You are on page 1of 521

Fundamentals of

Tree-Ring Research
By James H. Speer
Fundamentals of Tree-Ring Research
By: James H. Speer
Associate Professor of Geography and Geology
Indiana State University
Terre Haute, IN 47809

August 3rd, 2009

Draft: The text and figures from this book are currently in press with the University of Arizona
Press. The book, Fundametnals of Tree-Ring Research should be coming out in its final form in
Spring 2010. Please do not cite this version of the text without permission of the author.

Cover Image: This cross section is a Pinus occidentalis tree from the Dominican Republic and is
one of the most difficult pieces of wood that I have ever tried to date. This piece comes form a
site at above 3,000 meters elevation, but still has missing rings, false rings, and pinching rings
around the circumference of the sample. You can see on the cross section, my many pencil
marks that indicate the border of rings as I try to follow them around the section.

ii
Table of Contents

ACKNOWLEDGEMENTS ........................................................................................................................................ 1
PROLOQUE ................................................................................................................................................................ 2
CHAPTER 1: INTRODUCTION............................................................................................................................... 4
SOME INTERESTING APPLICATIONS OF DENDROCHRONOLOGY .................................................................................. 4
SOME BASIC PRINCIPLES AND DEFINITIONS IN DENDROCHRONOLOGY ...................................................................... 7
SUBFIELDS OF DENDROCHRONOLOGY ...................................................................................................................... 10
LIMITATIONS OF DENDROCHRONOLOGY .................................................................................................................. 10
OBJECTIVE ............................................................................................................................................................... 16
CHAPTER 2: SOME BASIC PRINCIPLES AND CONCEPTS IN DENDROCHRONOLOGY ..................... 19
INTRODUCTION ......................................................................................................................................................... 19
PRINCIPLE OF UNIFORMITARIANISM ......................................................................................................................... 20
PRINCIPLE OF CROSSDATING .................................................................................................................................... 22
PRINCIPLE OF LIMITING FACTORS ............................................................................................................................ 29
PRINCIPLE OF THE AGGREGATE TREE GROWTH MODEL........................................................................................... 31
CONCEPT OF AUTOCORRELATION............................................................................................................................. 34
CONCEPT OF THE ECOLOGICAL AMPLITUDE ............................................................................................................. 37
PRINCIPLE OF SITE SELECTION ................................................................................................................................. 39
PRINCIPLE OF REPLICATION ..................................................................................................................................... 41
CONCEPT OF STANDARDIZATION .............................................................................................................................. 41
SUMMARY ................................................................................................................................................................ 47
CHAPTER 3: HISTORY OF DENDROCHRONOLOGY .................................................................................... 48
INTRODUCTION AND THE EARLY YEARS .................................................................................................................. 48
THE 1700S AND THE 1709 FROST RING .................................................................................................................... 51
THE 1800S: TREE RINGS BECOME COMMON KNOWLEDGE ...................................................................................... 52
THE EARLY 1900S, DOUGLASS, AND HUBER ............................................................................................................ 63
THE MODERN ERA AND INTERNATIONAL ORGANIZATION ....................................................................................... 69
SUMMARY ................................................................................................................................................................ 72
CHAPTER 4: GROWTH AND STRUCTURE OF WOOD .................................................................................. 74
INTRODUCTION ......................................................................................................................................................... 74
TREE PHYSIOLOGY ................................................................................................................................................... 74
BASIC WOOD STRUCTURE ........................................................................................................................................ 79
CELL FEATURES AND TYPES..................................................................................................................................... 79
FORMS OF WOOD STRUCTURE .................................................................................................................................. 90
REACTION WOOD ..................................................................................................................................................... 98
GROWTH INITIATION AND ABSENT RINGS .............................................................................................................. 101
GROWTH THROUGHOUT THE YEAR ........................................................................................................................ 101
RING ANOMALIES ................................................................................................................................................... 105
SUMMARY .............................................................................................................................................................. 116
CHAPTER 5: FIELD AND LABORATORY METHODS .................................................................................. 117
INTRODUCTION ....................................................................................................................................................... 117
GEAR ...................................................................................................................................................................... 117
RANDOM VERSUS TARGETED SAMPLING ............................................................................................................... 122
PLOTS, TRANSECTS, OR TARGETED SAMPLING ....................................................................................................... 124
CORING A TREE ...................................................................................................................................................... 125
Testing for a Compressed Core ......................................................................................................................... 129
Taking and packaging a core ............................................................................................................................. 130
Removing an increment borer from the tree ...................................................................................................... 135

