − − −

− − − −

− − − − − −

− − − − − −

− − − − − −

− − − − − xxxxxxx

− − − − − −

− − − − − xxxxxxx

logs at that time, although that would imply that boring became episodic or much slower than previously from 20 July); that boreholes were completed by area to pre-­ determined grids rather than in strict numerical order; and that Schneider was directly involved in this project at its outset in May–June, again in mid-October, and in finally bringing the data together to generate his ‘opinion’ in January 1943. The ‘opinion’ is illustrated by seven maps derived from the logs (e.g. Figs. 7.5 and 7.8) that present the key results of the survey. All at scale of 1:2500, these are ground profile maps that depict ‘soil’ conditions at depths of 0.5, 1.0, 1.5, 2.0, 3.0, 4.0, and 5.00 m below the land surface. Each map depicts seven categories of ‘soil’ (Figs. 7.6 and 7.7), to indicate the load-bearing capacity of the ground and its relative permeability. The maps demonstrate that loess or a soil derived from loess covers nearly all of the airfield area, underlain by weathered rock to a depth of at least 5 m, and that unweathered bedrock gneiss (see Figs. 6.14 and 6.15) which occupies most of southern Guernsey must therefore lie at greater depth. The text is only two pages of typescript, but it is accompanied by a table that also distinguishes seven soil categories—with ‘Belastbarkeit’ (strength, i.e. load-­bearing capacity) values for four conditions (ranging from dry to wet) in each case—and repeats the assessment of relative permeability (to rain or groundwater) of each soil category in five terms from ‘good’ to ‘very bad’. Since the ‘opinion’ was not completed until 9 months after the start of the borehole programme, the project does not seem to have been given high priority. Apparent gaps in the time sequence of borehole emplacement presumably indicate either that at times the equipment (or Dr. Schneider) was required for higher priority use elsewhere, or that at these times the runway was actually required for use by aircraft, so could not be impeded by the boring apparatus and its operators. It seems unlikely that the simple and robust auger would have sustained damage causing it to be taken out of service for repair.

7  Guernsey and the German Air Force

265

Fig. 7.4  Number one of the 196 borehole logs on which the maps depicted as Figs. 7.5 and 7.8 and their five companion maps were based. Borehole site is the extreme SE corner of Fig. 7.5. The columns show successively: borehole depth (here to 7.70 m), soil description, geological identification, groundwater occurrence, bearing capacity (in kg/cm2), and relative permeability. The log is signed and dated by both the borer and the geologist and stamped in red ‘Geheim’ (secret) as well as with a standard reference block for appendix and report numbers assigned by the ‘L[uftwaffen] F[eld] B[au] Amt Kanalinseln’. From Rose and Willig (2012); reproduced courtesy of the Bundeswehr Geoinformation Centre

266

E. P. F. Rose

Fig. 7.5  Map, original at scale of 1:2500, showing ‘soil’ conditions beneath the Guernsey airfield (cf. Fig. 7.2) at depth of 0.5 m—the first of seven maps compiled by Dr. Schneider in January 1943 for depths ranging from 0.5 to 5.0  m. For key see Fig.  7.6. Boreholes (cf. Fig.  7.4) were sited according to three primary overlapping grid systems, as described in the text, the first beginning at the SE corner and extending NNW; the second covering the remainder of the eastern area; and the last east–west across the southern region. Boreholes on the first grid detected mainly ‘loess-loam’ (coloured pale brown on the map) but also an area of made ground (coloured blue); on the second grid they indicated that much more of the northeastern area was underlain by ‘loess-loam’; and on the third grid that most of the southern area was underlain immediately by true ‘loess’ (yellow on map)—a sediment of predominantly silt-sized particles. Two small diagrams near the top right map margin display wind directions recorded for the years 1938–1940 (top) and for the poor weather conditions of November–March only (bottom). From Rose and Willig (2012); reproduced courtesy of the Bundeswehr Geoinformation Centre

7.3.2  Ground Conditions and Airfield Building Construction The opening paragraph of this ‘opinion’19 again defines the nature of the task. It was planned to construct large hangers to field-type fortification standards, so as to accommodate fighter aircraft on land to the SW of the three proposed runways, together with a large number of substantial bunkers for a variety of purposes. The

 Gutachten über die Baugrundverhältnisse im festungsmässigen Bausektor für die Jagdgruppe (E-Hafen Guernsey). Der Geologe beim Lw.-Feldbauamt 14, Kanalinseln, Br. B.  Nr. 270/43  g. Kdos. Az. 63 c 26. O.U. 5 Juni 1943. Sachbearbeiter: Reg. Bauassessor d.B. d. Lw. Dr. Schneider [2 pp., 8 figs.] 19

7  Guernsey and the German Air Force

267

Fig. 7.6  Key to map depicted as Fig.  7.5. Seven colours are used to distinguish soil types (Lösslehm  =  loess-loam, Löss  =  loess, Verwitterung[s]lehm  =  weathered loam, Verwitterungston  =  weathered clay, Verwitterungsgrus  =  weathered sandy detritus, Brockenzone = fragmented zone, künstliche Aufschüttung = made [or worked] ground), with classification in engineering terms (leichter = weak, mittlerer = moderate, fester = strong) and relative permeability (ausreichend = adequate, gut = good, schlecht = bad, sehr schlecht = very bad, and mittelmässig = average). From Rose and Willig (2012); reproduced courtesy of the Bundeswehr Geoinformation Centre

‘opinion’ was therefore required to define ground conditions in order to guide definitive planning of the construction work. Very similar in style to the previous ‘opinion’, this also consists of a brief (two-­ page) text accompanied by seven maps at scale of 1:2500 to show ground conditions at various depths, based on a series of borehole logs. The covering letter (dated 5 June 1943) from the leader of Field Works Office 14 states that 49 borehole logs (numbered 197–245) accompany some copies of the report, but neither of the first two—one of which the copy seen must therefore be. However, unlike the previous report, a map (at scale of 1:2500) indicating depth to groundwater is also included (Fig. 7.9). The means of investigation as described in the text is very similar to that for the proposed runways. Since the precise location for the individual buildings had not been finalized, even for the large hangers, the region in general was surveyed by means of boreholes sited according to a 100-m grid (Fig. 7.9). The borehole numbers clearly follow on precisely from those put down for the airfield report completed in January, so presumably the site investigation for this

268

E. P. F. Rose

Fig. 7.7  Key to map depicted as Fig. 7.8. Categories as for Fig. 7.6, but colours slightly different. There is no loess-loam (i.e. top soil derived from loess) or worked ground at this depth. From Rose and Willig (2012); reproduced courtesy of the Bundeswehr Geoinformation Centre

project followed on directly—although in the absence of dated logs, the exact timing of this investigation cannot be proved from the ‘opinion’ itself. The date of the ‘opinion’ (5 June 1943) allows ample time for the boring of the 49 holes following completion of the airfield report in January 1943—for since the ground profile maps in the two ‘opinions’ are so similar in style, it may reasonably be inferred that boreholes were put down to a similar depth and generated similar log information. The ground condition maps are less elegantly coloured than those for the airfield, but the key distinguishes the same seven soil categories and tabulates the same equivalent information (for ground strength and relative permeability). Maps show ground conditions at depths of 0.5, 1.0, 1,5, 2.0, 3.0, 4.0, and 5.0 m below surface level, as for the airfield maps compiled previously. The ‘groundwater’ map, unique to this ‘opinion’, shows merely the depth to water table at those borehole locations (Fig. 7.9) where the hole reached 5 m in depth, with no attempt to interpolate contours for water table depth in so small an area. The text of the ‘opinion’ provides a paragraph each of general introduction, the kind of investigation undertaken, soil conditions, groundwater, bearing capacity of the ground, and conclusions. The conclusions specify foundation depths of about 1 m for normal bunkers, but over 6 m for the aircraft hangers, and requirements for drainage during construction because of the height of the water table.

7  Guernsey and the German Air Force

269

Fig. 7.8  Map, original at scale of 1:2500, showing ‘soil’ conditions beneath the Guernsey airfield at depth of 5.0 m—the last of seven maps compiled by Dr. Schneider in January 1943 for depths ranging from 0.5 to 5.0 m. For key see Fig. 7.7. The grid systems used to site boreholes for the four potential runways are clearer on this map than in Fig. 7.5. Ground strength has apparently increased with depth, unsurprisingly. From Rose and Willig (2012); reproduced courtesy of the Bundeswehr Geoinformation Centre

7.3.3  G  round Conditions and Water Supply at Anti-Aircraft Battery Sites The text of this third ‘opinion’20 is again brief (less than two pages of typescript), but accompanied by a map of the whole island showing sites for six heavy anti-­ aircraft gun batteries and their headquarters, all with their means of water supply, then detailed plans for these nine units, each followed by logs for associated boreholes. The map (Fig. 7.10, cf. Fig. 7.11) shows the brigade headquarters (HQ) (symbol: red spot surmounted by red triangle with horizontal line) near St. Martin in SE Guernsey, with a regimental HQ (red spot surmounted by square black flag) close by; a battalion HQ at Haut Landes, on the edge of St. Sampson some 3.5 km almost

 Geologisches Gutachten über den Baugrund und die Wasserversorgung der schweren Flakbatterien auf der Insel Guernsey. Der Geologe beim Lw. Feldbauamt 14 im Luftgau Westfrankreich. Az. 63 c 26. Br.B.Nr. 284/43  g.Kdos. O.U., den 27.7.43. Sachbearbeiter: Reg. Bauassessor d.B. Dr. Schneider. [2 pp., annex 2 pp., 15 figs.] 20

270

E. P. F. Rose

Fig. 7.9  Map, original at scale of 1:2500, compiled by Dr. Schneider in 1943, showing grid system for numbered boreholes plus depth in metres to water table (figures in blue) recorded in the deepest (c. 5  m) boreholes, for a potentially fortified region bordering the SW corner of the Guernsey airfield (cf. Fig.  7.5). From Rose and Willig (2012); reproduced courtesy of the Bundeswehr Geoinformation Centre

due north of the brigade HQ (red spot surmounted by black triangle); and six battery positions (red spot surmounted by gun symbol). The batteries are based at St. Sampson (immediately east of the regimental HQ), L’Ancresse, and St. Germain in the NE, and at Mont Saint, La Corbière, and Les Huriaux in the west and south. Each of the nine sites is illustrated by a plan at scale of 1:2500 (e.g. Fig. 7.12), showing the positions of boreholes and bunkers and, for battery positions, gun sites. The battery at L’Ancresse is additionally illustrated by a plan at scale of 1:500, credited to an Air Force Fortification Staff.21 Each of these ten plans is followed by borehole data: a single log for the borehole on the site of the Brigade HQ command bunker; two borings for positions of the St. Martin regimental HQ command bunker, and one for that at Hautes Landes; and eight borings for each of the six battery positions, two for command bunkers (e.g. BI and BII in Fig. 7.12) and six for gun positions (e.g. GI to GVI in Fig. 7.12). The logs for battery positions summarize 15 categories of information for each borehole: 1. ground conditions (log of principal soil/rock units to depth of about 5.0 m), 2. relative workability (for each principal soil/rock unit), 3. bearing capacity in kg/cm2, 4. slope stability, 21

 Luftwaffen Befest. Stab 7/XII.

7  Guernsey and the German Air Force

271

Fig. 7.10  Map of Guernsey, original at scale of 1:25,000, showing sites of six heavy anti-aircraft batteries plus a brigade, regimental, and battalion headquarters, as described in the text, with means of water supply, compiled by Dr. Schneider in 1943 (cf. Fig. 7.11). The key in the bottom right corner of the map shows from the top down sites for (1) Notversorgung = emergency supplies: (a) well with an existing pump (green spot), (b) potential well, up to 8.0 m deep (brown spot), (c) potential well, over 8.0  m deep (yellow spot), and for (2) festungsmässig  =  ‘fortress standard’ supplies: (d) captured spring (blue spot), (e) well (small red spot). These sites are legibly numbered 1–28 on the original map, and details summarized in an accompanying table (Table 7.2). From Rose and Willig (2012); reproduced courtesy of the Bundeswehr Geoinformation Centre

5. fragmentation, 6. ground permeability, 7. depth to groundwater table, 8. potential for abstraction galleries, 9. use as conduit for sewage, 10. suitability for latrines and WCs, 11. groundwater flow direction, 12. water supply, 13. groundwater quality, 14. high water risk, and 15. other construction materials.

272

E. P. F. Rose

Fig. 7.11  Map of Guernsey, showing positions of the airfield and six heavy (88 mm calibre) anti-­ aircraft gun batteries plus brigade, regimental, and light anti-aircraft (+) battalion headquarters as at July 1943, as in Fig. 7.10. Amended from Rose and Willig (2012)

An adequate, secure water supply was evidently important for each of the nine sites. Table 7.2 is a translation of the detailed information summarized in the ‘opinion’. The text of the ‘opinion’ summarizes information under two headings: building foundations and water supply. It makes clear that the building proposed was for gun emplacements on battery sites and for associated command bunkers, and that soil samples were obtained to 5 m depth during borehole construction by means of a shell-and-auger rig. The ground was assessed as suitable for the proposed construction everywhere except at the site in Lancresse Bay, in the far north of the island, although groundwater was likely to cause rapid corrosion of reinforced concrete. In Lancresse Bay, since ground conditions at the original site proved unfavourable, Schneider’s survey determined a new and more suitable location. For water supply to all sites, ‘field’ conditions could be met by joining the island’s piped system; for ‘emergency’ conditions, it was recommended that the nearest wells be all fitted with appropriate pumps; and to meet more substantial ‘fortress’ conditions, new wells would have to be drilled and springs captured as shown on the plans. The site in Lancresse Bay was again deemed exceptional: it required a surface water reservoir, since any groundwater in the region was likely to prove brackish.

7  Guernsey and the German Air Force

273

Table 7.2  Translation of table listing 28 sites to supply water to heavy anti-aircraft artillery units on Guernsey, compiled by Dr. Schneider in 1943 to accompany the map illustrated as Fig. 7.10 No. Type, place 1 Large well shaft west of Doyle Rock 2 Well shaft west of La Greve 3 Well shaft by Capelle 4 Well to be newly installed 5 Well to be newly installed 6 Well shaft by Blanches Rocques 7 Well shaft north of Kings Mills 8 Spring

9 10 11 12 13 14 15 16 17 18

19

20 21 22

Well shaft east of L. Marais Well shaft east of Mont Saint Masonry well shaft east of L. Marais Masonry well shaft east of L. Marais Well shaft west of Choffine Farm Captured spring

Depth to water 10.2

Hole depth 16.65

Equipment Pump

2.5

7.0

Wind pump

3.7 –

Wind pump –

8.8

14.0 10– 14.0 10– 14.0 21.5

10.8

15.5





– Wind pump

Some old fixtures are to hand. Valley slope is boggy, with discharge of medium strength. Improvement is possible through installation of an abstraction gallery 5.0 10.5 Hand pump ?

14.95

Wind pump

10.0

14.0



5.6

12.45



4.3

16.35

Motor and wind pump

Spring can be captured by an abstraction gallery to the south of the valley in the depression of the southern slope Well shaft 0.8 5.85 – Well shaft ? 8–10.0 – Well shaft south of Les 2.1 7.4 Wind pump Murches Existing water source Utilize by abstraction galleries in wet pasture land west of La Corbiere Well shaft SW of Les 4.05 16.1 Motor and wind pump Huriaux to be built up to fortress standard Well shaft SW of Les 6.45 8.65 – Huriaux Well shaft north of Les 8.0 11.7 Hand pump in house Huriaux Well shaft in La Fosse 9.5 14.9 Electric pump (continued)

274

E. P. F. Rose

Table 7.2 (continued) No. Type, place 23 Captured spring in La Fosse 24 Captured spring in La Fosse 25 Well shaft in St. Martin 26 Well shaft by Village de Puton 27 Well shaft NW of Calais 28 Spring NW of Calais

Hole Depth to water depth New installation, moderate yield

Equipment

New installation, potential for limited improvement 10.8

12.55

8.3

10.55

?

10.00

Motor pump very productive Motor pump

Hand pump very productive New installation, potential abstraction gallery possible

Depths in metres

Gavey (2001) has recorded that only two of the six batteries were built to ‘fortress’ standard, but that each was supplemented by three lighter (20 mm calibre) anti-aircraft guns and by machine guns. Each heavy battery was also provided with a Wurzburg Dora Radar, to supplement the main Luftwaffe radar station (named Alderschloss) at Fort George which mounted two Giant Wurzburg and two Freya radar units. Gavey lists these batteries as: 1. Flak Batterie Dolman, at L’Ancress Bay. This was built entirely to ‘fortress’ specification, and the guns could be depressed to serve additionally in a coast defence role. The battery positions survive intact, with the command bunker and radar position also still visible, within part of a present-day golf course. 2. Flak Batterie Gabelsberg, at St. Germain, above Vazon Bay (cf. Fig. 6.11). This was built only partially to ‘fortress’ specification, and now lies on private property. 3. Flak Batterie Kapellendorf, at Capelles, St. Sampson. This consisted of ‘field order’ emplacements, with little building work carried out. There are no visible relics. 4. Flak Batterie Heilingenberg, at Le Mont Saint, St. Saviour. Another location with only ‘field order’ emplacements, and no visible relics. 5. Flak Batterie Les Huriaux, at St. Martin. Here six guns were mounted in ‘field positions’, and a seventh was retained as a mobile piece. There are no visible relics. 6. Flak Batterie Rabenstein, at Les Laurens, Torteval. The second battery to be built to ‘fortress’ standard. All concrete installations survive, but are on private property.

7  Guernsey and the German Air Force

275

Fig. 7.12  Site plan for battery of heavy anti-aircraft guns to be positioned at Mont Saint on Guernsey (large red spot to NE in Fig. 7.10; cf. also Fig. 7.11), original at scale of 1:2500, compiled by Dr. Schneider in 1943. One of a set of plans identical in style for nine unit positions, as described in the text. From Rose and Willig (2012); reproduced courtesy of the Bundeswehr Geoinformation Centre

276

E. P. F. Rose

7.4  K.G. Schmidt on Guernsey The preface to a geological report on reconnaissance for tunnel systems on the Channel Islands as a whole22 (described in Sect. 5.3) states that its author, Professor K.G. Schmidt (described in Sect. 5.2), an official from the administrative centre23 of the Luftwaffe’s Air District Command24 for western France,25 visited sites for tunnels on Guernsey between 19 and 21 April 1942, accompanied by Reg. Baurat Theis, Reg. Bauass. a. K. Dr. Schneider, and Uffz. Dipl. Berging Schulte. Additionally, Schmidt visited two Guernsey tunnel sites on 3 May, and an Army tunnel system there whose visit was led by an engineer, Ing. Meckel. Schmidt was not himself seemingly to generate any geological reports specifically for Guernsey. However, it is evident that conditions in the tunnels then under construction in Guernsey as well as Jersey and Alderney influenced the guidance he provided in his report. Moreover, in the rank of Regierungsbaurat (equivalent to an Air Force major) he was senior to Schneider as a Regierungsbauassessor (equivalent to a captain). Schmidt presumably gave verbal instructions that would also have influenced Schneider’s work as a geologist on Guernsey in the subsequent months. It was presumably also because the Luftwaffe already had a resident geologist in Guernsey, Hans Schneider, that Schmidt himself was to compile geological reports only for Jersey (Sect. 5.5) and Alderney (Chap. 8), where the Luftwaffe lacked such expertise.

7.5  Franz Schulte on Guernsey By October 1943, the ‘Uffz. Dipl. Berging Schulte’ who had accompanied both K.G. Schmidt and Hans Schneider on their tunnel visit on Guernsey in April 1942 (see Sect. 5.6) had been promoted to succeed Schneider as the Luftwaffe’s geologist resident on Guernsey (Rose and Willig 2014). Evidence for this comes from an ‘expert opinion’ (Gutachten) dated 11 October 1943.26 This ‘opinion’ was written as

 Geologischer Bericht Nr. 146. Grundsätze zur Geologischen Vorerkundung von Hohlgangsystemen auf dem Kanalinseln. Berichterstatter: Reg. Baurat Prof. Dr. K.G.  Schmidt, Dipl. Berging. Luftgaukommando Westfrankreich—Verwaltung—Az.: 63 c 26 A 87—Verw. III/7—Br. B.  Nr. 5122/42 geh. O.U., den 11.5.1942 [Bundesarchiv-Militärarchiv file RH/3041]. 23  Verwaltung. 24  Luftgaukommando. 25  Westfrankreich. 26  Geologisches Gutachten für das unterirdische Marine-Torpedolager auf der Kanalinsel Guernsey. Bezirksgeologe Kanalinseln b. d. Oberbauleitung b. General der Luftw. Kanalinseln (Guernsey). Beantragende Dienststelle: Marine Baudienststelle Guernsey. Bearbeiter: Regierungsbauassessor F.  Schulte, Dipl.-Berging. O.U., 11 Oct 1943. [4  pp., 18 figs; now preserved in the Bundesarchiv-Miliärarchive, Freiburg, in file RH32/3041.] 22

7  Guernsey and the German Air Force

277

from the District Geologist27 for the Channel Islands28 within the Senior Works Management29 of the General commanding the Air Force on the Channel Islands, from a base in Guernsey. The office authorizing preparation30 of the ‘opinion’ was the Naval Works Office31 in Guernsey. The author32 was F.  Schulte, writing ‘on deployment’.33 The ‘opinion’ credits Schulte with his academic qualification as a mining engineer (Dipl.-Berging) as previously, and the status of Regierungsbauassessor that he had achieved by 9 April 1943 (Sect. 5.7), but he is recorded in the new appointment of ‘District Geologist’. This followed a reorganization of the Luftwaffe’s administrative structure on the Channel Islands, effected after June (the date of a covering letter noted in Sect. 5.7). The ‘opinion’ is classified as ‘secret’ (Geheim!) and stamped as annex (Anlage) 1 to a report34 of the Luftwaffe’s Feldbauamt 1/3—a reorganization of the Feldbauamt 14 that commissioned the report of 9 April. It is not stated on the report itself how many copies were made, or where they were sent, but presumably recipients included both the Naval Works Office on Guernsey (the top copy, to the commissioning office, as standard practice—believed to have been destroyed prior to German surrender) and the Army’s Inspectorate of Land Fortification in Berlin (the carbon copy now found in the Bundesarchiv-Militärarchiv, together with other Inspectorate documents).

7.5.1  Recommendations in Schulte’s ‘Opinion’ The ‘opinion’ opens with a statement of the requirement. A tunnel for the storage of Naval torpedoes is to be driven for about 100 m, with a diameter of 4.6 m; space is to be provided for a workshop, plus fan-ventilation and electricity generating systems; and in the workshop or elsewhere, it must be possible to rotate the torpedoes about their long axis. To this end, two areas were subject to site investigation and geological study, both close to St. Peter Port, Guernsey’s main town, on its eastern coast (Fig. 7.13). Both areas were near the coast, providing easy access to the sea (Fig. 7.14). Like other German military installations in this general region, the overall purpose was to enhance defence against amphibious attack.

 Bezirksgeologe.  Kanalinseln. 29  Oberbauleitung. 30  Beantragende Dienststelle. 31  Marine Baudienststelle. 32  Bearbeiter. 33  O.U. = Ortsunterkunft: location undefined for military postal service. 34  Br. Tgb. Nr. 129/43 geh. of the Oberbauleitung. 27 28

278

E. P. F. Rose

Fig. 7.13  Map of Guernsey, showing positions of the German underground facilities planned or under construction as at 29 June 1943. From Rose and Willig (2014), after Ginns (1993), reproduced courtesy of the Channel Islands Occupation Society (Guernsey)

Area 1 lay to the south of Havelet Bay: part of the plateau fringed by steep slopes between the promontory of ‘Les Terres Point’ which juts into the sea and a bathing place to the north. The area available here was very restricted: tunnels and/or tunnel complexes were proposed or already under construction for the German Army, Navy, or Air Force, leaving only a small space in which access could be created for the proposed torpedo store. Development was constrained to a westerly axis. Nevertheless, two possible routes, 1a (Fig. 7.15) and 1b (Fig. 7.16), were evaluated for a tunnel in this area. An advantage noted for this region was that it was to some extent sheltered and so protected by the ‘Terres Point’ promontory to the south. Area 2 lay further to the north, much closer to the harbour, to the SE of ‘Castle Pier’ and parallel to the cliff inland from ‘South Beach’ (Fig. 7.14). The surface of this area was already mostly overbuilt, occupied by the houses and gardens that formed the southern district of St. Peter Port (Fig. 7.17). Although the potential tunnel access was here more open to view from the sea than in area 1, the ‘opinion’ noted that the tunnel mouth and a route across it could be concealed by a dense line of trees, planted across a small parkland area.

7  Guernsey and the German Air Force

279

Fig. 7.14  German map showing the potential locations for an underground torpedo store south of St. Peter Port; original at scale of 1:2500. Annex 2 (Anlage 2) to a Geological ‘Expert Opinion’ by the Channel Islands Regional Geologist of the Luftwaffe’s Senior Works Management, Guernsey. The map shows, from north to south, location 2 (Fig. 7.17) inland from South Beach, at the western edge of Havelet Bay; overlapping locations 1a (Fig. 7.15) and 1b (Fig. 7.16) inland to the ENE of the promontory Les Terres Point; and sites numbered 1–7 at which orientations of planar features in the rocks were measured. From Rose and Willig (2014); reproduced by permission of the Bundesarchiv-Militärarchiv from file RH32/3041

Geologically, according to the ‘opinion’, both areas 1 and 2 lie at the NE corner of the vast mass of granite-gneiss known to form most of southern Guernsey. Although not cited in the ‘opinion’, a military geological map for Guernsey at a scale of 1:25,000 had been remotely compiled from published sources by the Army geologist Walter Wetzel in November 1941 (Fig. 6.2), and maps at this scale for the bedrock, bedrock plus superficial geology, and military geology were all compiled during 1942 on the basis of field survey by members of Wehrgeologenstelle 4 (Figs. 6.14, 6.15, and 6.16). Presumably Schulte had access to at least one of these maps when generating his ‘opinion’. However, he was clearly himself a competent geologist. His ‘opinion’ distinguishes and describes two types of gneiss: one coarse-grained, the other fine-­ grained, with an abrupt boundary between the two. The coarse-grained rock is recognized as an augen-gneiss, its dominant minerals being quartz, feldspar, and mica, in which the quartz and feldspar crystals are largest, especially the latter—and with the ‘augen’ (eye) structure typical of this well-known rock type. The fine-­

280

E. P. F. Rose

Fig. 7.15  German geological map and potential tunnel plan for location 1a of Fig. 7.14; original at scale of 1:500. Annex 4 to the Geological ‘Expert Opinion’ by Regional Geologist F. Schulte. The boundary of a lined area in the NW corner is labelled as that of the Air Force operational fuel store tunnel system (Hohlgangsystem Betriebsstofflager Luftwaffe), and the lined area at the top to the east of this the location of a fuel store for the German Navy (Prov. Hafen Betriebsstofflager Marine). The tunnel plan indicates, amongst other features, positions for a turntable for rotating the torpedoes (Wendeplatte), their storage room (Torpedolagerraum), electricity generating room (Elektrische Anlage), fan-ventilation system (Ventilator), and a workshop (Werkstatt). Nine geological features are explained by a key at the bottom, from left to right: (1) hard, coarse-grained gneiss (harter, grobkörnig Gneis); (2) crumbly, weathered gneiss (mürber, verwitterter Gneis); (3) fine-grained, foliated gneiss (feinkörnig flasriger Gneis); (4) hornblende-rich gneiss (hornblendereicher Gneis); (5) fault (Störung); (6) dolerite dyke (Diabasgang); (7) boundary line marking limit of underground construction (Begrenzungslinie ansgrenzender Hohlenbauten); (8) rock changes all vertical (Gesteinswechsel alle seiger); (9) strengthening rods required for 15 m from tunnel boundary (to withstand shock) (Sicherheitspfeiler zu ansgrenzendem Stollen von 15.00 m (ab Stoss gemessen)). From Rose and Willig (2014); reproduced by permission of the Bundesarchiv-­ Militärarchiv from file RH32/3041

grained rock is noted for its more obvious foliation: thin parallel bands rich in quartz and feldspar alternating with darker, more micaceous, layers. Schulte describes grain size variations in the rock, and variations in its general reddish colour, noting that overall the bands in the rock dip at about 50–60° to the NE, with strike direction therefore NW–SE. Quartz is a mineral much harder than mica, so Schulte predicted that drilling in the quartz-rich augen-gneiss would be more difficult than in the mica-rich foliated gneiss. In either case, during rock drilling there would be significant wear on the drill bits.

7  Guernsey and the German Air Force

281

Fig. 7.16  German geological map and potential tunnel plan for location 1b of Fig. 7.14; original at scale of 1:500. Annex 3 to the Geological ‘Expert Opinion’ by Regional Geologist F. Schulte. Labelling and key as for Fig. 7.15. From Rose and Willig (2014); reproduced by permission of the Bundesarchiv-Militärarchiv from file RH32/3041

Schulte recognized that the gneiss was cut by intrusions of a distinctive dark grey rock, of which the minerals hornblende, mica, and feldspar were the most common constituents. The intrusions were thin, varying from NE to NW in their strike, and steeply dipping. Their contact zones with the gneiss were rich in quartz and so ‘hard’. Indeed, the rock overall had great hardness: it showed little evidence of jointing. Veins of aplite also cross-cut the gneiss, but with different strike. Being very thin (50–250 mm), these were deemed to be insignificant with regard to tunnelling activities. Schulte also detected that the region was cross-cut by faults, with a high angle of dip, generally with strike ranging between NW–SE and north–south. Moreover, a zone of disturbance, first detected during excavation for an Army Munition Depot to the south of the area in March 1942, crossed the area, seemingly increasing in thickness towards the NW. He analysed the pattern of rock fracture in the general region by measuring joint directions in the rock at seven locations, plotting a rose diagram for each location to reveal the greatest frequency of joint direction, and a corresponding pole diagram for dip direction, following the procedures illustrated in Sect. 5.5. Although high values for joint frequency directions did not correspond exactly between the seven localities, there was a general tendency for them to show a north–south trend.

282

E. P. F. Rose

Fig. 7.17  German geological map and potential tunnel plan for location 2 of Fig. 7.14; original at scale of 1:1000. Annex 5 to the Geological ‘Expert Opinion’ by Regional Geologist F. Schulte. The rooms for the fan-ventilation system and the workshop, together with turntable for rotating torpedoes, are situated near the tunnel entrance, screened by a line of trees. Overall, the rock is labelled as granite-gneiss. The geological key adds only: A—Aplitgang and D—Diabasgang, for veins of aplite and intrusions of dolerite, respectively. From Rose and Willig (2014); reproduced by permission of the Bundesarchiv-Militärarchiv from file RH32/3041

Schulte noted that the top of the bedrock was badly weathered, typically to a depth of 6–9  m. The weathered zone was capped by a layer of loam35 seldom exceeding 1.5 m in thickness, overlying a very much thicker zone of sandier rock detritus.36 This had been proved in a boring put down to a depth of 17 m in the vicinity of St. Andrews, but it was predicted that there would be occasional pockets of weathering to depths greater than normal in the area as a whole. In some places the rock was so weathered that components had deteriorated to china clay.37 Groundwater, however, did not seem to be a major problem for construction, since the loam layer  Lehm.  Grus. 37  The process known as koalinization. 35 36

7  Guernsey and the German Air Force

283

formed a barrier to downward seepage—and previous tunnel works in the region had not experienced major inflows. For planning purposes, Schulte recommended that a thickness of 10 m should be estimated for the weathered zone; that since the thicker the overhead cover, the stronger the rock forming the tunnel walls and roof, any tunnel should be excavated in the direction that would achieve maximum cover in shortest distance; and that in general, a tunnel should be excavated so as to have a rock cover at least 20 m in thickness. Schulte qualified his description by noting that he was unable to make observations in those parts of the region already overbuilt, or occupied by minefields, and that some of the coastal rocks had been too badly corroded by saltwater for detailed mapping to be possible. Some trial pits to gain access to fresh subsurface rock for examination were therefore recommended. That said, Schulte provided the construction diagrams illustrated (Figs.  7.15, 7.16, and 7.17), in which the axes for the main tunnel, workshop, and fan-­ventilation plus electricity generating systems have been designed to take account of the geological conditions as well as the specified technical requirements for the depot. Tunnel axes are located diagonally across the main joint directions (as recommended in Schmidt’s general guidelines for tunnel excavation, described in Sect. 5.3), and to a large extent take account of the zone of disturbance and the faults observed. In selecting the most favourable direction and profile, Schulte noted that he has made an effort to avoid excessive construction costs! As to the material produced by excavation, Schulte recommended that only material derived from fresh rock should be used as an aggregate for road or other construction purposes. Weathered material, especially that with clay content, was unsuitable for use.

7.5.2  Significance of Schulte’s ‘Opinion’ Schulte’s ‘expert opinion’ on the site for a torpedo store on Guernsey contains more details of local geology than any other Channel Island report by a Luftwaffe geologist yet described (Rose and Willig 2012, 2013, 2014). His account was intended for engineers rather than geologists so omits observations of strictly geological interest, but he clearly recognized the granitic augen-gneiss later mapped by the British Geological Survey (Sect 2.7) under the name of Icart Gneiss and recognized also that this gave way northwards along the coast to a foliated gneiss, presumably the British Geological Survey’s granodioritic Perelle Gneiss or its equivalent. His terminology indicates that he had received a sound training in the principles of geology, enabling him to identify the major rock forming minerals, and processes of igneous intrusion, mylonitization, koalinization, and weathering that had influenced the appearance and strength of the bedrock. His maps (Figs. 7.15, 7.16, and 7.17) are the largest scale German geological maps known for any part of the Channel Islands and reveal his proficiency as a field geologist.

284

E. P. F. Rose

The date of the ‘opinion’ (11 October 1943) is of significance. It comes after Hans Schneider’s last known report (of late July) within the Channel Islands and probably indicates that Schneider had left the Islands by then—although according to Häusler (1995b) he was not transferred from appointment within the Western France region as such until December. Schulte’s move back to Guernsey from tasks on Jersey (Sect. 5.7) seemingly followed reorganization of Luftwaffe construction units on the Islands, and his transfer from Feldbauamt 14 to Feldbauamt 1/3. Seemingly also, he continued there until his next recorded transfer (Häusler 1995b), in March 1944, and if so, served continuously or intermittently for nearly two years in the Channel Islands—longer if he had arrived before the April 1942 site inspection noted above. It was in October 1943 that ‘serious work on the [Channel Island] tunnels came to a virtual standstill when the construction crews were recalled to the Continent to work on the main Atlantic Wall’ (Ginns 1993, p. 3). Details of work completed on Guernsey have been recorded (and extensively illustrated) by Ginns (1993), and by Gavey and Powell (2012). The tunnel designated Ho 3, south of St. Peter Port (Fig. 7.13), which was recorded in 1942 as a ration store, was re-assigned as a Naval Supply Store in 1943, and largely completed. Ho 4 further to the south, although shown even in records between July 1943 and March 1944 as a Luftwaffe Fuel Store, held fuel tanks that, once installed, ‘always contained fuel oil for submarines’ (Ginns 1993, p. 50). It comprised five galleries: four completed and one still under construction in 1944, when work stopped as the Allied liberation of Normandy halted delivery of construction materials. This fuel store and the planned torpedo store are consistent with German plans to establish a naval base at St. Peter Port. Ginns (1987) has noted that ‘During the war, St. Peter Port played host to Schnellboote (motor torpedo boats, known to the British as E boats) of the 5 Schnellbootflottille which was based at Cherbourg [in France]’ (p. 15); that convoys of ships transporting troops, supplies, guns, and construction materials were escorted by armed vessels based in France at St. Malo (p. 24); and that a Harbour Protection Flotilla of small wooden vessels ‘never more than about 30 boats in service at any one time’ (p. 33) was permanently based on the Channel Islands. Ginns (1987, p.  15) has also noted that ‘One of the more unusual operations carried out at St. Peter Port in 1943 was the conversion of nearby Havelet Bay from a rocky barren shore into a fine sandy beach’ by suction dredging of sand from a source offshore, inferring that this might be ‘for some sort of ship repairing operation’. Ho 8, to the south of Ho 4, was designated a fuel store in 1942, but was re-­ assigned as a Temporary Auxiliary Munition Store in 1943. An extension of a nineteenth century British tunnel, this was one of only two Guernsey tunnels to be effectively completed during the occupation. Ho 4 now houses La Valette Underground Military Museum, Ho 8 housed an aquarium and amusement arcade (closed in 2019), so Schulte’s record of the engineering geology of this region with its assessment of factors affecting long-term rock stability should still be of interest. Hitler’s fortification directive of 20 October 1941 (Sect. 3.2) gave the German Navy responsibility for installation of a heavy battery of artillery on Guernsey and

7  Guernsey and the German Air Force

285

for provision of adequate ammunition storage, within an overall Channel Island fortification programme to be integrated by the Army. It is known that the Navy in general had at least one geologist on its staff, to help evaluate beach conditions for sea-borne assault and matters of coastal defence (Häusler 1995a; Häusler and Willig 2000). However, geological guidance to assist construction of the Guernsey heavy battery (‘Nina’, later re-named ‘Mirus’) was provided by the German Army (as described in Chap. 6). It seems, therefore, that inter-service co-operation was routine, and that the Navy made appropriate use of Schulte as a locally based engineering geologist, albeit an official of the Luftwaffe rather than the Navy. After working on the Channel Islands, Schulte was to undertake similar military geological work in Southern Europe (Sect. 5.8) but his career postwar is not known.

7.6  Hans Schneider After Guernsey After completing his assignment to Air District Command Paris,38 Schneider served with the South Italian Air District Command39 from December 1943 to April 1945 (Häusler 1995b, p. 43)—following the Allied landings on the southern Italian mainland that began on 3 September 1943, the surrender of Italy on 8 September, and German attempts to halt the Allied advance northwards. He was thus consistently a Luftwaffe geologist throughout the war. Häusler summarizes Schneider’s work as site investigations for anti-aircraft gun emplacements, and hydrogeological investigations for water supplies for anti-aircraft gun emplacements and for airfields. In the west his work included investigations on the Channel Islands, and hydrogeological studies for V2 rocket bases in the Cherbourg region of Normandy. In Italy his studies included raw materials for construction of airstrips, notably near Florence and Sirmione. According to Häusler (1995b, p. 43), in 1944, as the Allies advanced steadily northward in Italy, Schneider became leader of a military geology team based in Bolzano,40 capital city of the province of South Tyrol in northern Italy, with geological control of tunnelling in the southern front of the Alps. By the winter of 1944 he was assigned to Einsatzgruppe VIII of the paramilitary construction agency Organisation Todt and based in Villach, the seventh largest city in Austria, for further Alpine work—until the end of hostilities in May 1945, and his internment as a prisoner-of-war. A biographical record at the University of Münster (Rose and Willig 2012) shows that following his release, in 1946 Schneider founded a hydrogeological agency in Bielefeld to undertake work for groundwater resource and water supply tasks. He began again to publish articles in this field (e.g. Schneider 1947), and his agency

 Luftgaukommando Paris.  Luftgaukommando Süd-Italien. 40  Bozen. 38 39

286

E. P. F. Rose

tasks took him to many countries in southern Europe and in Africa. By 1948 he was associated with the Hydrogeological Working Group of the Administration for Home Economy of the United Home Regions,41 the forerunner of the Federal German Ministry for Home Economy. He became an early member of the International Association of Hydrogeologists when this was founded in the 1950s. Becoming increasingly distinguished as a hydrogeologist, he lectured at the Westfälischen Wilhelms University of Münster from 1969 and in 1972 was appointed an honorary professor of hydrogeology in its Geological-Palaeontological Institute. A book (Schneider 1981) published in honour of his 65th birthday contained a contemporary photograph (Fig. 7.1) and a list of his 87 publications to that date, mostly dealing with hydrogeology. On 9 November 1990 the University of Münster honoured Schneider by commemorating the ‘golden jubilee’ of his doctoral thesis, published 50 years previously (Schneider 1940c). The jubilee was reported in the local newspaper, the Münsterische Anzeiger, on 10 November. On 15 January 1994 the newspaper carried news of another distinction: award of the Verdienstkreuz am Bande of the Verdienst Order of the Federal Republic of Germany, for his contributions to the science of hydrogeology. Schneider died at Bielefeld on 3 February 1999, at the age of 84. He had by that time written or contributed to nearly 100 books and articles (Anon 1999), including a book on groundwater development (Schneider 1952) that achieved three editions by 1988 and so influenced more than a generation of German hydrogeologists. He was well respected within Germany and especially Westphalia for his hydrogeological achievements. However, his publications were almost always in German, and consequently seem so far to have earned him little international recognition: his work is seldom cited in either the UK or the USA.

7.7  Conclusion It is evident that construction work for the Luftwaffe on the Channel Islands was supported by a geologist resident on Guernsey from at least April 1942 to October 1943, and that this assignment was held initially by Hans Schneider in post as a Regierungsbauassessor, subsequently by Franz Schulte on promotion from the rank of corporal to Regierungsbauassessor (i.e. to rank equivalent to that of a captain). Schneider was a geologist by doctoral degree, and after the war developed a career as a hydrogeologist leading to his appointment as a university professor. Schulte was a mining engineer by profession, but evidently with enough geological training and experience from Germany’s coal mining industry to qualify him for appointment as a Luftwaffe geologist. After the war, he seemingly left no record of

41

 Hydrogeologischen Arbeitskreis der Verwaltung für Wirtschaft des Vereinigten Wirtschaftgebietes.

7  Guernsey and the German Air Force

287

academic publications and is presumed to have returned to employment in the mining industry. Units of the Luftwaffe involved with construction on the Channel Islands came within the administrative district of western France, whose headquarters was based in Paris. This provided a more senior geologist, Professor K.G. Schmidt, in appointment as a Regierungsbaurat and so equivalent in rank to a major, to make a visit to the three largest Channel Islands in April 1942 and issue geological guidelines for future tunnelling work by the Luftwaffe. It seems that further visits were not required. However, the Paris headquarters received copies of the geological Gutachten (‘expert opinions’) generated by Schneider and by Schulte and presumably maintained quality control of their technical activities.