iii
CLEANING AN INCREMENT BORER ......................................................................................................................... 135
SHARPENING AN INCREMENT BORER...................................................................................................................... 137
Spanish Windlass Technique for Retrieving a Stuck Borer ............................................................................... 140
LABORATORY METHODS ........................................................................................................................................ 142
Preparing Core Samples .................................................................................................................................... 142
Preparing Cross Sections ................................................................................................................................... 152
ANALYSIS OF CORES AND CROSS SECTIONS ........................................................................................................... 153
Skeleton Plotting................................................................................................................................................ 156
List Method........................................................................................................................................................ 162
Memorization Method ....................................................................................................................................... 163
Measuring Methods ........................................................................................................................................... 164
WORK TIME DISTRIBUTION .................................................................................................................................... 169
CHAPTER 6: COMPUTER PROGRAMS AND STATISTICAL METHODS ................................................ 171
INTRODUCTION ....................................................................................................................................................... 171
STATISTICS IN DENDROCHRONOLOGY .................................................................................................................... 172
Series Intercorrelation ........................................................................................................................................ 172
Mean Sensitivity ................................................................................................................................................ 173
Gleichläufigkeit – Sign Test .............................................................................................................................. 174
Rbar ................................................................................................................................................................... 175
Expressed Population Signal (EPS) ................................................................................................................... 175
Subsample Signal Strength (SSS) ...................................................................................................................... 177
MEASURING PROGRAMS ......................................................................................................................................... 179
Measure J2X ...................................................................................................................................................... 179
DPL ....................................................................................................................................................................... 183
FMT ....................................................................................................................................................................... 183
COFECHA ............................................................................................................................................................ 186
Keystroke Tutorial of COFECHA ..................................................................................................................... 190
Reading the Output of COFECHA .................................................................................................................... 196
Conclusions from COFECHA ........................................................................................................................... 208
EDRM ................................................................................................................................................................... 210
ARSTAN ............................................................................................................................................................... 210
Keystroke Tutorial for ARSTAN for Windows................................................................................................. 215
Reading the Output of ARSTAN ....................................................................................................................... 220
Regional Curve Standardization (RCS) ............................................................................................................. 222
YUX ...................................................................................................................................................................... 223
CLIMATE ANALYSIS PACKAGES ............................................................................................................................. 223
PRECON ........................................................................................................................................................... 225
DENDROCLIM2002......................................................................................................................................... 225
OUTBREAK ......................................................................................................................................................... 228
SPECTRAL ANALYSIS.............................................................................................................................................. 231
EVENT .................................................................................................................................................................. 231
CONCLUSION .......................................................................................................................................................... 232
CHAPTER 7: DENDROARCHAEOLOGY ......................................................................................................... 234
INTRODUCTION ....................................................................................................................................................... 234
ARCHAEOLOGICAL METHODS ................................................................................................................................ 243
Sample Collection.............................................................................................................................................. 243
CHRONOLOGIES USED IN DENDROARCHAEOLOGY ................................................................................................. 253
APPLICATIONS OF DENDROCHRONOLOGY TO ARCHAEOLOGY................................................................................ 253
Construction dates ............................................................................................................................................. 255
Dating Artifacts ................................................................................................................................................. 259
Climate Reconstructions .................................................................................................................................... 260
Ecological Reconstructions and Anthropogenic Ecology .................................................................................. 263
FUTURE OF DENDROARCHAEOLOGY....................................................................................................................... 269