References Anon (1999) Professor Hans Schneider verstorben. GWF Wasser Abwasser 140(4):242 Cruickshank CG (1975) The German occupation of the Channel Islands. Oxford University Press, London Forty G (1999) Channel Islands at war: a German perspective. Allan Publishing, Shepperton Gavey E (2001) A guide to German fortifications in Guernsey. Guernsey Armouries, Guernsey Gavey E, Powell S (2012) German tunnels in Guernsey, Alderney & Sark. Festung Guernsey, Guernsey Geologen-Gruppe (1918) Ergänzungsheft der Wasserversorgungskarte des Gebietes der A.  A. C.  Ver. Abt. 2, Geologen-Gruppe. Archive Document, Heringen Collection, AGeoBw, Euskirchen Ginns M (1987) German harbour organisation and the Harbour Protection Flotilla. Channel Islands Occup Rev 15:5–34 Ginns M (1993) German tunnels in the Channel Islands. Archive book no. 7. Channel Islands Occupation Society, Jersey Häusler H (1995a) Die Wehrgeologie im Rahmen der Deutschen Wehrmacht und Kriegswirtschaft. In: Teil 1: Entwicklung und Organisation, vol 47. Informationen des Militärischen Geo-­ Dienstes, Vienna, pp 1–155 Häusler H (1995b) Die Wehrgeologie im Rahmen der Deutschen Wehrmacht und Kriegswirtschaft. In: Teil 2: Verzeichnis der Wehrgeologen, vol 48. Informationen des Militärischen Geo-­ Dienstes, Vienna, pp 1–119 Häusler H, Willig D (2000) Development of military geology in the German Wehrmacht 1939–45. In: Rose EPF, Nathanail CP (eds) Geology and warfare: examples of the influence of terrain and geologists on military operations. Geological Society, London, pp 141–158 Prien J (1997) Jagdgeschwader 53 history of the PIK-AS Geschwader Vol 1 March 1937 – May 1942. Schiffer Publishing Ltd, Atglen Ramsey WG (1981) The war in the Channel Islands: then and now. Battle of Britain Prints International Limited, London Rose EPF, Willig D (2009) Work by German military geologists on the British Channel Islands during the Second World War. Part 3. Reports with contributions by Walther Klüpfel and Rolf Thienhaus now preserved at the Training and Education Centre of the Bundeswehr Geoinformation Office, Fürstenfeldbruck. Channel Islands Occup Rev 37:105–118 Rose EPF, Willig D (2012) Work by German military geologists on the British Channel Islands during the Second World War. Part 4: site investigation on Guernsey for airfield extension and

288

E. P. F. Rose

heavy anti-aircraft battery locations by Luftwaffe geologist Hans Schneider. Channel Islands Occup Rev 40:188–212 Rose EPF, Willig D (2013) Work by German military geologists on the British Channel Islands during the Second World War. Part 5: work by Luftwaffe geologist Professor K. G. Schmidt and Hilfsgeologe Dr. K. Diebel, for tunnelling (in general and on Jersey) and water supply (on Alderney). Channel Islands Occup Rev 41:78–101 Rose EPF, Willig D (2014) Work by German military geologists on the British Channel Islands during the Second World War. Part 6: work by the Luftwaffe geologist Franz Schulte on Jersey and Guernsey. Channel Islands Occup Rev 42:152–172 Schneider E (1981) Festschrift zum 65. Geburtstag von Professor Dr Hans Schneider. Münstersche Forschungen zur Geologie und Paläontologie 54:1–232 Schneider H (1938) Zur Frage des Münsterländischen Kiessandrückens. Z Dtsch Geol Ges 90(10):603–615 Schneider H (1939) Die wasserwirtschaftlich nutzbaren Grundwasserhorizonte der Münsterschen Bucht. Gas- und Wasserfach 82(49):795–801 Schneider H (1940a) Über die Wasserführung des Schilfsandsteins des mittleren Keupers bei Osnabrück. Gas- und Wasserfach 83(46):576–578 Schneider H (1940b) Erfahrungen beim Bau von Bohrbrunnen in unverfestigten Gesteinen. Pumpen- und Brunnenbau, Bohrtechnik 1940(20):4 pp Schneider H (1940c) Die geohydrologischen Verhältnisse des Gebietes der Baumberge. Decheniana 100A:187–228 Schneider H (1941a) Der geologisch-hydrologische Aufbau der Baumberge. Gas-und Wasserfach 23, 24, 25:342–346, 358–364, 369–374 Schneider H (1941b) Über die Zusammenhänge von Korngrössenzusammensetzung, Durchlässigkeit, Mächtigkeit und Leistung von Grundwasserträgern mit freiem Spiegel in unverfestigten Sedimenten. Gesundheits-Ingenieur 64(5, 6):64–68. 81–84 Schneider H (1941c) Wie weit sind geohydrologische Untersuchungen bei der Anlage von Wasserfassungen förderlich? Pumpen- und Brunnenbau Bohrtechnik 1941(19/20):7 pp Schneider H (1941d) Versickerung und künstliche Grundwasseranreicherung. Gesundheits-­ Ingenieur 64(13, 14):191–196. 205–210 Schneider H (1941e) Ergebnisse zweier Belastungsversuche bei Fliesssand mit Muddeeinlagerungen. Der Bauingenieur XXII 11, 12:95–97 Schneider H (1942) Über die Bohrbarkeit und Wasserführung einiger Gesteinshorizonte in Nordwestfalen. Pumpen- und Brunnenbau, Bohrtechnik 38(25, 26):423–427 Schneider H (1947) Über die Notwendigkeit und Bedeutung regionalhydrogeologischer Untersuchen. DVGW (Deutsche Vereinigung des Gas- und Wasser-faches)-Rundschreiben 9:4 pp Schneider H (1952) Die Wassererschliessung: Grundlagen der Erkundung, Bewirtschaftung und Erschliessung von Grundwasservorkommen in Theorie und Praxis. Vulcan, Essen Schneider H, Wehrli H (1939) Bericht über den Lehrausflug am 28. August 1938 von Münster über Geist, Tilbeck, Billerbeck, Neuenkirchen, Rheine, Salzbergen nach Emsbüren. Decheniana 98A:209–211 Schneider H, Wehrli H (1943) Geologie des Emsbürener Höhenrückens nördlich Rheine. Neues Jahrbuch für Mineralogie, Geologie und Paläontologie 88:263–292 Weale J (2007) Jagdgeschwader 53 ‘Pik As’. Osprey, Oxford

Chapter 8

Alderney Edward P. F. Rose

Abstract  Geological support for the fortification of Alderney came principally from the German Army. During 1941 it came from Second Lieutenant Walther Klüpfel on Jersey, subsequently from the ‘Technical War Administration Officers’ (TKVRs) Walter Wetzel and Friedrich Röhrer in Paris. During 1942 it came first from Lance Corporal (later TKVR) Dieter Hoenes on assignment from the military geological team Wehrgeologenstelle 7, and for the rest of the year from TKVRs Dieter Hoenes and Bernard Beschoren whilst based on Guernsey leading Wehrgeologenstelle 4. Most reports focused on aspects of water supply, and the need to develop groundwater by means of shallow infiltration galleries rather than deep drilled boreholes. Geological studies were complemented by an earth resistivity survey carried out in March–April 1942 by a geophysics reconnaissance unit of the German Army’s military geological organization, led by TKVR Johann Kliemstein (seemingly assisted by TKV Inspektor Woerner): to help determine the depth to groundwater in weathered basement rocks and the extent of a likely freshwater lens above sea water in the coastal zone. Army geologists compiled thematic maps at the scale of 1:10,000, notably two military geological maps to guide extensive fortification in a way comparable with that on Jersey and Guernsey and a raw materials map and report to guide quarrying of bedrock and superficial sands/gravels to support the construction programme. They also compiled reports related to tunnelling for underground facilities, as on the other islands, although in a different geological setting. Additionally, Air Force construction teams at work on sites for anti-aircraft batteries and other facilities had access in 1942 and 1943 to the expertise of a geologist within the Luftwaffe’s field works office for the Channel Islands, based on Guernsey.

E. P. F. Rose (*) Department of Earth Sciences, Royal Holloway, University of London, Egham, Surrey, UK e-mail: [email protected] © Springer Nature Switzerland AG 2020 E. P. F. Rose (ed.), German Military Geology and Fortification of the British Channel Islands During World War II, Advances in Military Geosciences, https://doi.org/10.1007/978-3-319-22768-9_8

289

290

E. P. F. Rose

8.1  Introduction The first German troops arrived on Alderney on 2 July 1940 (Cruickshank 1975; Pantcheff 1981; Ramsey 1981; Forty 1999). At 8 km2 a much smaller island than Jersey or Guernsey, virtually its whole population of about 1500 people had been evacuated to England during June. Alderney was therefore unique in being the only large Channel Island under German occupation without a resident British civilian population. Also, and partly consequentially, it was unique amongst the Islands in that an element of the labour force used for its fortification was drawn from a concentration camp, operated by SS-Baubrigade 1. Until the beginning of 1942, Alderney was administered separately from the other Channel Islands, from Cherbourg in France. Initially too, its garrison was relatively small: rising to only about 450 men by June 1941. However, as with the other islands, the garrison expanded rapidly thereafter. By November 1941 there were nearly 2500 German troops on Alderney: over 1100 from the Army, some 200 from the Navy, and 1100 from the Air Force. The total was soon to exceed 3000. By the end of 1941 the post of Island Commandant was held by an officer of the rank of lieutenant-colonel rather than, as previously, by a much more junior officer: a senior captain. The first of four camps to accommodate a labour force for fortification work was set up by a volunteer group of French workmen in January 1942, initially to accommodate workers of the Organisation Todt. These were to be augmented from July 1942 by ‘forced’ labourers controlled by SS-Baubrigade 1. German forces were to use place names as shown on maps then in use and illustrated in this chapter, rather than in some cases currently or previously (e.g. St. Annes rather than St. Anne for the principal town, and Longy rather than Longis Bay). All the principal landmarks were ultimately given German names.

8.2  Army Geologists in 1941 From August 1941, Walther Klüpfel was responsible for providing military geological advice for Alderney as well as Jersey: work described in Chap. 4 (and by Rose 2005b). In October, Walter Wetzel compiled, from the Senior Fortress Works Office at the Headquarters of the Admiral of France,1 in Paris, a report for Alderney—a month before compiling one for Guernsey (Rose 2005b, and Sect. 6.2.2). Finally, in December, Friedrich Röhrer dealt with Alderney as well as Guernsey (Sect. 6.3.2) in a preliminary report on the supply of drinking water to military sites on headland promontories, based on a reconnaissance visit.

 Oberfestungsbaustab beim Komm. Adm. Frankreich.

1

8 Alderney

291

8.2.1  Walther Klüpfel Although Walther Klüpfel, having recently arrived in Jersey from France, had by August 1941 been made responsible also for provision of advice on the geology of Alderney (as noted in Sect. 6.2), there is no known documentary evidence that any advice was actually required. Since the German garrison initially was far smaller than the civilian population that it replaced, water supply is unlikely to have been a problem. Moreover, since major construction work did not begin until 1942, there was no pressing need either for site investigation or an appraisal of sources of raw materials to facilitate building works. There is no evidence from Klüpfel’s diary that he visited Alderney in 1941, or from other sources that he had a likely need to do so. Nevertheless, at some time Klüpfel made a review of published literature on the geology of Alderney. This facilitated a brief article on its geology (Klüpfel 1944) in a picture book on the island by Otto Bessenrodt. The article takes account, as explicitly stated, of geological literature published between 1811 and 1936, notably that by Ansted, Bigot, Hill, Bonney, Plymen, and Mourant (reviewed in Chap. 2). Articles in the book were originally published in the Deutsche Guernsey-Zeitung, a newspaper for German forces on the Channel Islands, during the summer of 1943. Klüpfel’s article must therefore have been written before this time, but there is no indication in its text of how much before. Presumably writing took place either in 1941 before the geologists of Wehrgeologenstelle 4 became responsible for geological work on Alderney or in 1943 after they had left the Channel Islands at the end of 1942 and Klüpfel was the only Army geologist still in residence (Sect. 6.4.2). Klüpfel’s papers preserved at the British Geological Survey do contain a manuscript geological map of Alderney (Fig. 8.1). This has a German key although no indication of its authorship. At 1:10,000 the map is larger in scale and more detailed than earlier published maps of Alderney (described in Chap. 2). The sandstone at the eastern end of the island is dated as Cambrian (consistent with geological understanding in 1940: Sect. 2.6.4); mapped with a basal conglomerate; and shown with a boundary offset by faulting. A small outlier is mapped to the SW. The central and west igneous region of the island is subdivided into a northern (Château à l’Etoc) granite, central diorite/gabbrodiorite, western granodiorite complex, and southwestern ‘older’ granite. All of these areas, but most especially those of the western island, are shown with minor intrusions, most commonly of quartz porphyry. Whether this map was made by Klüpfel or merely came into his possession is not known. Its base map can be identified from its grid lines as an edition of a 1:10,000-­ scale topographical map that was first issued in August 1941 (Rose 2005a). Preparation of the geological overlay could not have been before this date. Klüpfel’s notebooks and accompanying data sheets show him to have been very fully occupied on Jersey when not on leave on the mainland in 1941 and 1942. The notebooks end on 17 September 1943, and the data sheets give a patchy record of his activities from mid 1943 onwards, so it is possible that he made one or more unrecorded visits to Alderney late in his military career. According to Ginns (1994, p. 126), the pace of construction work on the Channel Islands slackened from April 1943 onward, so

292

E. P. F. Rose

Fig. 8.1  German geological map of Alderney, original at scale of 1:10,000, undated. Key (translated), from the top down: lamprophyre dykes; aplite dykes; diabase dykes; quartz porphyry stocks; granite of Château à l’Etoc; older kataclastic/porphyritic granite; granodiorite, hornblende granite, granite, etc.; diorite, gabbrodiorite; Cambrian sandstone; and Cambrian basal conglomerate. From Rose (2005a); reproduced by kind permission of the British Geological Survey, reference CP19/095

he might more easily have found the time to do so. The map differs in several details from the geological boundaries illustrated on smaller-scale maps published pre-war (as illustrated in Sect. 2.6) and in the detail of numerous minor intrusions, so is evidence of significant geological field mapping made by (or at least available to) German geologists. An additional German geological map of Alderney is preserved at the US National Archives, amongst the maps compiled by Wehrgeologenstelle 4. This is a tracing overlay (Fig. 8.2) that lacks a grid, so the base map used to compile it cannot be identified. The title and scale cartographically inked upon the trace (Alderney 1:10,000) have been amended in handwritten English to read ‘Corrected geol. map of Alderney 1:10,000, 1941’. The cartographically inked legend, however, is entirely in German. The western end of the island is mapped as a region of hornblende granite intruded by diabase, the smaller eastern end as a region of sandstone and quartzite. A small patch of coastal gravel borders part of the igneous region, an even smaller patch of peat is shown upon the sandstone. The map is so similar to the nineteenth-century geological mapping of the island (e.g. Fig. 2.27) that even if based on field survey rather than published sources, it is less detailed than the small-­ scale map already published by Parkinson and Plymen (1929). The author of the trace map is not known, but if the 1941 date is correct, it was compiled at a time

8 Alderney

293

Fig. 8.2  German geological map of Alderney, original at scale of 1:10,000, dated 1941. Key (translated), from the top down: (1) water-bearing valleys; (2) dry valleys (a water supply can be obtained from groundwater at suitable locations); (3) peat; (4) pebbly beach sand; (5) sandstone and quartzite; (6) quarry in sandstone/quartzite; (7) hornblende granite; (8) quarry in granite; (9) diabase (= dolerite); (10) old mine shaft; (11) line indicating best orientation and place for a dam to impound valley water. From Rose (2005a); reproduced courtesy of the US National Archives and Records Administration

when Walther Klüpfel had responsibility for German geological work on Alderney, and before the arrival of any members of Wehrgeologenstelle 4.

8.2.2  Walter Wetzel The Führer Adolf Hitler’s directive of 20 October 1941 (described in Sect. 3.2.3) that stimulated intense planning activity for the Channel Islands was followed instantly by a military geological description of the island of Alderney.2 Also dated 20 October, this was compiled by Walter Wetzel of the German Army’s military geological centre/team Wehrgeologenstelle 9, based at the headquarters of the Admiral of France in Paris (see Sect. 6.2.1).

 Wehrgeologische Beschreibung der Insel Alderney. Sachbearbeiter TKVR Prof. Wetzel. (Oberfestungsbaustab beim Komm. Adm. Frankreich. Wehrgeologenstelle 9.) 20 Oct 1941 [2 pp., eventually 3 maps at 1:10,000—Baustoff- und Minier-Karte, Wasserversorgungskarte, Wehrgeolog. Bodenkarte, with 2 cross sections—but text explains “Anlage: 1 geolog. Karte (Weiterer Karten werden nachgereicht)”]. [Bundesarchiv-Militärarchiv file RH32v.3041, also v. 3043 and v.3082]. 2

294

E. P. F. Rose

The description summarized information under four headings: 1. General rock conditions. This explained that the island is composed of several kinds of ‘hard’ rock, with a thin cover of unconsolidated sediments that becomes patchy in the north. Distribution of the different ‘hard’ rocks present is indicated on an accompanying geological map. A cover of loess-loam reaches 2 m in thickness on the southern part of the plateau that forms most of the island’s surface. Deposits of more varied nature occur within stream valleys. Coastal sands and gravels are significant inland of Saline and Longy bays. 2. Raw materials. This notes that the British had long worked quarries in both granite and diorite for building construction, with more recent quarries in the sandstone. Coastal sands inland of Saline Bay were potentially suitable for building sand. 3. Water supply. Potential for water supply is inferred to be limited. Rainfall mostly percolates underground, and flows below the valleys towards the sea. In some cases this valley water emerges as springs in the lower part of the valleys. These have been tapped in many cases to provide existing supplies. Water is also present in fractured rock. However, the description notes considerable uncertainty with regard to successful boring to abstract this ‘crevice’ water. 4. Fortifications. Mining is considered to be practicable almost everywhere in the various ‘hard’ rocks, progress depending on rock strength. Slit trenches and temporary field positions could be dug without difficulty in the loess-loam cover, especially in the southern half of the island. At one spot an exploratory mineshaft is present, which might be utilized if necessary. The text was accompanied by a ‘military geology ground map’ (Fig. 8.3). Despite its title, the map is clearly merely a coloured and enlarged version of the map of bedrock geology previously published by Parkinson and Plymen (1929) (Fig. 2.31), transcribed on to a 1:10,000 topographical base map, although it bears no such attribution. Its two geological cross sections are identical with those published. The map key is to geological and geomorphological features only. A ‘water-supply map’ (Fig. 8.4) to accompany the report was completed the following month, in November. The key is prefixed by a warning that the island does not possess a water distribution system, and that the wells supplying individual households are mostly of low productivity. A ‘building materials and tunnelling map’ (Fig. 8.5), its scope comparable with the less-accurately entitled ‘military geology’ map of Guernsey (Fig. 6.2), was compiled at about the same time. Some of Alderney’s tunnels were to be started in former quarries as recommended in the report. The ‘military geology ground map’ was used in March–April 1942 as the basis for an earth resistivity survey led by TKVR Kliemstein (Sects. 8.3.2 and 8.3.4): evidence that the report itself was of some practical use in the months following its completion. Its identification of water supply as a potential problem is likely to have helped stimulate a visit to Alderney by Friedrich Röhrer some 6  weeks later, in December, to report specifically on an aspect of that topic.

8 Alderney

295

Fig. 8.3  Wehrgeologische BodenKarte. Military geological ground map of Alderney, original at scale of 1:10,000, drawn to accompany a military geological description by Walter Wetzel dated 20 October 1941. Key to geology, from the top down: yellow diagonal lines = alluvium, brown diagonals = sandstone, dark red = porphyry, brown = kersantite, red vertical lines = granite, green verticals  =  diorite. Symbols: brown  =  dry valleys, green  =  valleys with water courses, thin red line  =  faults, dashed red line  =  lithological boundaries. Geological boundaries apparently after Parkinson and Plymen (1929): see Fig. 2.31. From Rose (2005a); reproduced by permission of the Bundesarchiv-Militärarchiv from file RH32/3082

8.2.3  Friedrich Röhrer Röhrer’s report of 19 December 19413 dealt with the supply of drinking water to headland sites both on Guernsey (see Sect. 6.3.2) and Alderney. After its opening paragraph, it concluded that the rocky headland sites on both islands were too small to yield much fresh water from the lens that might be expected to overlie the sea water in these regions on both islands. There were therefore only two possibilities to consider with respect to either island: 1 . connection with the main water supply or, 2. transporting water by bowser from the nearest suitable source. After discussion of the situation on Guernsey, the report concludes (in translation):

 Untersuchung der Inseln Guernsey und Alderney, die Trinkwasser-versorgung der auf Landzungen vorgeschobenen Stellungen betreffend by KVR Röhrer, for Insp. Land. O. West. 3

296

E. P. F. Rose

Fig. 8.4  Wasserversorgungskarte. Water supply map of Alderney, original at scale of 1:10,000, compiled to accompany map shown as Fig. 8.3. Key: small blue circle = sites of individual wells, large blue circle = sites of several adjoining wells, green line = valleys containing a watercourse, short double red line = site of a small dam. From Rose (2005a); reproduced by permission of the Bundesarchiv-Militärarchiv from file RH32/3082

‘On the island of Alderney the existing water supply systems can be expanded in a similar manner [to those on Guernsey]. Four pumping stations are already available. The most important of them is located in the NE of the island in Berry’s Quarry and pumps the water standing here in a freshwater pond. It primarily supplies Fort Albert and the arsenal. The second pumping station is located in the west of the island, to the west of [the town of] St. Annes, for valleys leading down to the north coast. It supplies water from the valley creeks to Fort Tourgis. The third pumping station is located in Nunnery Farm, but has never been in operation. The parts of the machinery lying around seem to be quite new, but have been damaged or destroyed. It was probably intended to pump water supplied by the first pumping station to Nunnery Farm, up to Fort Essex by means of this station. The fourth pumping station is near the arsenal. Its purpose is not yet entirely clear, and will be established only after a detailed examination of the pipelines’. ‘Using the first two pumping stations it should be possible to create a central water supply system by pumping to water towers on the highest parts of the island, for example, the “Rond But”. However, the same considerations apply here as on the island of Guernsey’.

8 Alderney

297

Fig. 8.5  Baustoff- und Minier-Karte. Construction materials and mining conditions map of Alderney, original at scale of 1:10,000, compiled to accompany maps shown as Figs. 8.3 and 8.4. Key: yellow = areas of gravel and sand, brown = sandstone, blue-green = diorite, red = granite (red), area shaded diagonally with brown lines = bedrock covered by a firm loam, area outlined in blue pecked line = area probably most difficult for tunnelling, brown symbol = quarries: good starting points for tunnels, red symbol = old ‘Pinge’ (mining attempts). From Rose (2005a); reproduced by permission of the Bundesarchiv-Militärarchiv from file RH32/3082

‘As to option 2, it must be said that the supply of water by means of bowsers can also be taken into consideration, but there is a significantly greater degree of uncertainty than associated with option 1’.

8.3  Army Geologists in 1942 1942 marked the year of most intensive geological work on the Channel Islands. For Alderney, work in the first 3 months was conducted principally by Dieter Hoenes, with geophysical input in March–April from TKVR ‘Dr.’ Kliemstein, perhaps with some assistance from TKV Inspektor Woerner. Thereafter until the end of the year responsibility lay with Wehrgeologenstelle 4 (Rose 2007).

298

E. P. F. Rose

8.3.1  Dieter Hoenes Dieter Hoenes (Sect. 6.4.2) was evidently assigned to military geological tasks on Alderney within a few weeks of the submission of Röhrer’s report: he compiled at least seven reports of his own, two in January and five in February (Table 8.1). The January reports were effectively preliminary site investigations, for a hospital and for an anti-tank wall. These were written whilst Hoenes was attached to the Alderney section of Fortress Engineer Staff 14,4 the first of two sections in a unit whose headquarters and second section were based on Jersey (as noted in Chap. 3). Although appointed as a technical expert,5 Hoenes completed his first report in the rank of lance corporal.6 This was unusual: reports were usually only issued by an officer or an official of equivalent status. However, promotion had presumably been

Table 8.1  Reports (numbered sequentially here for convenience of citation in the text) prepared by Dieter Hoenes early in 1942, with translation of title, author’s rank and assignment, date of completion, and indication of length; all preserved within file number RH32v.3043 at the Bundesarchiv-Militärarchiv 1. Bericht über die Gesteinsbeschaffenheit im Gebiete westilich der La Tohue-Bucht (Standort des geplanten Insel-Lazarettes) [Report on the composition of the rocks in the region westward of La Tohue Bay (Site of the proposed island hospital)]. Sachbearbeiter Gefr. Dr. Hoenes. Fest. Pi. Abschn. Gruppe I/14. 20 Jan 1942 [3 pp.] 2. Bericht über die geologischen Verhältnisse des Untergrundes der geplanten Ufermauer [Report on the geological conditions beneath the planned sea wall]. Sachbearbeiter TKVR Dr. Hoenes. Fest. Pi. Abschn. Gruppe I/14. 22 Jan 1942. [6 pp., map at 1:10,000 missing] 3. Vorläufiger Bericht über die Wasserverhältnisse der Kanalinsel Alderney [Preliminary report on water conditions in the Channel Island of Alderney]. Sachbearbeiter TKVR Dr. Hoenes. Wehrgeol. Stelle 7 bei In. West. 1 Feb 1942. [5 pp.] 4. Planung für eine Wasserversorgungsanlage der Insel Alderney [Planning for a water supply system for the island of Alderney]. Hoenes. 1 Feb 1942. [1:10,000 map] 5. Bericht über die geologische Verhältnisse in den für die Anlage von Hohlgangsbauten vorgesehen en Gebieten der Kanalinsel Alderney [Report on geological conditions affecting the excavation of underground facilities to be provided in parts of the Channel Island of Alderney]. Sachbearbeiter TKVR Dr. Hoenes. Wehrgeol. Stelle 7 bei In. West. 5 Feb 1942. [12 pp.] 6. Bericht über die Wasserversorgung der Kanalinsel Alderney [Report on the water supply of the Channel Island of Alderney]. Sachbearbeiter KVR Dr. Hoenes. Wehrgeol. Stelle 7 bei In. West. 12 Feb 1942. [12 pp.] 7. Abschrift. Niederschrift über die Erkundung der Wasserversorgung der Insel Alderney am 12, 13 und 14 Feb 1942 [Copy. Minutes of an exploration for water supply on the Island of Alderney on 12, 13 and 14 February 1942]. Festungs-Pionierkommandeur XIV. 19 Feb 1942. [6 pp.]

 Abschnittgruppe I/14.  Sachbearbeiter. 6  Gefreiter. 4 5

8 Alderney

299

authorized, for only 2 days later Hoenes’s second report was completed in the rank of Technical War Administration Officer.7 Transfer from the Fortress Engineer Staff to a military geology centre/team (Wehrgeologenstelle 7) of the German Army’s military geological organization (see Chap. 3) quickly followed. Hoenes was thereafter assigned to tasks not by Fortress Engineer Staff 14 on Jersey but, like Röhrer, by the Inspectorate of Land Fortification (West), based at Paris. Apart from his final report, on an exploration for water supply conducted under the auspices of Fortress Engineer Command XIV (the Guernsey-based unit in direct overall command of the fortification programme on the Channel Islands), all reports by Hoenes that were completed in February were issued from Wehrgeologenstelle 7. The February reports included a relatively substantial (12-page) document on the excavation of underground facilities: tunnelling was evidently already a significant feature of the fortification programme on Alderney, and its geological aspects were being taken into consideration. The other four items all related to water supply: a topic that was to be of major concern in future geological studies. From the dates given on his reports, Hoenes was apparently active on Alderney for several weeks, from at least mid-January to mid-February. His early reports were to be superseded by longer-term studies in the months that followed.

8.3.2  The ‘Geologist’ Johann Kliemstein A report by TKVR Lautmann8 of the Inspectorate of Land Fortification West records that he was accompanied during a visit to Alderney on 29 March 1942 by two engineer majors, one representing Fortress Engineer Command XIV and the other Fortress Engineer Staff 14, and by four geologists: TKVR Professor Röhrer, the geologist at the Inspectorate of Land Fortification West (as described in Sect. 6.3); TKVR Dr. Beschoren, of Wehrgeologenstelle 4 (Sect. 6.4); and two geologists listed as such but without ascription to any unit, TKVR Dr. Kliemstein and TKV Insp. Woerner. ‘Dr.’ Kliemstein must be the man listed by Häusler (1995a, b) as Dipl.-Ing. Johann Kliemstein: the only German military geologist of World War II with that surname, and for whom the significant career details fit. The ‘Dr.’ is presumably a courtesy title, based on the supposition that Kliemstein had graduated with a qualification equivalent to a doctorate: the traditional end-point of a university education in Germany. Dip.-Ing. Franz Schulte was similarly credited as ‘Dr. Schulte’ when he visited Walther Klüpfel on Jersey in October 1942 (as noted in Sect. 5.6).

 Technischer Kriegsverwaltungsrat: TKVR.  Die Besichtigungsreise zu den Kanalinseln vom 24.3 bis 2.4.42, of 4 April 1942: now preserved at the Budesarchiv-Militärarchiv in file RH32v.3041. 7 8

300

E. P. F. Rose

Fig. 8.6  Dipl.-Ing. Johann Kliemstein, a postwar photograph. From Oberste Bergbehörde (1961)

Kliemstein (Fig. 8.6) was by 1942 an experienced geophysicist with a military background; with experience also of some 3  years of service within the German Army’s military geological organization; and at nearly 43 years of age, a man arguably in the prime of life (Oberste Bergbehörde 1961). He had been born on 2 June 1899 at Belowar, a town then within the Austrian Empire but within independent Croatia at the present day. He had been educated at school in Vienna (the Empire’s capital city), subsequently at a military college at Mährisch Weisskirchen (a town in Moravia, now in the eastern part of the Czech Republic), and finally, from 1917, at Austria’s prestigious Military Academy housed in the castle at Wiener Neustadt, from which he graduated with honours in 1918—but as World War I came to an end.9 From 1919 to 1925 Kliemstein attended the University of Leoben: Austria’s specialist university for mining, metallurgy, and materials science. He qualified as a mine surveyor in 1924, and (with distinction) as a mining engineer in 1925. His student years included significant practical experience in the mining industry, notably in coal mines both in Austria and the Ruhr district of Germany, and from 1923 to 1925 he served also as a teaching assistant in the department of geodesy and mine surveying at the university. From 1925 to 1934 he was employed as a mining surveyor at Seegraben, the oldest coal mine in Austria, and from 1932 also as an honorary lecturer at the University of Leoben. In 1935, and therefore some 3 years before Austria’s annexation by Germany, Kliemstein moved to Germany, to Berlin, to obtain commercial employment in testing and developing geophysical instruments. In 1937 he married. In 1938, still in

 Founded in 1751, the ‘Theresian’ Military Academy—where officers are trained for the Austrian armed forces—is one of the oldest military academies in the world. 9

8 Alderney

301

Berlin, he achieved appointment to the Prussian Geological Survey, shortly before this was amalgamated (in April 1939) with Germany’s other regional geological surveys to form a national geological survey.10 His assignment was to geophysical and geological projects concerned with mineral resources, especially ores and petroleum. His employment with the (national) Survey continued to 1942, but from 1939 he was seconded to serve as a military geologist (Häusler 1995b, p. 26). TKVR Kliemstein served from early in the war with the military geology ‘group’ led by Fr. Schuh11 on the Western Front (Häusler 1995a, p.  75). Schuh was supported at his headquarters by a young geologist as an adjutant, and by specialists for water works, mining, road construction, electrical engineering, blasting, and geophysics, whilst geologists as such were dispersed to outstations. Kliemstein was the specialist in geophysics. He was recorded as with Schuh in May 1940 (Häusler 1995b), with him in Brussels the following month (Häusler 1995a), and from July 1940 still with him in Brussels with the ‘group’ which was now assigned to the engineer reconnaissance staff for Belgium and northern France.12 In August, Häusler (1995b) records Kliemstein as assigned to Poland under the Commander-in-Chief on the Eastern Front,13 but by July 1942 he was in post as a military geologist (Wehrgeologe) with a geophysical ‘earth electricity’ reconnaissance troop based at the Inspectorate of Land Fortification (West).14 Kliemstein’s appointment to the ‘troop’ must have been earlier than this, because on 16 April 1942 he completed a report as from that unit. This provided the results of an earth resistivity survey carried out in March and April to help assess groundwater conditions on Alderney15—although his initial as author is shown as ‘H’ rather than ‘J’, and it is as ‘H. Kliemstein’ that he is listed as a former (1938–1939) member of the Prussian Geological Survey in Berlin.16

8.3.3  TKV Inspektor Woerner There is no record of a military geologist by the name of Woerner in Häusler’s (1995a, b) listing of the German military geologists of World War II. However, it is evident from a later report by Dieter Hoenes (Sect. 8.4) that he was present on

 Reichsamt für Bodenforschung.  Wehrgeologengruppe SCHUH. 12  WG-Gruppe/Pi-Erkundungsstab für Belgien und Nordfrankreich. 13  General der Pioniere bei Oberbefehlshaber Ost; Militärbefehlshaber im Generalgouvernement (Polen). 14  E-Erktrp 2 beim Insp. d. Landesbef. West (Wehrgeologe). 15  Bericht über die erdelektrischen Erkundungsmessungen zur Ermittung der Grundwasserverhältnisse auf der Insel Alderney, by H.  Kliemstein. (Erdelektrischer Erkundungstrupp 2 bei Kommandodienststelle Fest. Pi. Abschn. Gr. I/14.) 16 April 1942. [14 pp., 17 figs, 1 map.] Bundesarchiv-Militärarchiv file RH32v.3924. 16  www.pgla.de/direkt.htm. 10 11

302

E. P. F. Rose

Alderney by at least 12 February, and that he was of sufficient geotechnical proficiency to be mentioned by rank and name amongst a group of geologist and engineer officers or officials concerned with aspects of the island’s water supply. His ‘TKV Inspektor’ status is not typical of the ranks assigned to members of the Organisation Todt: it was a rank held by officials serving with the German Army, junior to that of TKVR: Technischer Kriegs Verwaltungs Rat. Possibly he was a member of Kliemstein’s ‘earth electricity’ reconnaissance ‘troop’ making preparations in February for the survey that began in March but, if so, he failed subsequently to achieve promotion to full ‘Wehrgeologe’ status. Alternatively, he may have held an appointment within the local Fortress Engineers—in which case, one would expect his assignment to have been indicated, as normally in German military reports. What is clear is that although he was perceived as a geologist and an official with some authority on Alderney, he did not make a signed contribution to the documentary record of geological work either on Alderney or elsewhere.

8.3.4  The Earth Resistivity Survey The purpose of the earth resistivity survey is set out in the opening paragraph of Kliemstein’s report: it was to clarify the hydrological conditions on the island by a survey complementary to the geological studies that had already been carried out. Two main questions were to be addressed: 1. Are the deeply weathered crystalline rocks (granite and diorite) and the Cambrian sandstone to be regarded as potential aquifers and if so, at what depth within them does usable groundwater occur? 2. Can it be proved by means of earth resistivity that a lens of fresh water lying upon sea water extends far into the interior of the island? As to the geological and hydrological conditions revealed by outcrops, springs, and wells, the reader is referred to the reports by Wetzel of 20 October 1941 and by Hoenes of 20/22 January, and of 1 and 5 February 1942. The map which accompanies the report and served as a base map for the survey is a copy of Wetzel’s military geological map of Alderney (Fig. 8.3). The report presents the results of 13 profiles across parts of the island, and then summarizes results in terms of three regional complexes (granite, diorite, and sandstone) with discussion of how far the main questions had been answered. The survey is seemingly the first time that resistivity methods had been applied to the search for groundwater on the Channel Islands. Geophysical expertise was, however, by then already well established within the German Army. A reconnaissance ‘troop’ specifically for resistivity surveys had been formed in September 1940 under the auspices of a ‘military geology group’ based in western Germany at Baden-Baden (Häusler 1995a, p.  76). The ‘group’ had a headquarters at Baden-­ Baden, three branches at towns in the Lower Rhine, five reconnaissance ‘troops’ based at towns in part of the then German-occupied region of northern France, and

8 Alderney

303

two specialist ‘troops’: one for water supply and the other for ‘earth electricity’ (Fig. 3.3). Moreover, resistivity techniques were used by German military geologists from 1941  in their search for groundwater in North Africa to support the desert campaigns led by Erwin Rommel (Willig and Häusler 2012). In contrast, the British Army had no unit of geophysicists to use for ground investigations: for such studies in the arid regions of North Africa and the Middle East it relied upon a unique unit of the South African Engineer Corps: 42nd Geological Section (Rose 2012, 2018a, b).

8.3.5  Wehrgeologenstelle 4 In March 1942, whilst Kliemstein’s survey was in progress, Hoenes was transferred to Wehrgeologenstelle 4 (Rose 2007). He was to be based on Guernsey as deputy to Bernard Beschoren for a programme of work on the Channel Islands that was to last for the rest of the year (as described in Chap. 6). The team was to generate a further 14 reports for Alderney (Table 8.2), of which eight were compiled by Hoenes alone, and four (perhaps five) jointly with Beschoren. Apparently no further reports were completed for Alderney in March or early April, as Wehrgeologenstelle 4 established its base on Guernsey. One was completed on 26 April; four in May; one on 19 June; one on 7 July, four in August; one on 7 October; one on 25 November; and finally, one on 15 December. Apart from the month of September, when the geologists were either on leave or focused on work in Guernsey, geological work therefore continued on Alderney until almost the end of 1942. Of the 14 known geological reports for Alderney generated by the team, mostly by Hoenes (Table 8.2), 11 dealt with various aspects of water supply, one with tunnelling, one with burying a cable network underground, and one with naturally occurring raw materials for the construction of fortifications.

8.4  Water Supply Hoenes compiled a report on 27 May (Table 8.2, item 5) that provided an authoritative assessment of the current state of water supply on Alderney. This noted that planning for a water supply began in January and February 1942. Reports by Hoenes himself of 1 February (Table 8.1, items 3 and 4) provided the geological basis for subsequent assessments. Preliminary assessment indicated that water supply would be a problem because of the small size of the island, its geology, climatic conditions, and dense occupation, and that a comprehensive supply network would need to be established. Strongest underground water flow appeared to be within the transition zone between the superficial sandy loam that covered much of the island plateau and the igneous bedrock: a zone that included that of weathered basement rock. This shallow groundwater was recorded to issue as springs in the upper parts of valleys

304

E. P. F. Rose

Table 8.2  ‘Reports’ generated by Wehrgeologenstelle 4 for Alderney, numbered sequentially here for convenience of reference in the text, with translated title, date of completion, author, page length, and current file number at the Bundesarchiv-Militärarchiv 1.

2. 3. 4.

5.

6.

7.

8.

9.

10.

11.

12.

13.

14.

Grundwasser und Quellschutzgebiete auf der Kanalinsel Alderney [Groundwater and spring protection areas on the Channel Island of Alderney]. 26 April, by Hoenes, 3 pp., [1 map now missing] [RH32v.3043] Wasserversorgung Alderney [Water supply Alderney]. 10 May, by Hoenes, 3 pp. [RH32v.3043] Bericht über den Wasserinhalt der Steinbruche auf Alderney [Report on the water content of quarries on Alderney]. 18 May, by Hoenes, 1 p. [RH32v.3043] Bericht über den Stand der chemischen und bakteriologischen Untersuchen der Wasserverkommen auf Alderney [Report on results of the chemical and bacteriological examination of water supplies on Alderney]. 23 May, by Hoenes, 2 pp. [RH32v.3043] [Gutachten 1] Bericht über den derzeitigen Stand der Wasserversorgung von Alderney [Report on the current state of water supply from Alderney]. 1oc Br. B. Nr. 46/42 geh, 27 May 1942, by Hoenes, 9 pp., 2 maps at 1:10,000 [RH32v.3028, also v.3043 where the endorsement ‘Gutachten Nr. 1’ has been added in ink] Geologischer Bericht: Die geologischen Verhältnisse von Ho1 (Mannez Quarry) and Ho2 (Fort Essex) auf “America” [Geological Report. Geological conditions [at the tunnels] Ho1 (Manez Quarry) and Ho2 (Fort Essex) on “America” [codename for Alderney]]. 19 June, unsigned. [RH32v.3043] Bericht über den Befund der chemischen Untersuchungen der Wasservorkommen von Alderney [Report on the results of chemical analyses of water supply from Alderney]. 7 July, by Beschoren & Hoenes, 4 pp., 1 map, annex 3 pp. [RH32v.3028, also v.3043] Gutachten 4. Gutachten über die zu erwartenden Untergrunds-verhältnisse bei der Anlage des festungsmässingen Kabelnetzes [‘Expert opinion’ on the anticipated subsoil conditions when creating the projected cable network for the fortress] Az.39 Geol. 10 b Br.B.Nr.72/42, 15 Aug 1942, by Hoenes, 3 pp. plus 1:10,000 map showing route of cable network on Alderney [RH32v.3043] Gutachten 5. Gutachten über weitere Arbeiten zur Erkundung von Wasserverkommen auf der Kanalinsel Alderney [‘Expert opinion’ on further work to explore water supply on the Channel Island of Alderney]. 1oc Br. B. Nr. 73/42 geh, 15 Aug 1942, by Beschoren, 3 pp. [RHv.3028, also v.3043] Gutachten 6. Gutachten über die geologischen Verhältnisse im Gebiet der bei Mill-Farm auf Alderney geplanten Stauanlage [‘Expert opinion’ on the geological conditions in the region of a dam planned near Mill Farm on Alderney]. 1oc Br. B. Nr. 75/42 geh, 21 Aug 1942, by Beschoren & Hoenes, 2 pp. [RH32v.3031, also v.3043] Anlage des O T Lagers bei “Haus Keller” innerhalb des Quellschutzgebietes westlich Fort Essex [Construction of the Organisation Todt camp ‘Haus Keller’ within the groundwater protection area west of Fort Essex]. 25 August, by Beschoren & Hoenes, 3 pp., 1 fig. [RH32v.3028, also v.3043] Gutachten 9. Erläuterungen zur Baustoffkarte der Kanalinsel Alderney [Notes on the raw materials map of Alderney]. 1oe Br. B. Nr. 94/42 geh, 7 Oct 1942, by Hoenes, 8 pp., 1 map [RH32v.3027, also v.3043] Bericht über weitere Arbeiten zur Sicherung der Wasserversorgung von Alderney [Report on further work to secure the water supply of Alderney]. 25 November, by Hoenes, 4 pp., 1 fig., 1 map Wasserversorgung 1:10,000 Anlage zum Bericht von TKVR Dr. Hoenes 25 Nov 42. [RH32v.3043] Gutachten 10. Weiterer Ausbau der Wasserversorgung Alderney [Further expansion of the water supply of Alderney]. 1oc Br. B. Nr. 115/42 geh, 15 Dec 1942, by Beschoren & Hoenes, 5 pp. [RH32v.3029. also v.3043]

8 Alderney

305

Fig. 8.7  German sketch plan of an infiltration gallery. Sickergräben  =  infiltration trenches; Sammelschacht  =  storage chamber; Entnahme  – Leitung  =  offtake conduit; Drahtzaun um den Schutzbezirk = wire fence to define protected area (at about 50 m distance from the installation). Narrow, branching slit trenches excavated in gently sloping ground (see contours) lead to a storage tank in which water is collected prior to offtake via a conduit pipeline. From Geologen-­ Gruppe (1918)

that cut through the island plateau. For the development of a future water supply system, it was recommended that this be done by the construction of infiltration galleries, thereby sealing off some of the valley outlets where appropriate. Perhaps curiously, this report makes no mention at all of Kliemstein’s earth resistivity survey. The construction of infiltration galleries (Figs. 8.7 and 8.8) was one of the means of water supply widely used by the German Army during World War I, and comprehensively described in textbooks that benefited from that experience (e.g. von Bülow et al. 1938). The report of 27 May notes that meetings were held on 12 and 13 February between Hoenes and Major (Ing.) Boerzel, Reg. Bauassessor Scheidt, KV Insp. Woerner, Kulturbauingenieur Schneider and Reg. Bauinspektor Baumgartner of the Organisation Todt (see Table 8.1, item 7). (Schneider was presumably the man soon to be appointed as a geologist to the Luftwaffe’s Field Works Office for the Channel Islands, on Guernsey: as described in Sect. 7.2. Woerner was to be listed in a later report as a geologist: Sect. 8.3.3). As a result, precise guidelines for further planning and for the design of equipment were established, as summarized in reports of 12 and 19 February (Table 8.1, items 6 and 7). The 19 February recommendation that groundwater should be collected by means of infiltration galleries constructed at the heads of valleys, and be piped from there, was endorsed in the 27 May report. The system would require the pumps and

306

E. P. F. Rose

Fig. 8.8  German cross-section (left) and longitudinal section (right) through part of a completed infiltration gallery (cf. Fig. 8.7), excavated successively through (a) loam, (b) loamy gravels, and (c) clay. Gelochte Rohre = perforated pipes; Sickerkanal = infiltration channel; Filterpackung aus Kies oder Klarschlag (gewaschen)  =  filter pack of gravel or crushed rock aggregate (washed); Gestampfter Ton oder Lehm = compacted clay or loam. Water from the permeable gravels filters though the gravel pack ultimately into the pipe leading to the storage tank. The impermeable clay or loam seal at the top prevents contamination from the ground surface. From Geologen-­ Gruppe (1918)

filtration plant necessary to supply water towers to be situated on the highest points of the island. These would be linked as two loops, one for the west of the island and the other for the east. (Maps to show the developing supply system with its infiltration trenches, pipelines, water towers, and pumping stations, were to be generated at several stages during 1942, e.g. Fig. 8.9.)

8.4.1  W  ater Supply Works Described in the 27 May 1942 Report The meetings of 12 and 13 February and consequent report of 19 February had established that a large number of infiltration trenches would need to be constructed in order to provide an insight into groundwater conditions, and the chemical and bacteriological quality of the water so obtained would need analysis. Prospecting was to be the responsibility of the Organisation Todt. Trenching for the infiltration galleries should have begun immediately, with completion in about 5 weeks. However, due to shortage of labour, only one trench was actually completed within this time. This was a trench to 3 m depth, sited at the Organisation Todt base

8 Alderney

307

Fig. 8.9  Wasserversorgung. Water-supply map of Alderney, original at scale of 1:10,000, compiled by Dieter Hoenes, presumably in 1942. The topographical base map has annotations inked upon it to show (from top down in the key): positions for planned infiltration galleries; course of planned conduits; existing conduits; reservoirs; and motor pumps. From Rose (2005a); reproduced courtesy of the US National Archives and Records Administration

north of Fort Essex. It cut through about 2.5 m of dune sand, overlying sandy loam. Water was abundant above the loam, particularly in the southern branch of the trench, with a yield of 0.5  l/s. By then exploratory boreholes had proved that unweathered bedrock lay here at a depth of about 8  m: dune sand 0–3  m, loam 3–5 m, weathered diorite 5–8 m, fresh diorite below 8 m. From this profile it was evident that there were two water-bearing zones: one above the loam, the other as usual at the top of the unweathered bedrock. The upper zone could be utilized by the planned infiltration trenches, whereas in the final expansion of the system several wells would be drilled to exploit the deeper water zone. Subsequently, between 12 March and 18 April, about 30 workers were used to develop five sites for water supply: 1. A trench to 3 m depth in a small valley on the western slope below Fort Platte Saline (east of Mill-Farm) yielded ample water, at the rate of 0.3–0.4 l/s, from the unconsolidated material above bedrock. When sufficient labour was available, this trench would be extended and deepened. The water was free of bacteria and suitable for use as drinking water.

308

E. P. F. Rose

2. A trench dug to 2.5 m depth in the upper part of a small valley further to the east also yielded a good water flow. Here the water-bearing zone was in the top 0.5 m of weathered material, overlying the rock talus slope. There were considerable fluctuations in water flow, which were suspected as indicating contamination by sewage outfalls from the town of St. Annes, so the work was stopped. However, on examination, the water proved to be free of bacteria. 3. During this period a spring was discovered independently by the Organisation Todt in a valley south of the gasworks, and an infiltration trench dug. Together they produced a yield of 0.8 l/s and fed an 18-m3 reservoir that supplied water to the OT kitchens near the gasworks. 4. Instead of the designated trenches at Rose Farm, the OT arbitrarily created two more basic trenches and a collection conduit in the upper part of the valley to the west. These installations were to be regarded as provisional. It was recommended that they should be reconstructed properly as infiltration galleries to ensure that the water so obtained was residue-free. However, the two excavations already produced an abundant water supply. Together they yielded 1 l/s, a supply that merited installation of a pump. Water samples contained the bacterium E. coli, so the water needed to be boiled before use. One trench penetrated soil from 0.0 to 0.4 m, blue-grey, loamy rock residue with fragments of granite from 0.4 to 1.2 m, brown-yellow rock detritus from 1.2 m, with a strong water flow at 2.5 m depth. The other, western trench penetrated soil from 0.0 to 0.2 m, brown sandy loam from 0.2 to 0.9 m, light grey very fine sand from 0.9 to 1.1 m, fine sand of dark brown colour 1.0–1.45 m, yellow-brown weathered rock fragments from 1.45 m, with strong water flow at 2.5 m. 5. Infiltration galleries were dug under Hoenes’ supervision in the shallow valley south of the road at Rose Farm during the second half of April. The same conditions were revealed as for locality 4. Typically: soil from 0.0 to 0.3  m, light brown sandy loam from 0.3 to 1.0 m, fine light grey sand from 1.0 to 1.6 m, and brown weathered granite rubble with rock fragments from 1.6 m. The thickness of the strata varied somewhat, especially the fine sand. The strongest water flow was found in the top 0.7 m of the weathered granite. An abundant water supply (yield 0.4 l/s) was found in the most advanced of two trenches. The old British pump currently connected to a well shaft at the head of the valley would be removed and used to help complete work on the gallery facilities. The conclusion at this time (27 May) was that, below Alderney’s plateau surface, the strongest groundwater flow was to be found within the fragmented weathered zone that marked the top of the igneous bedrock. This ‘near-surface’ water occurred across the area, but was best exploited by construction of infiltration trenches to seal off its exit to the valleys. During prolonged periods of low rainfall the yield might fall or even cease. This needed to be taken into account in further planning.

8 Alderney

309

8.4.2  Priority Action The 27 May report then went on to describe immediate measures to be taken for an enhanced water supply. This was necessary because of the substantially increased garrison strength on the island, high demand for water for construction purposes, and the onset of summer and so a period of drought. However, it was realized that at the time of writing the OT was not in a position to release more labourers to complete the programme of work in hand in the necessary short time. A series of priority measures was therefore recommended: 1. Construction of a filling point west of Rose Farm at the junction of the road to Fort Tourgis, to provide an additional source of supply for the town of St. Annes and the batteries and strongpoints of the western part of the island. Water from the two existing infiltration trenches was to be stored in an iron tank (of about 9 m3) sunk into the ground, with flow achieved by means of a motor pump. 2. After this, an additional extraction point for the area was to be created on the road south of Rose Farm, using the existing outlet, in order not to waste the considerable quantities of water available. Water was to be collected in two metal containers, from which it could be pumped by hand or by means of a wind-­ driven pump. For the time being, points 1 and 2 were not to be used to provide water for construction purposes, only for drinking. 3. As a source of water for concreting, there was an outlet in the quarry at the small harbour, connected to two water tanks, and with a motor pump installed. If as a result of drought in the summer months the existing outlets no longer supplied enough water, the final resort must be to extract water from the disused quarry in the NE of the island, which contained about 50,000 m3. In this case it would be necessary to provide more pumps and storage tanks, from which bowsers could be supplied. The quarry was able to meet the total island water requirements for several periods of drought. However, it was at risk of contamination by enemy action (gas attack, or introduction of bacteria), which would make the water unusable. A following section of the report dealt with measures for protection of ground and spring waters from contamination. It emphasized that a supply of water safe to drink was only possible if measures to prevent contamination were strictly observed. Groundwater and source protection areas had been delineated by Hoenes and set out in a report issued on behalf of the Island Commander on 26 April 1942 (Table 8.2, item 1). Some refinements had been made in a letter issued subsequently, in May (Table 8.2, item 2). In groundwater protection areas, there were to be no OT camps, troop billets, fuel stores, butcheries, laundries or de-lousing stations. If at all possible, they were to be kept free of latrines. Where the essential siting of strong points and artillery batteries prevented this ideal, cesspits had to be constructed from impermeable concrete, and to be emptied regularly. By building the OT camp south of Ho-Höhe at ‘Haus Kellar’, the water intake from an infiltration trench planned for the valley below it had been put at risk, as indicated in memoranda dated 26 April and 15 May. Relocation of the camp was

310

E. P. F. Rose

recommended, in the interests of securing safe water supplies. If that were no longer possible, then the water intake system ought not to be completed. This would have implications for the planned island water supply grid, because of the importance of the infiltration trench and associated wells. However, the OT had already agreed to implement a series of measures to keep the camp clean, and to install sealed cesspits. Demarcated water protection areas were to remain free of all construction projects. They were not to be entered, and therefore to be surrounded by a wire fence to be constructed by an appropriate Engineer Company. Fertilizing with manure or liquid manure or the grazing of cattle within a radius of 300 m of a protected area was prohibited. A report on chemical and bacteriological analysis of water samples taken from 12 localities on Alderney had been completed earlier in May (Table 8.2, item 4), and was summarized in the 27 May report. Nine of 14 samples were free of contamination, but the field laboratory on Guernsey had detected the bacterium E. coli in five samples. More chemical analyses were urgently required to facilitate planning of the filtering and storage systems. The water was expected to be highly acidic and, if so, this would affect the choice of cement to be used for the various installations now planned. A letter of 23 May to the geologist at the Inspectorate of Land Fortification West had asked that a chemist be appointed for this task. The report of 27 May ended with a short list of recommendations for the installation of an infiltration gallery south of Rose Farm, an associated storage tank, a water tower west of St. Annes, and infiltration galleries at four other localities, together with recommendations concerning water supply for the continuing fortification programme. The project was dependent on the availability of adequate labour to complete the infiltration galleries, well points, reservoirs, pumping stations, and filtration plants associated with the existing plan. A detailed design for this in plan and profile view, with conduit and reservoir dimensions, was to be prepared by the Luftwaffe’s Field Works Office for the Channel Islands, on Guernsey. The water catchments, reservoirs, and pipelines on which work had already begun were to be made compatible with this final scheme. Facilities were to be constructed to ‘shatterproof’ field standards, with the potential subsequently to increase the concrete thickness of the outer walls to 2 m (‘fortress’) standard.

8.4.3  Final Hydrogeological Studies The recommended chemical analyses were duly completed, in July (Table 8.2, item 7). ‘Expert opinions’ were compiled in August on the need for further work to develop the island’s water supply (Table 8.2, item 9); on site investigation for a potential dam site (Table 8.2, item 10); and on the predicted problems associated with the development of a camp for Organisation Todt workers within one of the groundwater protection areas designated on Alderney (Table 8.2, item 11)—a situation noted above. This ‘opinion’ focused on the location of an infiltration gallery and its associated storage tank in the valley downslope to the NE of the camp (Fig. 8.10).

8 Alderney

311

Fig. 8.10  Plan showing proposed water supply for the camp housing Organisation Todt workers NW of Fort Essex on Alderney. Annex to a report dated 25 August 1942 compiled by Dieter Hoenes of Wehrgeolenstelle 4. Reproduced courtesy of the Bundeswehr Geoinformation Centre

312

E. P. F. Rose

A further ‘opinion’ on water supply was issued on 25 November (Table 8.2, item 13), and Wehrgeologenstelle 4’s final ‘opinion’ for the Channel Islands as a whole on 15 December provided the final assessment of works needed for further expansion of the water supply on Alderney (Table 8.2, item 14). This ‘opinion’ by Hoenes recorded that discussion had taken place with members of the Army’s engineer staff, together with site investigations, with regard to five projects: 1. Water supply for the filtration system in the valley north of St. Annes. This facility was planned to lie between the valley and ‘Camp Deubau’: a camp name that does not feature in later known literature, but which possibly refers to an early camp known to have been sited in this area—later used for OT storage (Trevor Davenport, pers. com. 2020). The report provided a detailed assessment of the water conditions based on knowledge of springs and groundwater, and of the conditions likely to affect sewerage. 2. Water supply for the camp operated by the Firm Sager and Wörner at the Divisional radio station. To supply the camp, a concrete storage tank should be built here, and filled daily by water bowsers. Emplacement of either wells or infiltration galleries around the camp had no prospect of success. 3. Well shaft at ‘Exit no. 2 West’. Here there were 2 m of water in a shaft next to a tunnel entrance created for training purposes by one of the Engineer Mining Companies. The shaft could not be developed as a well because of its unfavourable position: water was drawn off by the tunnel. 4. Infiltration gallery and water point NW of Fort Essex. The most important installation for water supply to the dry eastern part of the island was the infiltration gallery nearing completion NW of Fort Essex and the primitive well already in operation. The southern branch of the gallery had been found to bring in the most water, from the southern slope, and this was therefore to be developed by a forked branch and by deepening by about 1.5–2.5 m. 5. Water supply to the ‘Schirrhof’. Two storage tanks of about 15 m3 at Fort Albert were fed by pumping from an old quarry to the NE. One of these, filled once or twice a day, had a pipeline leading to the ‘Schirrhof’. The demand for water at the ‘Schirrhof’ was therefore already high, and increasing it to provide for a laundry with its necessarily high water consumption was undesirable. Given the scarcity of water in the eastern part of the island, it was recommended that the laundry be sited instead in the middle or western part, where water supply was more plentiful. The final paragraph of this final report noted that for the future, it had been agreed that the Luftwaffe’s Field Works Office for the Channel Islands (based on Guernsey) and the Army’s Fortress Engineer Section I/14 (based on Alderney) would consult each other on any further questions regarding water supply on Alderney. Water supply problems were thus of long-term concern to German forces on Alderney, and groundwater studies were the predominant focus of reports compiled by their military geologists. Unfortunately, as described in Chap. 9, British studies immediately postwar were to be carried out without knowledge of and so access to the German data (Robins and Rose 2005; Robins et al. 2012).

8 Alderney

313

8.5  Tunnels A definitive account of the German tunnels on Alderney, both in terms of their history of construction and present-day appearance, has been published by Ginns (1993), an account significantly updated and extended by Gavey and Powell (2012) on the basis of compass and tape surveys by Dr. Trevor Davenport. All the tunnels are illustrated by plans to indicate their layout, and photographs to illustrate key features of their appearance past and present, so far as this is possible. According to the German monthly building progress reports, it was planned that some 10,000 m2 of storage space would be created. In 1942 only about 250 m2 had actually been achieved, rising through 1943 from 2882 m2 in April to 3896 m2 in June and 4795 m2 in September. It remained at this level thereafter, following withdrawal of most of the workforce to France. The figure is equivalent to between 15,000 and 20,000 tonnes of rock excavated—mostly after the geologists of Wehrgeologenstelle 4 had left the Channel Islands. Tunnelling was evidently regarded as a significant component of the Alderney fortification programme even though little was achieved in 1942. Thus Hoenes compiled a relatively substantial report on 5 February 1942 on the geological conditions that would affect the excavation of Alderney’s underground facilities, whilst he was serving with Wehrgeologenstelle 7 rather than Wehrgeologenstelle 4 (Table 8.1, item 5). When the Luftwaffe geologist Professor K.G. Schmidt visited the island on 23 April 1942 (as described in Sect. 5.3), Hoenes guided him to sites where Army work was planned or in progress. The earliest known contemporary map showing tunnel locations is dated little more than a week later: 1 May 1942 (Gavey and Powell 2012). This indicated that nine tunnels were then planned or under construction (Fig. 8.11 and Table 8.3). A map of 1 July 1943 indicated that by then the proposed Ho 9 had been deleted from the plan. (Ho 9 was intended as a Flak (anti-aircraft artillery) ammunition store and so a Luftwaffe rather than an Army responsibility.) However, a tunnel at this site, Rose Farm, was in fact under construction by 1945. According to a map compiled by Gavey and Powell (2012), updated here by Trevor Davenport (Fig. 8.12), in total 24 underground facilities had been or were being excavated on Alderney by the end of the war—although only Ho 1 to Ho 9 are known from surviving German maps. Apart from the information provided by the German maps of 1 May 1942 and 1 July 1943, and some monthly building progress reports that are still extant, there is little other documentary evidence regarding the construction of the Alderney tunnels. Following evacuation of the local civilian population, there were no eyewitness accounts from local people comparable with those made on Jersey and Guernsey. Systematic destruction of records prior to surrender was particularly thorough, possibly because some were associated with activities of the SS-Baubrigade. Thus although the records left by geologists are but few, they are nevertheless of significance in that they help to fill this information gap. The report by TKVR Lautmann cited above (Sect. 6.3.4) records a visit to Alderney on 29 March 1942 when accompanied by TKVR Professor Röhrer (the military geologist at the Inspectorate of Land Fortification, West) (see Sect. 6.3);

314

E. P. F. Rose

Fig. 8.11  Map of Alderney indicating sites of nine tunnels planned or under construction in May 1942; see Table 8.3. From Ginns (1993); reproduced by kind permission of the Channel Islands Occupation Society (Jersey) Table 8.3  Tunnels on Alderney (cf. Figs. 8.11 and 8.12) as planned at 1 May 1942, with intended usage: data from Gavey and Powell (2012)

Ho 1. Munition and ration store Ho 2. Munition store Ho 3. Island hospital Ho 4. Bakery and ration store Ho 5. Fuel store and electricity generating station Ho 6. Munition store Ho 7. Munition store Ho 8. Butchery and ration store Ho 9. Munition store (Flak)

Major (Ing.) Boertzel of Fortress Engineer Command XIV; TKVR Dr. Beschoren, of Wehrgeologenstelle 4; Major Beger of Abschnitt Gruppe I/14 (the Alderney section of Fortress Engineer Staff 14); and the ‘geologists’ TKVR Dr. Kliemstein and TKV Insp. Woerner. The report recorded (very briefly) that for tunnelling on Alderney, the rock was everywhere consistently suitable for blasting; major input was not required; and that work was already in progress. Kliemstein and Woerner were apparently present because they were already serving on Alderney to effect an earth resistivity survey, and were designated ‘geologists’ because their ‘earth electricity troop’ formed part of the German Army’s military geological organization. K.G. Schmidt’s report mentions geological conditions at Trois Vaux Bay, at the western end of the island (NW of Telegraph Bay: Fig. 8.13). This must refer to the site planned for Ho 7 (Fig. 8.10). He notes that tunnel construction here would be influenced by a single direction of weakness in the rock: dominant north to NW joint planes, and fault zones. The geological constraint may help to explain why this

Fig. 8.12  Map of Alderney indicating sites of all tunnels known to have been planned or excavated prior to surrender in May 1945. © Trevor Davenport, and reproduced by kind permission

Fig. 8.13  View across Telegraph Bay (see Fig. 8.12), on Alderney’s SW coast, showing the high cliffs in this region that deterred potential amphibious assault and so reduced the need for defensive fortification in the area; also the weathered granite that formed part of the local bedrock. The telegraph tower (top left) was constructed in 1811, enabling signals to be relayed visually to the island of Sark and on to Guernsey—early warning of potential attack during the Napoleonic Wars (1803–1815) and so initially of strategic importance. Photo: E.P.F. Rose

316

E. P. F. Rose

Fig. 8.14  Plan of Ho 1, at the present day. From Gavey and Powell (2012); reproduced by kind permission of Steve Powell and Festung Guernsey

tunnel, shown on the planning maps of 1942 and 1943 (Fig. 8.11) was never constructed (Fig. 8.12). In contrast, the site he examined at Fort Essex, with two directions of weakness, was excavated as Ho 2. Hoenes completed a report specifically on Ho 1 and Ho 2 on 19 June (Table 8.2, item 6), after his re-assignment in March from Wehrgeologenstelle 7 to Wehrgeologenstelle 4. Ho 1 (Fig. 8.14) was intended as munition and ration store. The layout of the large storage tunnels in Alderney is similar to that of such tunnels in Guernsey: two entrances curving into a main gallery. The entrance tunnels are all about 3 m wide by 2.2–3.0 m high, and the main galleries vary between 3.0 m and 5.6 m wide. At Ho 1 about 25 m of the southern part has been destroyed by postwar quarrying, but the accessible northern entrance is 3  m wide, and a 2.2-m-high tunnel continues straight into the quarry face for 33.6 m before curving to the left for 10.3 m through 90°. Excavation is entirely within the Alderney Sandstone (a rock unit described in Chap. 2: cf. Fig. 8.1). Like Alderney tunnels in general, Ho 1 has been bored to a square profile rather than with a curved roof. An arched profile gives greater roof stability, but Ho 1 was constructed without lining or roof supports and has nevertheless survived until the present day without sign of roof collapse. It is surprising that no side branches or storage chambers were excavated, but since the tunnel remains both sound and dry, it was presumably put to some use. Ho 2 (Fig. 8.15) was another of the large storage complexes, also planned as a munition store, excavated into an old quarry face and through Alderney Sandstone. The tunnel lies beneath the hill surmounted by the British-built Essex Castle: a largely nineteenth-century building more appropriately known as Fort Essex (see Sect. 1.6). The tunnel has again been cut to a square profile, between 2.5 and 3.0 m wide by 2.6 m high, mostly without roof supports, and mostly dry. However, three

8 Alderney

317

Fig. 8.15  Plan of Ho 2, at the present day. From Gavey and Powell (2012); reproduced by kind permission of Steve Powell and Festung Guernsey

chambers, also unlined but with high curved roof sections, have been cut from the sides of the central section: the largest of the three, the main south chamber, 23.2 m long and 5.2 m wide. This central part of the system is wetter, and towards the original concrete portal the rock is more fractured and susceptible to roof falls. The roof near this western portal was sufficiently weathered to have needed timber support, but the wooden props have now long since decayed. Parts of a 0.6 m gauge railway line still in place (Fig.  8.16) show that the tunnel was connected to this transport system. Some indication of the progress of construction of these two tunnels during the time of Wehrgeologenstelle 4’s assignment to the Channel Islands is provided by a report for the period 16 March to 8 October 1942 compiled by a unit used for the work: Second Engineer Mining Company.17 The preamble notes that the company was not only helping to develop facilities on both Guernsey and Alderney, but also testing the use of tungsten carbide-tipped bits as suitable for continuous use in boring in ‘hard rock’ conditions. On Alderney, the company had been used for construction at three locations: two air raid shelters at the harbour; Ho 1 at Mannez Quarry; and Ho 2 at Fort Essex. The report notes that in terms of its geological structure the island is partly of granite, partly diorite, partly of a red coarse-grained sandstone dipping at about 20–45° to the east or SE. The tunnels at Mannez Quarry and Fort Essex lie in this sandstone. At Mannez the rock is fine-grained with a siliceous inter-granular cement, giving it a strength equivalent to granite, with  Auszug aus dem Bericht der 2. Pionier-Minier-Komp. Einsatz im Hohlgangsbau auf den Kanalinseln GUERNSEY– ALDERNEY, von 16.3–8.10.42. [7  pp.] Extract of a report stamped ‘secret’ (Geheim), and as received by the military geologist at the Inspectorate of FortificationsWest (Insp. d. L.  West. Wehrgeologe) for filing as ‘Anlage zu Az 39, Geol 10, Nr 56/43 geh’. Currently filed in the archives of the Bundeswehr Geoinformation Centre, Euskirchen. 17

318

E. P. F. Rose

Fig. 8.16  View within central region of Ho 2, at the present day, towards the original concrete portal but with entrance to the 5.4 m wide side chamber to the right. From this point the tunnel is wetter and the rock more fractured. From Gavey and Powell (2012); reproduced by kind permission of Steve Powell and Festung Guernsey

bedding and jointing poorly developed. At Fort Essex the cementation was weaker, and the rock more amenable to drilling. At the harbour excavation was through igneous rocks, predominantly dark-coloured diorite. The rock there was greatly fractured, with clay seams up to finger-width filling many of the fractures, which made drilling almost impossible. It was often necessary to abandon a hole half-finished and begin a new one. Drilling conditions were evidently better in the sandstone than the fractured and weathered igneous rocks. The two air raid shelters are presumably the two tunnels that Pantcheff (1981) records as being used for ammunition storage in the harbour area, one of which (following interpretation by Trevor Davenport) is illustrated as the bricked-up entrance to ‘Ho Crabby Harbour’ at the SE corner of the inner harbour by Gavey and Powell (2012, p. 285). Gavey and Powell (2012, p. 224) note that with the exception of Ho1 and Ho2, all of the larger tunnels on Alderney (those that have been excavated through weathered igneous rocks rather than sandstone) are now extremely dangerous because of potential roof collapse. They should not be entered.

8 Alderney

319

8.6  German Fortifications Alderney’s extensive fortifications have been comprehensively described by Partridge and Davenport (1993) and by Davenport (2003). It is now evident that geology was one of the factors taken into consideration during the construction programme. The US National Archives contain a single copy of a ‘military geology’ map of Alderney (Fig.  8.17). Hand-coloured like the Guernsey map of similar title (Fig. 6.16) and presumably produced at a similar time, the title and key to symbols and colours have been inked upon a ‘second edition’ topographical base map by a cartographic draughtsman in a style almost identical to that on the Guernsey map, and so presumably under the same general supervision. Responsibility for compilation18

Fig. 8.17  Wehrgeologische Karte. Military geological map of Alderney, original at scale of 1:10,000, prepared by Wehrgeologenstelle 4. Key (from the top down): (left column) coastal features: tidal range 8.0 m, low tide level, high tide level, coastal cliffs, beach sand (grouped together with blown sand), gravel banks (grouped with ground bearing rocky ridges and sporadic thin deposits of beach sand), crags and platforms of granite and diorite and crags and platforms of sandstone; (right column) inland features: two categories of ‘hard’ rock (granite and diorite, sandstone), residual weathered clay over granite and diorite, alluvial clay with intercalated debris and sand deposits, blown sand on granite and diorite, blown sand on sandstone, and slopes with thick colluvium. From Rose (2005a); reproduced courtesy of the US National Archives and Records Administration, College Park, MD

18

 Entwurf.

320

E. P. F. Rose

on Alderney is given as ‘Dr. Hoenes’, and for draughtsmanship19 as ‘Ludwig, Uffz.’ rather than as ‘Dr. Beschoren’ and ‘Gehrt, Gefr.’ as on the Guernsey map. Corporal Ludwig, however, had drawn Beschoren’s map of Guernsey’s potential water supply, so it seems evident that all these men functioned as members of the same team. As for the Guernsey map, features of coastal geology and geomorphology are distinguished from those of inland areas. Tidal range on Alderney is shown as 8.0 rather than 8.5 m as on Guernsey, and differences in geology and scale between the two maps result in some other minor differences. Beach sand is grouped together with blown sand rather than separately, gravel banks are grouped with ground bearing rocky ridges and sporadic thin deposits of beach sand, crags and platforms of granite and diorite and crags and platforms of sandstone are distinguished (rather than of ‘hard rock’ and ‘gneiss’ as on Guernsey). How far the map influenced Alderney’s fortification programme is not known. That the programme was, however, similar to fortification both on Jersey and Guernsey has been documented in considerable detail by Trevor Davenport (2003). A map of the principal fortifications (Fig. 8.18) reveals that Alderney was, for its size, fortified to a relatively greater degree than the other Channel Islands although in the same manner: defended by batteries of coastal artillery, infantry strongpoints and resistance nests, land mines, an anti-tank wall and anti-aircraft batteries. Alderney had five batteries of coastal artillery: three operated by the Navy, one by the Army, and one by the Army’s 319 Division. Three batteries were of medium artillery, to work in conjunction with the 155 and 203 mm calibre guns mounted near Cap de la Hague near Cherbourg, on the Contentin Peninsula of France, and the 305 mm calibre guns of the ‘Mirus’ battery on Guernsey. The three Alderney batteries were protected by machine guns, mortars, light anti-aircraft guns, flamethrowers, searchlights, barbed wire and, to some extent, minefields. The two additional batteries were of light artillery, capable of providing barrage fire to any part of the island’s shore in support of the coastal fortifications. A Naval Direction and Range Finding Tower (Fig.  8.19) was built above Mannez Quarry at the NE end of the island: five others were planned but not constructed. Alderney was also defended by 13 strongpoints and 12 resistance nests (Fig. 8.18): amongst the 54 sites illustrated by detailed plans and numerous photographs, from World War II and the present, by Davenport (2003). As for Jersey and Guernsey, the core of each of these sites typically comprised a 105 mm calibre gun: 13 of which were sited in casemates, and three more in field positions. They were supported, as on the other islands, by anti-tank guns (of 47, 50, or 75 mm calibre): 16 of these were sited around the island, in casemates or in field positions. All the sites had concrete bunkers to serve a variety of purposes. An example is Bibette Head (Strongpoint ‘Biberkopf’: Fig. 8.20), the most powerful coastal strongpoint on Alderney, on the island’s northern coast commanding the western entrance to Braye Bay, and also the small sandy Saye Bay to the east. An unusual feature of this

19

 Gezeichnet.

8 Alderney

321

Fig. 8.18  Map of principal German fortifications on Alderney as at June 1944. For coastal batteries, data as for Fig. 6.9. From Robins et  al. (2012), after Rose (2005a) and a larger-size figure within a poster illustrating Atlantic Wall defences 1940–1945: Alderney published in 1984 by Colin Partridge through Ampersand Press, Alderney. Reproduced courtesy of the Geological Society of London

Fig. 8.19  The Naval Range and Direction Finding Tower above Mannez Quarry at the NE of Alderney. Photo: E.P.F. Rose

322

E. P. F. Rose

Fig. 8.20  Plan of Strongpoint ‘Biberkopf’, at Bibette Head east of Bray Bay and so Alderney’s dominant harbour, on the central northern coast of the island. © Trevor Davenport, and reproduced by kind permission

complex is the attempt made to camouflage concrete bunkers by a cemented rock cover up to 1 m thick (Fig. 8.21). Minefields were used on Alderney as on Jersey and Guernsey (Gander 1991) to form a defensive barrier to off-beach movement, on Alderney using about 30,000 mines (see Table 4.4). Also, a 600-m-long anti-tank wall was constructed adjacent

8 Alderney

323

Fig. 8.21  View NW across the 105 mm calibre gun casemate at the northern end of Bibette Head (see Fig. 8.20), showing cemented rock cover intended to camouflage the site. Photo: E.P.F. Rose

to the island’s main beach, at Longy Bay (Fig. 8.22), to prevent armoured vehicles gaining access inland. A 100-m-long gap in the wall was blocked by three rows of mined anti-tank obstacles (illustrated by Davenport 2003, p. 89), and as the coastline is low in this area, it was mined with a mixture of anti-tank and anti-­ personnel mines. One aspect of fortification that did merit specific geologist input concerned the emplacement underground of a cable network to provide communication between the fortified sites. Before the war, there was no telephone network in Alderney, so installation of a military telephone system was a major innovation. Telephone cables needed to be buried deeply enough to minimize risk of fracture by naval or aerial bombardment. Hoenes guided burial by compiling an ‘expert opinion’,20 on 15 August 1942 (Table 8.2, item 8). Six copies were prepared, five as routinely (the two for the geologist at the Inspectorate of Land Fortification West, and one each for Fortress Engineer Command XIV on Guernsey, Fortress Engineer Staff section I/14 on Alderney, and Wehrgeologenstelle 4 itself) plus an additional one for the ‘signals’ staff on Alderney.21 Underground conditions to be anticipated during the excavation of trenches to lay the cable network were indicated upon a map attached as an annex to the text. This distinguished ground in terms of seven categories: 20 21

 Gutachten 4.  Fest. Nachr. Stab I Alderney.

324

E. P. F. Rose

Fig. 8.22  View NE from Essex Castle across Longy Bay, showing part of the major anti-tank wall constructed on Alderney, and the Naval Range and Direction Finding Tower at Mannez Quarry (Fig. 8.19) in the far distance. Photo: E.P.F. Rose

1 . Uncemented sand, thickness greater than 1.5 m. 2. Uncemented sand, thickness generally under 1.5  m, on deeply weathered bedrock. 3. Sandy loam, thickness greater than 1.5 m. 4. Sandy loam, thickness generally under 1.5 m, on deeply weathered bedrock. 5. and 6. Regions in the SW of the island, near Fort Tourgis, where the thickness of superficial deposits above the weathered bedrock was calculated to be significantly less than in the four areas previously distinguished, or where superficial deposits might be totally lacking. Excavation in these areas would be into the weathered bedrock, which included resistant veins and masses of quartz porphyry. However, the bedrock would be so strongly jointed that it could be excavated without the use of explosives. 7. The most difficult underground conditions, which lie near Berry’s Quarry and Mannez Quarry. There only a very thin regolith, some 75 mm thick if present at all, lies upon the strong unweathered sandstone. Depending on the course of the cable trenches, which had not yet been determined, it might be necessary to use explosives to facilitate excavation in this region. Warning was given that it would be necessary to line the cable trenches in places, most importantly where these crossed groundwater protection areas. Ideally, groundwater protection areas should only be crossed at their margins, and the heads

8 Alderney

325

of valleys containing springs should be avoided. Excavated trenches should be kept open for the minimum time necessary, to avoid flooding by rainfall, and any potential contamination by the workforce should be carefully avoided.

8.7  Quarrying for Raw Materials On 7 October, Hoenes completed another ‘expert opinion’:22 an explanation for the raw materials map of Alderney that accompanied it (Fig.  8.23). The penultimate ‘opinion’ to be completed by Wehrgeolenstelle 4 for Alderney, 2 months before the team’s departure from the Channel Islands, this aspect of its work evidently had relatively low priority. Alderney had many existing quarry sites, in use or disused, prior to German occupation. The apparent need was thus to assess their relative use rather than to make a survey for potential new sites. However, as on Jersey (described in Sect. 4.6) and Guernsey (Sect. 6.7), an appraisal of potential local sources of materials for construction works was a routine task assigned to German military geologists. Six copies of the report were made: one for the military geologist at the Inspectorate of Land Fortification (West) in Paris; one for the German Army’s military geological staff at Wannsee, in Berlin; and four for units based on the Channel Islands (Fortress Engineer Command XIV on Guernsey, Fortress Engineer Staff 14 on Guernsey, its section on Alderney,23 and Wehrgeologenstelle 4 itself). It comprised five pages of text that summarized descriptive features of the rocks, followed by a list and brief description of the 20 quarries shown by number on the map (Table 8.4). The text was divided into two parts: four pages on the occurrence of ‘hard’ (i.e. strong) rocks, one on the occurrence of loam, sand, and gravel. The introduction to ‘Part I’ noted that quarries numbered 1–18 on the map were all in strong rocks, which were principally of three types: the igneous rocks diorite and granite, and a sandstone that was almost a quartzite in character. Their properties could be assessed from a number of dilapidated sites that had served in earlier times as sources of raw material for road metal. For a variety of reasons, it was deemed that most of the existing quarries were unsuitable for re-opening. Number 11, for example, had been filled with water in the decades following its closure to serve as a reservoir supplying the surrounding forts. It then contained about 40,000  m3 of drinking water. Number 1 contained about 90,000 m3 of water that could be used for sanitary purposes rather than drinking. The importance of these sites for water supply precluded any future use for the extraction of building materials. In other cases, such as sites 2 and 15 and the large quarries (Berry’s and Mannez) in the sandstone, quarrying was deemed inadmissible for military reasons. (For example, Mannez Quarry was the

22 23

 Gutachten 9.  Abschnitt Gruppe I/14.

326

E. P. F. Rose

Fig. 8.23  Baustoffkarte. Map for Alderney showing quarry sites for building construction materials, original at scale of 1:10,000, prepared by Wehrgeologenstelle 4 in 1942. Key: red spot within a circle = quarries currently in operation, and equipped with plant; red spot = disused quarries, whose reinstatement was not possible without expense; blue spot = quarry flooded with water, as reservoir for the island’s supply; yellow spot within circle = sources of sand and gravel, currently in operation; brown spot = sources of building sand, currently disused. For numbered quarry sites, see Table 8.4. From Rose (2005a); reproduced by permission of the Bundesarchiv-Militärarchiv from file RH32/3027

site for construction of the island’s principal Naval Direction and Range Finding Tower:24 Fig. 8.19. It was also the site for a battery of 88 mm calibre anti-aircraft guns,25 and the entrance to the Ho 1 tunnel system then under construction: Fig. 8.14). Other quarries were so small that they could not be brought to high performance in a short time. Moreover, these small quarries had almost always been excavated into near-surface weathered rock, so that the rock quality did not meet the specifications required for building material. Following this introduction, the report described features of the diorite, granite, and sandstone in turn:

24 25

 Marinepeilstanden und Messstellen 3: MP3.  Batterie Höhe 145.

8 Alderney

327

Table 8.4  List of quarries on Alderney, as at July 1942, numbered as on Fig. 8.23 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11.

12.

13. 14. 15. 16. 17. 18. 19. 20.

Small harbour quarry, in dark medium-grained granite-gneiss. Western part filled with water to 23 m depth. Eastern part might be worked Old quarry north of 1, in dark medium-grained granite-gneiss. Rock quite highly fractured. Not recommended for bringing into production for military reasons Disused quarry in diorite. Not recommended for re-opening for military reasons Small dilapidated quarry in diorite. Contains a freshwater spring, so not recommended for re-opening Disused and dilapidated quarry, in poor quality diorite. Quarrying not possible because of its protected status as a storage area Large quarry in Harbour Bay, in fine-grained diorite. A good source for aggregate for concrete and cement, and for crushed rock Small dilapidated quarry in weathered diorite. Not recommended for re-opening Small dilapidated quarry in weathered diorite. Not recommended for re-opening Small disused quarry in dolerite. Can be brought back into production for local construction projects Small disused quarry in diorite. Not recommended for re-opening Large quarry to the east of Fort Albert, its northern part used as a major reservoir. Bringing the southern part into production is not recommended, because of poor stone quality and the risk of contamination of the water To the west of Berry’s Quarry, this site has sandstone easily worked under a thin overburden. The rock is of good quality and shows the usual uniform jointing. Field railway tracks provide access to the full-gauge railway to the port; compressed air facilities are available Berry’s Quarry, in sandstone. Not recommended for re-opening for military reasons Mannez Quarry, in sandstone. Not recommended for re-opening for military reasons Small quarry on steep slope north of Fort Essex. Not recommended for re-opening for military reasons Small disused sandstone quarry on the south coast. Re-commissioning possible for construction projects in the immediate vicinity Small disused quarry to the west of 16, in similar stone. Not recommended for re-opening Two small disused quarries in sandstone on the south coast. Re-commissioning possible for construction projects in the immediate vicinity Disused English borrow pit for fine gravel, during a time of former stone production Gravel pit in operation for the production of concrete aggregate, with conveyor belt and light railway terminal

1. Diorite formed the bedrock in the middle third and NW part of the island. Different types could be distinguished according to grain size, mineral composition and colour, but all have essentially the same properties. The rock is usable except where subject to weathering or fracturing by faulting. Grain size varies from about 1 mm in fine-grained varieties to 3 mm in medium-grained, giving the rock a dense, tough character and roughened fracture surfaces as well as hardness that makes it particularly suitable as a source of crushed rock aggregate for concrete. In most places stone suitable for paving blocks and rubble infill is also obtainable, but larger blocks for forming embankments could only be

328

E. P. F. Rose

obtained where the joint pattern is favourable. Frequent fault zones adversely affect rock properties: the rock is crushed in these fracture zones, and strong water flow through them promotes chemical weathering. The width of the fracture zones varies from a few cm to over 50 m. Within them, the rock is completely useless as a construction material. The diorite is intensively weathered throughout its area of outcrop. Within the zone of weathering, the rock has crumbled to gravel with a loam matrix, subject to rapid variation in character. The fresh rock lies at a depth of about 8 m. If new quarries were to be opened in this area, they must therefore be excavated to below this depth. 2. Granite. Three types were distinguished. (a) Dark, gneiss-like granite. This occupies an area bounded by the harbour, Fort Tourgis, Fort Clonque, Rose Farm, and the main town, St. Annes. In composition it lies between diorite and typical granite. It is coarser-grained than most of the diorite, mechanically less resistant, and more easily weathered. Where it is present in large outcrops, as along the NW coast between Fort Grosnez and Fort Clonque, it exhibits properties that make it suitable for production of crushed rock, but not aggregate for concrete. Its best texture is seen at Site 1, near the small harbour, where it could be used to produce aggregate for concrete. In parts of the island surface and the edge of Saline Bay this type of granite shows an extraordinarily deep zone of weathering. (b) Pale, mostly coarse-grained gneiss-like granite. The distribution of this rock comprises a strip of land along the southern half of the western part of the island, between Trois Vaux Bay and a point south of ‘Rond But’ (e.g. Fig. 8.13). It weathers readily to a coarse, crumbly rock with schistose partings and so is completely useless for most practical purposes. Numerous fault zones and their deep weathering also adversely affect its potential usefulness. Quarries are naturally completely lacking in this rock mass. Within the two gneiss-like granites there are intrusions of particularly strong quartz porphyry, particularly in the western part of the island between Fort Tourgis and Trois Vaux Bay and on the southern coast near ‘Springs’. These are potentially suitable as a source rock for road repairs. (c) Normal coarse-grained granite between Fort Château à L’Etoc and Bibette Head. This younger granite is of uniform coarse-grained texture (grain size 4–15 mm). In contrast to the other two granites it lacks foliation and shows a more regular jointing. Its coarseness makes it unsuitable for crushing to produce aggregate for concrete, but it can be quarried to produce blocks of cut stone. There are no quarries at present. 3. Sandstone. Two kinds with slightly different properties are distinguished:

8 Alderney

329

(a) Quartzitic sandstone. This had been excavated in the two major quarries in the NE of the island: Berry’s and Mannez (numbers 13 and 14, respectively, on Fig. 8.23). The rock is fine-grained (grain size about 0.5–3.0 mm). Since its inter-granular cement is also of quartz, it usually has a dense, quartzite-­ like character. Its hardness is so great that, like the mining works already within it, it should be worked with diamond drills. Jointing is here very regular, in contrast to that seen in the other rock types on the island. On these properties the sandstone is eminently suitable for the production of cement and crushed stone aggregate. There is no indication from prior use as to its suitability as an aggregate for concrete, but it is likely to be inferior to the diorite in this respect. (b) Coarse-grained sandstone: occurring near Fort Essex and on the south coast (sites 16, 17, and 18 on Fig. 8.23). Here the sandstone is more crumbly in character. Only when fresh can it be used for the production of broken stone. Use as aggregate for concrete was out of the question. ‘Part II’ of the report distinguished four categories within the superficial deposits overlying bedrock; 1. Loam covers almost the entire island plateau. It occurs in two situations: (a) Rubbly weathered loam, overlying the diorite and granite. Its thickness may reach 7 m; large rock fragments are common within it. (b) Sandy loess-loam, above the weathered zone, to a thickness of 0.5–2.0 m. In earlier times, the rubbly weathered loam was used for brick-making, as could be seen from two remaining derelict brick kilns. 2. Sand and gravel. Three categories were distinguished: (a) Sandy beach gravels. The main distribution area for this material lies within Saline Bay. Borings put down by the Organisation Todt distinguished material in three grain size categories: up to 1 mm; 3–8 mm quantitatively predominant; and 10–70 mm in layers or patches. Cobbles in this zone were derived from the diorite, granite-gneiss, and quartz porphyry. Loamy intercalations are very rare. The predominant sand contains a significant quantity of potentially soluble shell fragments. The thickness of this mixed sand-gravel deposit lies between 6 and 8 m. It extends for some 600 m between Forts Doyle and Tourgis, with a width of 30–50 m. The deposits had already been extracted for concrete aggregate at a site west of Fort Doyle. Deposits of a different character occur within the harbour region. Here the sands are predominantly fine-grained (grain size mostly less than 1 mm). However, layers and patches of pebbles 10–50 mm diameter also occur here. (b) Beach ridges of gravel are widespread on the north coast (Cats Bay, and between Forts les Hommeaux and Houmet-Herbé). They consist of well-­ rounded sandstone boulders of varying size (from a few centimetres to about 0.5 m in diameter).

330

E. P. F. Rose

On the south and west coasts there are gravel patches with diorite, quartz porphyry, and granite pebbles. It would be possible to process the gravel to generate aggregate, but better material was available from quarry site 6 (see Fig. 8.23) and this was therefore recommended. (c) Dune sands occupy the hinterland of Saline Bay, Harbour Bay, and Longy Bay. They also occupy the hillsides SE of Harbour Bay to a thickness of 0.5–2.0 m. Their composition is of fine sand, with grain size less than 1 mm. The accompanying map (Fig. 8.23) illustrated key information so as to be intelligible to any non-geologist, at a glance.

8.8  Air Force Geologists Like Jersey and Guernsey, Alderney had a functioning airport that was the first part of the island to be occupied by German troops: by the Air Force (Luftwaffe). However, on Alderney the airport was not put to subsequent operational use. Rather, it was deemed to be a feature likely to facilitate an airborne assault by British troops should they attempt to re-capture the island. Accordingly, the runway was blocked by trenching to prevent powered aircraft or gliders from landing, and other obstacles were erected to deter paratroopers (cf. Table 4.4). There were no aircrew or, so far as is known, ground crew, based at the airfield—and no significant attempt was made to extend it. However, 22 Flak (anti-aircraft) batteries (Fig. 8.18) were installed with some speed: four heavy batteries, with guns of 88 mm calibre; three medium batteries, with guns of 37  mm calibre; and 15 light batteries, with 20  mm calibre guns— nearly 100 guns in total (Davenport 2003). These batteries were all operated by the Luftwaffe, and were additional to the Flak weapons operated by the Army to defend the main coastal battery sites (Sect. 8.6). Davenport (2003) shows that many of these sites, most notably those for the heavy batteries, had guns sited on hard standings made from concrete (cf. Sect. 7.3.3), and there were associated bunkers for personnel and/or storage. Significant construction work was therefore associated with these sites, and with the few tunnels designated for storage use by the Luftwaffe (Table 8.3). According to a report26 of 11 May 1942, the Luftwaffe geologist Professor K.G. Schmidt (Sects. 5.2 and 5.3) inspected water supply facilities on the island of Alderney during the morning of 23 April 1942 in a group led by TKVR Dr. Hoenes of Wehrgeologenstelle 4. Other members of the group comprised Reg. Bauass. a. K. Dr. Schneider of the Luftwaffe Field Works Office for the Channel Islands27 (see

 Geologischer Bericht Nr. 148. Bericht über Wasserversorgung der Insel Alderney. Berichterstatter: Reg. Baurat Prof. Dr. Schmidt. Luftgaukommando Westfrankreich. – Verwaltung—Az.: 63 c 26 A 92—Verw. 111/7. Br. B. Nr. 5651/42 geh. O.U., den 11.5.1942. [2 pp.] Filed in the archives of the Bundeswehr Geoinformation Centre, Euskirchen. 27  Feldbauamt Kanalinseln. 26

8 Alderney

331

Sect. 7.2), Bauleiter Baumgärtner of the Organisation Todt on Alderney, and Reg. Bauinsp. Hegemann of the Luftwaffe Works Management unit on Alderney28 (Rose and Willig 2013). The heading of this report is identical to a report for tunnel systems described in Sect. 5.3, except for the new number, slight changes in file reference numbers, and omission of Schmidt’s qualification in mining engineering (presumably deemed superfluous to a report on water supply rather than tunnelling). The distribution list, allowing for slight differences in abbreviation, is exactly the same as for the other report—and the copy now preserved at Euskirchen is again one of those sent to Professor Röhrer as the military geologist at the Inspectorate of Land Fortification, West. The two categories of water sources identified in principle are (1) captured springs, arising from the zone of weathered bedrock beneath the Quaternary cover, near the heads of valleys, and (2) groundwater from the gravel in some of the valley bottoms. Infiltration galleries are noted to be potentially of more use than conventional wells, and attention is drawn to the existence of such a gallery west of Fort Essex. With reference to the intended construction of a 700-man camp to house workers of the Organisation Todt, warning is given of the need to avoid contamination of the shallow groundwater aquifer by sewage from the camp. Finally, the report concludes with a statement that Hoenes was to prepare a map in which water supply regions would be shown to guide construction of dams and other works associated with water. From this report, it is clear that the Luftwaffe had a Works Management unit29 on Alderney; that this was subordinate to a Luftwaffe Field Works Office30 for the Channel Islands as a whole, based in Guernsey; that there was close co-operation between the geologists of the Luftwaffe and the German Army; and that principal responsibility for advising on water supply problems on Alderney had been assigned to the Army’s TKVR Hoenes. It is presumed that if the Luftwaffe on Alderney needed further geotechnical advice, it would have sought this either from Hoenes or from Hans Schneider and his successor Franz Schulte as Luftwaffe geologists based in Guernsey between at least April 1942 and October 1943 (as described in Sect. 7.5).

8.9  Conclusion Thus between the summer of 1941 and that of 1944, it is known that at least five Army geologists contributed technical advice to guide fortification on Alderney, as they had done also on Jersey or Guernsey: Walther Klüpfel, Walter Wetzel, Friedrich Röhrer, Dieter Hoenes, and Bernard Beschören. Between them, they completed at least 24 technical reports for Alderney that are known to have survived to the present day (Tables 8.1, 8.2, and as shown in footnotes above), and at least eight  Bauleitung Alderney.  Bauleitung. 30  Feldbauamt. 28 29

332

E. P. F. Rose

geotechnical maps that have also survived (Table 8.5). The focus was similar to that on the other islands: primarily on water supply, but otherwise on tunnel stability, quarrying for aggregates to be used in construction, and site investigation for some of the fortification projects. Röhrer was to die soon after the end of the war, but the others returned to careers in academic geology or in geological survey (Sects. 4.8 and 6.11). Alderney was unique amongst the Channel Islands in that it was subject to an earth resistivity survey in the search for potable groundwater, by the geophysicist (serving as a ‘geologist’) Johann Kliemstein. From 1943 to 1945, as German armed forces in general retreated rather than advanced into new territory and the need for reconnaissance surveys accordingly diminished, Kliemstein was released from military service. He was appointed to the central mine survey office of a regional power company, to help update its antiquated systems (Oberste Bergbehörde 1961). After the war, he spent a short time as a technical draughtsman for Professor Dr. Posselt at the University of Leoben, maintaining his long association with that institution. This was followed by commercial employment from 1946 to 1955, as a construction manager surveying for tunnelling projects associated with new road works. Finally, on 1 October 1955, he was appointed to the Austrian mining authority31—where he led a regional office in Vienna. In his last years especially he was able to put his considerable experience in surveying and mining to most effective use. His sense of duty and reliability made him a valued employee, much respected. Kliemstein died unexpectedly, after a short illness, on 26 April 1961, aged only 62. The Air Force geologist Professor K.G. Schmidt visited Alderney in April 1942. He did so in company with the geologist, Hans Schneider, then resident on Guernsey with the Luftwaffe’s Field Works Office with responsibility for the Channel Islands as a whole. It seems likely that Schneider, and his successor by October 1943 Franz Schulte, might have provided geotechnical guidance for Alderney as part of their overall area of responsibility, but there is no known documentary record of this. Schneider returned to a geological career postwar (Sect. 7.6), and Schulte may well have returned to a career as a mining engineer (Sect. 5.8.3). Table 8.5  Thematic maps prepared by German military geologists for Alderney, with authorship where known and present location; all at scale of 1:10,000 1. 2. 3. 4. 5. 6. 7. 8.

[Untitled] geological map (elementary: no author): US National Archives [Untitled] geological map (advanced: no author): British Geological Survey Wehrgeologische BodenKarte, by Wetzel: Bundesarchiv-­Militärarchiv, file RH32v.3082 Wasserversorgungskarte, by Wetzel: Bundesarchiv-Militärarchiv, file RH32v.3082 Wasserversorgung, by Wehrgeologenstelle 4: US National Archives Baustoff- und Minier Karte, by Wetzel: Bundesarchiv-­Militärarchiv, file RH32v.3082 Baustoffkarte, by Wehrgeologenstelle 4: Bundesarchiv-­Militärarchiv, file RH32v.3027 Wehrgeologische Karte, by Wehrgeologenstelle 4: US National Archives

Sequential numbering adopted here is for convenience of reference in the text.

 Oberste Bergbehörde. Later re-named Oberste Montanbehörde, Bundesministerium für Wissenschaft und Forschung (BMWF) Minoritenplatz 5, A—1014 Wien, Austria. 31

8 Alderney

333

Their efforts overall contributed to a fortification programme on Alderney that was to transform this small island into a fortress called with some justification the ‘Gibraltar’ of the Channel (Davenport 2003).

References Cruickshank CG (1975) The German occupation of the Channel Islands. Oxford University Press, London Davenport T (2003) Festung Alderney: the German defences of Alderney. Barnes, Jersey Forty G (1999) Channel Islands at war: a German perspective. Allan Publishing, Shepperton, Surrey Gander T (1991) Land mine warfare in the Channel Islands. Channel Islands Occup Rev 19:13–31 Gavey E, Powell S (2012) German tunnels in Guernsey, Alderney & Sark. Festung Guernsey, Guernsey Geologen-Gruppe (1918) Ergänzungsheft der Wasserversorgungskarte des Gebietes der A. A. C. - Verm. Abt. 2, Geolog. Gr. [Archive document, Heringen Collection, Bundeswehr Geoinformation Centre, Euskirchen] Ginns M (1993) German tunnels in the Channel Islands. Archive book no. 7. Channel Islands Occupation Society, Jersey Ginns M (1994) The Organisation Todt and the Fortress Engineers in the Channel Islands. Archive book no. 8. Channel Islands Occupation Society, Jersey Häusler H (1995a) Die Wehrgeologie im Rahmen der Deutschen Wehrmacht und Kriegswirtschaft. In: Teil 1: Entwicklung und Organisation, vol 47. Informationen des Militärischen Geo-­ Dienstes, Vienna, pp 1–155 Häusler H (1995b) Die Wehrgeologie im Rahmen der Deutschen Wehrmacht und Kriegswirtschaft. In: Teil 2: Verzeichnis der Wehrgeologen, vol 48. Informationen des Militärischen Geo-­ Dienstes, Vienna, pp 1–119 Klüpfel W (1944) Zur Geologie der Insel Alderney. In: Bessenrodt O (ed) Die Insel Alderney: Aufsätze und Bilder. Deutsche Guernsey-Zeitung, Guernsey, pp 10–13 Oberste Bergbehörde (1961) Dipl.-Ing. Johann Kliemstein. In: Montan-Rundschau 1961 Heft 10. Montag, Wien, p 319 Pantcheff TXH (1981) Alderney, fortress island: the Germans in Alderney, 1940–1945. Phillimore, Chichester Parkinson J, Plymen GH (1929) The Channel Islands. In: Evans JW, Stubblefield CJ (eds) Handbook of the geology of Great Britain. Allen, London, pp 514–528 Partridge C, Davenport T (1993) The fortifications of Alderney. Alderney Publishers, Alderney Ramsey WG (1981) The war in the Channel Islands: then and now. Battle of Britain Prints International Limited, London Robins NS, Rose EPF (2005) Hydrogeological investigation in the Channel Islands: the important role of German military geologists in World War II. Q J Eng Geol Hydrogeol 38:351–362 Robins NS, Rose EPF, Cheney CS (2012) Basement hydrogeology and fortification of the Channel Islands: legacies of British and German military engineering. In: Rose EPF, Mather JD (eds) Military aspects of hydrogeology, vol 362. Geological Society, Special Publications, London, pp 203–222 Rose EPF (2005a) Specialist maps of the Channel Islands prepared by German military geologists during the Second World War: German expertise deployed on British terrain. Cartogr J 42:111–136 Rose EPF (2005b) Work by German military geologists on the British Channel Islands during the Second World War. Part 1: pioneering studies by Walther Klüpfel (Jersey and Alderney), Walter Wetzel (Guernsey and Alderney), and Friedrich Röhrer (Guernsey). Channel Islands Occup Rev 33:93–120

334

E. P. F. Rose

Rose EPF (2007) Work by German military geologists on the British Channel Islands during the Second World War. Part 2: Bernhard Beschoren, Dieter Hoenes, and the role of Wehrgeologenstelle 4 on Guernsey and Alderney. Channel Islands Occup Rev 35:93–114 Rose EPF (2012) Groundwater as a military resource: development of Royal Engineers Boring Sections and British military hydrogeology in World War II. In: Rose EPF, Mather JD (eds) Military aspects of hydrogeology, vol 362. Geological Society, Special Publications, London, pp 105–138 Rose EPF (2018a) Officers of 42nd Geological Section, South African Engineer Corps: geologists and geophysicists who created a unique unit that supported the British army during the Second World War. Scientia Militaria 46:19–35 Rose EPF (2018b) Military prospecting for groundwater by geology and geophysics: work by 42nd Geological Section (South African Engineer Corps) in Africa, the Middle East and the Mediterranean region during the Second World War. In: Bezuidenhout J, Smit H (eds) African military geosciences: military history and the physical environment. Sun Press, Stellenbosch, pp 131–162 Rose EPF, Willig D (2013) Work by German military geologists on the British Channel Islands during the Second World War. Part 5: work by Luftwaffe geologist professor K. G. Schmidt and Hilfsgeologe Dr. K. Diebel, for tunnelling (in general and on Jersey) and water supply (on Alderney). Channel Islands Occup Rev 41:78–101 von Bülow K, Kranz W, Sonne E (1938) Wehrgeologie. Quelle and Meyer, Leipzig Willig D, Häusler H (2012) Aspects of German military geology and groundwater development in World War II. In: Rose EPF, Mather JD (eds) Military aspects of hydrogeology, vol 362. Geological Society, Special Publications, London, pp 187–202

Chapter 9

Groundwater Investigations: German and British Nicholas S. Robins

Abstract  Major investigation of groundwater resources in the Channel Islands has taken place in three phases. The first comprised intensive activity by German armed forces in the early 1940s, faced by demographic growth with consequent increased demand for domestic water supplies plus water for mixing considerable quantities of concrete and for optimised agricultural use. The second followed recognition (in Jersey and Guernsey) during a 1976 drought that groundwater resources were finite and might not keep up with future demand. The third responded to a drought from 1989 to 1991. Phase 1 work showed that German geologists and engineers had a thorough understanding of the likely groundwater potential in weathered basement aquifers beneath a thin Quaternary cover, an understanding far in advance of British work elsewhere in the United Kingdom by that time. German investigations in Jersey focused on mapping the different hydraulic properties of the various lithological units that had been recognised earlier, whereas those in Guernsey concentrated on the availability of groundwater and its depth in the low-lying coastal regions of the island. Investigations in both islands used well and borehole inventories as a basic data source together with geological maps and reports. Work in Alderney recognised the poor potential of the shallow bedrock aquifer and recommended the use of horizontal galleries to enhance the area of the abstraction well open to the aquifer. Phase 2 studies, by British civilians, adopted much the same style of investigation, although their emphasis was towards evaluating overall resource potential as well as finding new locations for groundwater development. Phase 3 studies, led by the British Geological Survey, introduced a new investigatory tool, that of hydrochemistry and the distribution of particular chemical ions in order to help determine groundwater flow regimes. Unlike the earlier studies, the phase 3 investigations involved a three-dimensional scheme rather than the two-­ dimensional mapping framework of the earlier studies, to evaluate the overall ­productivity of the island aquifers. German maps and reports were not available N. S. Robins (*) Formerly British Geological Survey, Wallingford, Oxfordshire, UK e-mail: [email protected] © Springer Nature Switzerland AG 2020 E. P. F. Rose (ed.), German Military Geology and Fortification of the British Channel Islands During World War II, Advances in Military Geosciences, https://doi.org/10.1007/978-3-319-22768-9_9

335

336

N. S. Robins

during phase 2 and 3 studies, but the British work proves that they have stood the test of time extremely well. German work was detailed and thorough and has revealed a remarkably broad extent of technical understanding.

9.1  Introduction Island communities in general face serious problems with the maintenance of a sustainable and potable water supply. Despite their maritime climate, many islands are low-lying and potential rain-bearing clouds tend to pass over them unaffected. A significant issue, however, is demography. While effective rainfall (rainfall minus evaporation) remains essentially constant in the medium term, in most inhabited islands populations tend inevitably to increase, placing pressure on the available resource. In the Channel Islands during World War II, the situation was exacerbated by the influx of German troops and their associated workforce: providing an adequate freshwater supply was a high priority for German military engineers. In addition, weakly mineralised if not potable water (some polluted by organic matter in solution), had to be provided for enhanced agricultural use and to make up the concrete used to construct the numerous fortifications that are a feature of all the main islands (as described in Chaps. 3, 4, 5, 6, 7, and 8). Surface water is available in Jersey, Guernsey and Alderney, but it is flashy in nature, tending after significant rainfall to rise rapidly to spate conditions, because of the limited size of catchment areas. Capture of the surface water for disinfection and distribution, therefore, relies on reservoir storage, although some flow is allowed downstream to maintain the ecology of the stream. Historically, as populations increased in the Channel Islands, so the local water supply managers turned to groundwater to solve their deficits. Island groundwater in general, however, has stringent management issues associated with it, as Robins (2013) has noted: 1 . Both demand and supply of water require control. 2. Freshwater resources are generally but not always shallow and are vulnerable to contamination from on-site sanitation, poor land use practices and seawater intrusion. 3. The Ghyben–Herzberg ‘freshwater lens’ is a theoretical model that suggests that denser sea water will lie beneath lighter fresh water in many island environments. However, it cannot be applied universally and does not apply satisfactorily in fractured aquifers. An alternative analysis of the freshwater resource potential can be made using dynamic dispersion models. 4. Runoff and groundwater discharge beneath the sea shore cannot be measured directly, as they can in a mainland catchment, and this provides an additional and significant unknown variable in island groundwater investigations. 5. Impending sea-level rise elevates base levels and reduces the freshwater storage capacity, particularly in islands of low-lying topography.

9  Groundwater Investigations: German and British

337

How do these five issues respectively impact the Channel Islands? 1. The water resource available is finite and demand has to be managed as well as supply. This requires regulation and compliance from both water supply undertakings and private abstractors: not an easy task in islands where Norman Law was retained until recently. This gave the right to each islander to believe that ‘what is on or under my property is mine for the taking’. Under German occupation, such rights could be set aside. 2. The water table in bedrock in Jersey, Guernsey and Alderney is shallowest at valley bottoms, where the valley bottom stream coincides with the water table, and deepest beneath interfluves or surface water divides, and coastal cliff edges. Soil cover may be thin so that the biological protection to surface pollutants is minimal: the bedrock aquifer comprises fracture porosity that allows rapid infiltration of pollutants to the water table with little opportunity for dispersion. 3. A saline wedge generally dips beneath the foreshore under an island so that a freshwater lens sits on top of saline water. However, this does not occur on the periphery of most of the three major Channel Islands. This is due to the weakly permeable nature of the aquifers and their decreasing permeability with depth as the weathering impact declines and overburden pressure increases. Besides, most of the groundwater resource in Jersey, Guernsey and Alderney is above mean sea level with little opportunity for seawater contamination. Discharge of fresh groundwater around the coasts is generally sufficient to prevent any local ingress of sea water.1 The modern technique of dispersion modelling is available to present-day managers to help understand the resource potential, but it is data intensive and is a tool yet to be applied in the Channel Islands. 4. It is now realised that baseflow discharge from an island aquifer flows not only into streams, where it can be measured directly, but also beneath the foreshore and away to the sea. It can be measured by standard streamflow hydrograph separation techniques in the surface waters (Fig. 9.1), but it cannot be measured easily where it flows to the sea. Indirect measurement relies on Darcian slice models2 which may be in error, particularly beneath high sea cliffs where the water table is deep. 5. Projected sea-level rise will not impact significantly the groundwater resources of the Channel Islands as any increase in base level will serve to increase the elevation of the water table with little change in the aquifer volume available for groundwater storage. It may increase hydraulic gradients when higher elevation weathered material becomes saturated so that discharge rates to streams and the sea could increase making the groundwater systems more dynamic and more susceptible to years following poor winter rains.

 The head difference between groundwater and seawater controls sea water ingress. For the most part this difference is sufficiently positive to prevent sea water penetrating the shallow coastal aquifers. 2  A slice of unit thickness in which the discharge of groundwater to the sea is given by the product of the hydraulic conductivity and the hydraulic gradient within the slice. 1

338

N. S. Robins

Fig. 9.1  The concept of streamflow hydrograph separation: schematic showing likely streamflow derived from groundwater baseflow and from overland flow during an intense rainfall event. From Robins (2020) with permission

Just what comprises the weakly permeable aquifer that is common to Jersey and Guernsey and parts of Alderney? It is now known that the groundwater is stored in a weathered upper part of the crystalline basement rocks and that there is a gradual change in the degree of weathering with depth within the regolith,3 from the soil layer (pedolith) through the saprolite zone of chemical weathering and into the basal saprock zone of fractured rock with its numerous corestones. The weathering front at the base of the latter zone forms the interface between the solid rock and the overlying regolith. In the Channel Islands as elsewhere, there may be an interlayering of weathered and fractured zones, the weathering relating to zones of water table oscillation reflecting water levels during specific palaeoclimatic events and palaeohydrology levels controlled by erosion (Ollier and Pain 1996). In general, transmissivity and storativity are now known to decline as the degree of weathering diminishes with depth, passing from granular regolith into fractured rock. A conceptual model of groundwater occurrence and transport in weathered and fractured Precambrian rocks in general (Fig. 9.2) has been developed from the work of Jones and others in Tanzania in the late 1970s (Bro and Cowiconsult 1980; Jones 1985) and is now well known. It is also known that the depth and degree of weathering depend on a range of influences. For example, the basement in Africa is more weathered on an older African Erosion Surface than a Post-African Erosion Surface (Davies and Robins 2007), with weathering commonly developed to a stage beyond a granular phase to an impermeable clay grade on the older erosion surface.

 The layer of loose, unconsolidated material that covers the bedrock.

3

9  Groundwater Investigations: German and British

339

Fig. 9.2  The weathering profile of basement crystalline aquifers, based on the concept of Jones (1985). From Robins (2020) with permission, also with permission of Jeff Davies

In Jersey the upper regolith has been removed by erosion leaving, for the most part, the fractured bedrock. The regolith is still preserved in some valleys. Mourier Valley, for instance, that extends to the sea on Jersey’s northern coast displays a ghost quartz structure in weathered granite in a road cutting. Where the regolith is still present shallow wells can draw from limited groundwater storage while some go dry in late summer. Deeper boreholes are more sustainable and draw on the saturated dilated fractures which are a product of weathering (Fig. 9.3). The transmissive properties and storage potential (transmissivity4 and storativity) of the fractured bedrock support a modest aquifer capable of sustaining yields typically of between 1 and 3 l/s, occasionally greater as documented by Cheney et al. (2006). In times of water shortage the water table drops deeper into the fractured bedrock but, as transmissibility inevitably declines with the reduction in aquifer thickness, the aquifer cannot be pumped dry as borehole yields fall away in times of water scarcity. This affords some degree of self-protection to the aquifer.

 Transmissivity: the product of the permeability of an aquifer and its saturated thickness.

4

340

N. S. Robins

Fig. 9.3  Schematic illustration of the aquifer typical of Jersey showing residual regolith over weathered bedrock over bedrock. From Robins (2020) with permission, also with permission of Jeff Davies

The freshwater reserves available on Jersey, Guernsey and Alderney are replenished by direct rainfall recharge. Part of the rain that falls on the ground is returned to the atmosphere through transpiration of plants and evaporation from the soil (evapotranspiration). The remainder is termed effective rainfall, and this is divided between a component which runs off the land into the streams, called run-off, and a component which infiltrates through the soil to the water table to become groundwater. The effective rainfall replenishes the water resource so that run-off goes to the surface waters and infiltration to groundwater. Together, surface water and groundwater create the water resource: a single dynamic body of water in which surface water interacts with groundwater and groundwater with surface water. The island water supplies rely on the winter rainfall to replenish groundwater reserves and restock surface water impounded by dams (and in Guernsey also within former quarries). Natural transport of surface water and groundwater is due to gravity: streams flow down valleys. Ultimately, residual flow will reach the sea and discharge into it. Infiltration moves vertically downwards through the soil zone and the unsaturated zone of the aquifer until it reaches the water table. The direction and speed of groundwater flow depend on the prevailing slope of the water table (the hydraulic gradient) and the ability of the rock to allow water to pass through it (the hydraulic conductivity). Little of this modern-day understanding was available to the German military engineers in the early 1940s, charged with developing the available groundwater resources to their best advantage on the islands. Investigation of the groundwater resources in Jersey and Guernsey that led to a modern understanding was ultimately driven by different needs in three separate phases (Robins and Rose 2005). Phase 1

9  Groundwater Investigations: German and British

341

comprised the intensive studies carried out by German military geologists during the 1940–45 occupation. Phase 2 studies took place in the 1970s, long after British sovereignty had been restored, in response to an increasing reliance on, and demand for, groundwater coupled with water scarcity during a 1976 drought. Phase 3 studies took place between 1991 and 1998, in response to another extensive drought that lasted between 1989 and 1991. In Alderney the first phase German investigation was followed directly by third phase work. Maps and reports from the first phase studies for all three islands have only recently been rediscovered so that second and third phase investigations were all carried out in ignorance of the German achievements. The surviving German documents reveal studies on crystalline basement aquifers with thin Quaternary cover far in advance of work that had been carried out elsewhere in the British Isles by that time. German work on Jersey centered on a map portraying the different hydraulic properties of the various rock types as a two-­ dimensional model, whereas that on Guernsey concentrated on availability of potable water and depth to water in the coastal areas of the island. Both investigations depended on well and borehole inventories as primary data sources. The modern third phase studies also used well inventories, but were supported with hydrochemical evidence to determine groundwater provenance. These studies considered the groundwater flow systems as three-dimensional models using conceptualised flow systems and island-wide water balances to understand the hydrogeology of the islands. Nevertheless, the work carried out by the German geologists showed a comprehensive and advanced understanding of the hydrogeology of the islands which would, in hindsight, have been of great benefit to the second and third phase studies if it had been available when they were carried out.

9.2  German Investigations: Jersey Jersey comprises a plateau which lies at an elevation of between 60 and 120 m and which is divided by a series of north-south valleys that cut deeply into the land (Fig. 1.2). The valleys drain towards the south from the higher land in the north. The main valleys are, from west to east: St. Peter, St. Lawrence (Waterworks Valley), Les Grands Vaux and Queen’s. The northern coast is cliff-lined, whereas the eastern and western coasts include large tracts of low-lying sandy bays, the larger ones receiving significant groundwater discharge. The south coast is dominated by St. Aubin’s Bay, but there are cliff-lined bays to the west and a low-lying rocky foreshore to the east. The spring tides are large and range in height up to 12 m. The prevailing winds are from the west and SW bringing moisture from the Atlantic although rain clouds often pass over the islands without discharging. Occasional north-easterly winds blow in from continental Europe and these are generally dry. The average long-term annual island-wide rainfall is 877 mm but there is significantly less rainfall over the west and SW of the island than in the east. Mean annual potential evapotranspiration ranges between 648 and 754 mm.

342

N. S. Robins

When the German military geologist Walther Klüpfel arrived in Jersey in July 1941 (as described in Chap. 4) he had little to guide him regarding the groundwater potential available on the island. His water supply work was necessarily innovative, pioneering what was to become one of the first detailed groundwater investigations to be carried out at a practical scale applicable to the field engineer. Klüpfel did have some basic maps of the geology of the island at scales of about 1:100,000 (Noury 1886) and 1:80,000 (Bigot and De Lapparent 1894), a number of more recently published geological syntheses and regional studies (as described in Sect. 2.3), for example Mourant (1933), plus 1:50,000 and 1:25,000 topographical maps published by the British Ordnance Survey (and reprinted by German forces). Of Jersey’s groundwater occurrence and distribution there was little understanding and not much documentation. The prevailing belief of many islanders, then as now, was that groundwater flows in underground rivers from mainland France and is independent of rainfall over the island itself. Indeed, some people still believe that the groundwater not abstracted on Jersey then flows under the English Channel to supplement the Chalk aquifer in Essex; beliefs which Klüpfel would have encountered if he talked to local water diviners who were commonly if controversially regarded as experts in the detection of underground water (cf. Cheney et al. 2006). Fortunately Klüpfel, whose professional hydrogeological experience dated from World War I (Sect. 4.2.2), was not influenced by such implausible views and initiated a classic investigation of the groundwater resource potential of the island. Apart from publications on Jersey geology, there were few additional accounts that might have been of use, but notably an early reference to groundwater occurrence at Fort Regent (Jones 1840) and a description of the island-wide water supply system (Jenkins 1913). A key legacy of Walther Klüpfel’s work was a well and spring inventory that included several hundred groundwater sources (see Sect. 4.7). Work for the inventory was carried out at intervals between January 1942 and July 1943 and incorporated fields that were broadly similar to those used subsequently by Robins and Smedley (1991) in the post-war British ‘phase 3’ investigation. Klüpfel recorded: location name, grid reference, depth of well, depth to water, type of pump, likely pumping rate, water temperature and date of observation. Not all entries contained all these fields, for example, grid references (Lambert Zone 1 North extrapolated from Northern France) were recorded for only 66 of the sites, the remainder relying on the location description and field notes as the identifier. Ground elevation was not recorded, even though the available topographical maps permitted interpolation to within 2 m. These data coupled with depth to water would have allowed a groundwater level contour map to have been created to provide valuable information on the groundwater flow system and groundwater discharge to streams and discharge foci to the sea. However, Klüpfel clearly confined his investigations to a remit limiting his work to the development of groundwater sources to satisfy immediate need rather than study the overall long-term resource potential. Klüpfel made notes at each well and spring regarding seasonal variations in water level and spring flow, and whether the water was brackish to taste or whether it was clear and fresh. The groundwater chemistry became more important in the post-war surveys by which time groundwater contamination from organic pollutants, such as agricultural pesticides and fertilisers, had also become an issue.

9  Groundwater Investigations: German and British

343

The more recent phase 3 survey depended on a conceptual understanding of the groundwater flow system. Klüpfel, however, initially combined knowledge of the geology of Jersey with data from his emerging well and spring inventory to create a simple hydrogeological map (Fig. 4.28). This comprised 16 recognisable units or rock types, each with perceived differing water bearing properties (Sect. 4.7.1). The map, at a scale of 1:25,000, was prepared in January 1942, and as such it predated by several decades any similar hydrogeological mapping in comparable basement lithologies in the British Isles. Klüpfel produced two further syntheses in November 1942. The first (Fig. 4.29), a map at 1:50,000 scale (Sect. 4.7.2), which reduced the complex hydrogeological subdivisions of the original map into just three components: 1 . Inland areas with shallow groundwater. 2. Areas where groundwater is deeper but may be accessed by drilling. 3. Areas of coastal sands. The second (Fig. 4.30), a map at 1:25,000 scale (Sect. 4.7.2) shows six categories: 1 . Surface water divides. 2. Main geological boundaries. 3. Well and spring locations. 4. Coastal sand aquifers. 5. Water discharge zones to sea. 6. Channelling of groundwater flow to valleys. Thus, although Klüpfel did not make a water table elevation map he was able to make judgements regarding the groundwater flow system which proved correct when a British potentiometric map (a water table elevation contour map) was finally produced by Robins and Smedley (1998). His thinking was very much in two dimensions in the form of a plan or map; his reports did not include cross-sections. Besides, his main concern was still where does the groundwater occur, not where does the water come from and where does it go to. However, he was clearly conscious of groundwater transport and the balance between abstraction and rainfall recharge, the funnelling of groundwater along preferential valley bottom flow paths in a southerly direction, and selected zones of groundwater discharge to the sea. In a report dated October 1942 (Sect. 4.7.2), Klüpfel described work carried out at 1:25,000 scale and indicated that some of the inland areas were being mapped at 1:10,000 scale. Such larger scale maps, however, have seemingly not survived the war. Some of Klüpfel’s observations are remarkable as they show an advanced understanding of crystalline rock hydrogeology. Klüpfel’s October report ‘On the occurrence of water on the island of Jersey’ records that the island is divided into four aquifer types, three with groundwater in an intergranular matrix and one in f­ ractured bedrock. The three primary permeability pore water types are the coastal superficial sands, the valley fill alluvium and head deposits, and the porous regolith mantle over fractured bedrock. He states that deeper groundwater available in the fractures in bedrock is rarely available in significant volumes; typical discharges are about 0.3 l/s. He also observed that areas where the water table in bedrock is relatively

344

N. S. Robins

shallow are generally the most productive, and he recognised the importance of structural deformation and fracture enhancement. None of these features of basement hydrogeology was recognised formally by the British hydrogeological community until the weathering profile was formalised later by Jones (1985). The only near-contemporary work in the United Kingdom comparable with that of Klüpfel was that completed slightly later by Woodland (1946) in a series of hydrogeological maps prepared for the Chalk aquifer of East Anglia at a scale of 1:253,440 (quarter inch to the mile). These summarised borehole yield potential, depth to water, and groundwater quality. Earlier potentiometric maps had been prepared for a variety of British aquifers and attempts had been made to quantify prevailing flow patterns, notably by Walters (1929) and by Hartley (1935). However, Klüpfel’s maps of Jersey are significant as they are the only large-scale hydrogeological and groundwater potential mapping prepared for any part of the British Isles at that time. Both his work and that by Woodland predated quantitative hydrogeology, and persisted with the ideas that geology alone, unsupported by the hydraulics of aquifers, dictated groundwater occurrence, contrary to modern understanding (Downing 2004). Nevertheless, Klüpfel’s work on Jersey was that of a pioneer which went a long way towards satisfying the German military thirst for water during the Occupation. Had Klüpfel’s understanding been given as a foundation to the post-war phase 2 and phase 3 investigations that were carried out later, it would have saved considerable effort in conceptualising Jersey’s groundwater flow system.

9.3  German Investigations: Guernsey The south of the island of Guernsey is bounded by steep-sided cliffs and there is a distinct plateau area within the area of the High Parishes (Fig. 1.10). There is a marked topographical change to the north of the airport (cf. Fig. 6.9) from where the land elevation declines. There are also a number of steep-sided valleys, mostly trending NW that are the surface expressions of major NW–SE trending faults in the island bedrock. The low-lying ground in the north of the island is characterised by gently undulating topography. The NE is the flattest part of the island and includes an area of reclaimed salt marsh. Topographic gradients are more subdued to the north of St. Peter Port in Vale and St. Sampson parishes. Rainfall intensity decreases from the south to north, 838 and 837 mm respectively at the Airport and Haut Nez in the south of the island to 792 and 766 at L’Ancresse and La Turquie in the extreme NE of the island (1961–90 long-term averages). As demonstrated in Chap. 2 (Sect. 2.4), no large-scale geological maps were available for Guernsey at the start of World War II. Apparently the most accessible small-scale map was a diagram in a paper by Parkinson and Plymen (1929) which updated a similar diagrammatic map by Hill and Bonney (1884), based primarily on data from coastal outcrops. The 1929 map was used in November 1941 by the German geologist Walter Wetzel (as described in Sect. 6.2.2). He produced a water supply map at a scale of 1:25,000 (Fig. 6.3), seemingly by means of desk study in

9  Groundwater Investigations: German and British

345

Paris, which showed all the key water sources and the existing water supply mains. Not only were the surface water courses shown, but also ‘valleys with an underground water course’ were marked. This presumably recognised that groundwater flow was likely to concentrate in the valleys where it could discharge as baseflow. Wetzel’s map would have provided a foundation for preliminary site investigations by two successive German military geologists, Friedrich Röhrer (Sect. 6.3.2) and his assistant Scherer (known only from his surname: Sect. 6.3.3). These generated reports on aspects of groundwater or water supply in early 1942, notably with respect to the massive coastal artillery battery named Nina (later re-named Mirus). A priority task that faced the German military geologists of Wehrgeologenstelle 45 when they deployed on Guernsey in April 1942 for more intensive investigations (as described in Chap. 6) was that of providing a more detailed classification of the island’s rock types and more accurate delineation of their boundaries. Surviving hand-drawn geological maps (Figs. 6.14 and 6.15), like Wetzel’s drawn on topographical base maps at scale of 1:25,000, show that German geologists recognised the southern metamorphic complex and the northern igneous complex much, although not precisely, as established earlier and more firmly post-war by Roach (1966) and Roach et al. (1991). The German maps subdivided the bedrock primarily into quartzite, hornblende gabbro, augengneiss and gneiss. They also delineated some of the superficial cover: coastal sand, beach areas and widespread loess in the island’s interior. This basis would have influenced the preparation by geologists of Wehrgeologenstelle 4 late in 1942 of three groundwater prospect maps, again on topographical base maps at scale of 1:25,000 (Sect. 6.8). The first was a groundwater map prepared in September 1942 which showed well sites colour-coded in four categories of depth to groundwater and occurrence of saline or brackish water (Fig. 6.22). This was accompanied by a second map (Fig. 6.24) that illustrated groundwater conditions within the fortified coastal zone (Fig.  9.4), and was followed in November 1942 by a third map (Fig. 6.25) that indicated potential water supply prospects. This last map showed the location of all the known wells and springs and was accompanied by a report that contained an inventory of 279 wells and springs, each cross-referenced and numbered on the map. The inventory contained site location details, well type, depth to water, depth of well and comments usually regarding pumping equipment, yield and variation in yield with season. It also identified those springs in coastal areas that could be developed by digging infiltration galleries. Comparison of the German well inventory with the smaller British inventory created in the 1990s during the phase 3 investigations (Robins et al. 2002) reveals only two sites common to both surveys. This anomaly reflects the focus on drilled boreholes in the later survey (rather than shallow hand dug wells as in the former), as most of the boreholes on Guernsey were only drilled after the war. The two coincident sites, however, did not record common data as it was not possible to measure depth to water and well depth in the post-war survey due to installed pumping equipment.

 Military geology centre/team number 4.

5

346

N. S. Robins

Fig. 9.4  Groundwater occurrence and depth to water table in the coastal regions of Guernsey mapped by the German military geologists of Wehrgeologenstelle 4 in 1942: cf. Fig. 6.23; from Robins and Rose (2005), courtesy of the Geological Society of London

German work is unusual for the British Isles in general in that it includes references to infiltration galleries: these were proposed for construction in many places in Guernsey and Alderney but not to the same extent in Jersey (Fig. 4.24 illustrates a rare example from Jersey). Although few galleries were actually excavated in Guernsey during the war, much earlier wells constructed to augment water supply there (during the 1890s) had typically incorporated lateral adits to enhance their area open to the aquifer, and hence increase their yield. For example, at Forest near Guernsey airport is an old well which is 13 m deep and has three lateral adits, one 32 m long to the south, another 22 m long to the east and the third 23 m long towards the west (Hawksley 1977). Knowledge of such lateral adits may have stimulated the idea for galleries, but construction of infiltration galleries was a technique well-­ established within the German Army from at least the time of World War I (Geologen-Gruppe 1918). Data collected for the well and spring inventory were put to remarkable use. Concentrating on the coastal region of the island, as this is where most of the engineering was to take place for defence construction work, a detailed map of groundwater occurrence and groundwater status was prepared. It was initially in the form of a tracing overlay to the well and spring location map (Fig. 6.23; cf. Fig. 9.4) in the manner of a modern-day GIS data display system. In addition, the map also showed:

9  Groundwater Investigations: German and British

347

1 . Surface water courses. 2. Freshwater ponds. 3. Major springs. 4. Areas beneath which there is potable groundwater. The potable water category is subdivided between water available at depths 7 m below ground. The detail is understandably greatest along the west and northern coastal belts which are low-lying and covered by shallow Quaternary and contemporary sediments, and least for the cliff-­ lined southern and south-eastern coastal areas including the urbanised area of St. Peter Port. Such precise mapping was only possible because of the concentrated distribution of data collected for the well and spring inventory. The map, at a scale of 1: 25,000, almost certainly represents one of the most detailed pieces of hydrogeological mapping conducted in the British Isles, or indeed elsewhere, by 1942. It was the key to the siting of proposed infiltration galleries and temporary ‘Abyssinian Wells’,6 although there is no surviving evidence that such construction actually took place; also, to the construction of dry and, therefore, easily maintained, anti-tank ditches, of which many metres were successfully constructed (those at Les Mielles in Jersey were and still are notoriously wet).

9.4  German Investigations: Alderney Alderney is the smallest of the three islands in which German engineers took a significant interest. It is cliff-lined to the south and west and the main plateau area of the island falls away steeply to the north and east (cf. Fig. 1.3). Climatic conditions are broadly similar to those prevailing in Jersey and Guernsey. As for Jersey and Guernsey, geological features of the island had been studied for a long time pre-war (as described in Sect. 2.6). However, published geological maps were available at the start of the German occupation only at small scale e.g. 1:80,000 (Service de la Carte 1901), although (as for Jersey and Guernsey) there was a relatively recent and detailed if diagrammatic map (Parkinson and Plymen 1929). Several water supply maps were produced for use by German forces (as described in Chap. 8). The first was by Walter Wetzel (Sect. 8.2.2) in October 1941 (Fig. 8.4): this shows the location of wells and groups of wells, surface waters and small dams. The key to the map advises that there was no reticulation of mains water and individual sources to domestic properties were generally small. Other maps were made by Dieter Hoenes in 1942 (Sect. 8.4), notably one that shows possible locations for infiltration galleries to be constructed and of planned and existing water transfer  Perforated pipes driven into the ground to abstract groundwater at shallow depth: an American technique developed by the British Army during its campaign in Abyssinia (present-day Ethiopia) 1867-1868, adopted and widely used by the German Army in World Wars I and II. 6

348

N. S. Robins

conduits (Fig. 8.9). There was no overall attempt in Alderney to investigate the actual occurrence and distribution of groundwater, other than an unsuccessful attempt to locate a saline interface beneath the island, and no source inventory or groundwater mapping was undertaken. Investigation of the occurrence of a saline interface took place in March and April 1942. It was an attempt to try to understand the groundwater potential in the fractured rocks beneath the island and to assess if there was a salt-water layer beneath the fresh groundwater (Sect. 8.3.4). It was already understood that the freshwater-seawater interface around the larger islands of Jersey and Guernsey was steeply inclined and did not penetrate beneath the main body of the islands. One exception was the small saline wedge that occurs beneath the low-lying and partly man-made land in NE Guernsey. Electrical resistivity surveying using the Wenner array of equally spaced electrodes was, therefore, deployed on Alderney to better understand the occurrence of the groundwater beneath the island (Sect. 8.3.4). Electrical resistivity surveys had been widely used (to determine sites for boreholes to abstract groundwater) in North Africa both by German forces and those of the Allies. Both sides had developed a high respect for this technique when used in arid or semi-arid terrain. However, the resistivity surveying on temperate Alderney was not a success; electromagnetic (EM) techniques would be favoured today rather than resistivity for this kind of terrain and even then a successful outcome would not be guaranteed (Misstear et al. 2006). Water supply studies initiated by the Germans related to the major demographic problem faced by the occupying forces. The pre-war population was about 1500, and had been almost completely evacuated. However, with the arrival and expansion of the German garrison and its associated fortification workforce, latterly including many prisoners of war, the population rose at its peak to nearer 7000. The provision of adequate potable water was a priority task for military geologists of the German Army, initially Dieter Hoenes (Sect. 8.3.1) and subsequently Dieter Hoenes and his superior Bernhard Beschoren when deployed as the leaders of Wehrgeologenstelle 4 (Sect. 8.3.5). Hoenes wrote two reports relating to water and water supply in Alderney before assignment to Wehrgeologentelle 4, one dated 1 February and the second 12 February 1942. Following assignment, he generated at least three further reports, dated 26 April, 10 May and 18 May 1942. Beschoren and Hoenes jointly later prepared three particularly significant Gutachten (reports expressing expert opinion) on water supply and groundwater occurrence in the island (Sect. 8.4), dated 27 May, 15 August and 15 December 1942. Before hydrogeological studies became firmly vested in Wehrgeologenstelle 4 and so the Army, there was even a brief report on water supply by K.G. Schmidt, a uniformed geologist employed by the Air Force (Sect. 8.8). Concern developed regarding water quality in Alderney, and led to additional reports by Hoenes on 23 May and 7 July 1942 (Sect. 8.4). Analyses carried out in a field laboratory in Guernsey reported bacterial contamination in samples from three out of seven sources, and chemically aggressive acidic waters that were low in bicarbonate from a shaft near the centre of the island and from a spring. These

9  Groundwater Investigations: German and British

349

waters were reducing (possessed an active oxygen demand) and ‘traceable concentrations of ammonia’ were detected. A report dated 25 August 1942 dealt with a base camp for the workforce situated within the island’s recognised groundwater catchment area, and later, on 25 November, a final report was issued by Hoenes relating to the construction of groundwater infiltration galleries. Only one had been completed due to the inadequacy of the labour force allocated to the project. When tested, its yield was