iv
CHAPTER 8: DENDROCLIMATOLOGY .......................................................................................................... 270
INTRODUCTION ....................................................................................................................................................... 270
METHODS FOR DENDROCLIMATOLOGY .................................................................................................................. 273
APPLICATIONS OF DENDROCLIMATOLOGY ............................................................................................................. 279
Climate Indices .................................................................................................................................................. 283
Climatic Gradient Studies .................................................................................................................................. 285
Latitudinal Gradient ........................................................................................................................................... 285
Treeline Studies ................................................................................................................................................. 286
Dendrohydrology: Water Table Height and Flood Events ................................................................................ 288
SEGMENT-LENGTH CURSE ..................................................................................................................................... 293
ARCHAEOLOGICAL USES OF CLIMATE RECONSTRUCTIONS .................................................................................... 294
USE OF CLIMATE RECONSTRUCTIONS FOR FUTURE PREDICTION ............................................................................ 295
CHAPTER 9: DENDROECOLOGY ..................................................................................................................... 296
INTRODUCTION ....................................................................................................................................................... 296
METHODS FOR DENDROECOLOGY .......................................................................................................................... 297
Stand-Age Structure........................................................................................................................................... 297
Ring Width Analysis.......................................................................................................................................... 299
Tree Scars .......................................................................................................................................................... 299
Basal Area Increment ........................................................................................................................................ 300
APPLICATIONS OF DENDROECOLOGY ..................................................................................................................... 300
Gap Phase Dynamics ......................................................................................................................................... 300
Forest Productivity and Succession ................................................................................................................... 301
Old Forests......................................................................................................................................................... 304
Dendropyrochronology ...................................................................................................................................... 306
Dendroentomology ............................................................................................................................................ 322
Wildlife Populations and Herbivory .................................................................................................................. 337
Distributional Limits of Species ........................................................................................................................ 338
Treeline and Subartic Studies ............................................................................................................................ 339
Interactions of Multiple Disturbances................................................................................................................ 340
Other Applications in Dendroecology ............................................................................................................... 342
CONCLUSION .......................................................................................................................................................... 343
CHAPTER 10: DENDROGEOMORPHOLOGY ................................................................................................ 344
INTRODUCTION ....................................................................................................................................................... 344
SOURCES OF INFORMATION .................................................................................................................................... 348
Reaction wood ................................................................................................................................................... 348
Death dates ........................................................................................................................................................ 349
Establishment dates ........................................................................................................................................... 351
Wound Events.................................................................................................................................................... 351
Coarse Woody Debris (CWD) ........................................................................................................................... 352
Roots .................................................................................................................................................................. 353
SUBFIELDS OF DENDROGEOMORPHOLOGY ............................................................................................................. 353
Dendrovolcanology ........................................................................................................................................... 353
Dendroglaciology .............................................................................................................................................. 356
Mass Movement ................................................................................................................................................ 358
Dendroseismology: Plate Boundaries, Faults, and Earthquakes ........................................................................ 360
LIMITATIONS IN DENDROGEOMORPHOLOGY .......................................................................................................... 362
CONCLUSIONS ........................................................................................................................................................ 363
CHAPTER 11: DENDROCHEMISTRY............................................................................................................... 364
INTRODUCTION ....................................................................................................................................................... 364
DENDROCHEMISTRY ........................................................................................... ERROR! BOOKMARK NOT DEFINED.
Methods of Elemental Analysis ......................................................................................................................... 365
Conclusions on Dendrochemistry ...................................................................................................................... 372

v
RADIOMETRIC ISOTOPES......................................................................................................................................... 372
STABLE ISOTOPES ................................................................................................................................................... 373
Limitations ......................................................................................................................................................... 379
Standard Procedures .......................................................................................................................................... 380
Fractionation ...................................................................................................................................................... 382
Other Usable Elements ...................................................................................................................................... 388
CONCLUSIONS ........................................................................................................................................................ 389
CHAPTER 12: FRONTIERS IN DENDROCHRONOLOGY ............................................................................ 390
INTRODUCTION ....................................................................................................................................................... 390
STABLE ISOTOPES ................................................................................................................................................... 391
MULTIPLE PROXIES ................................................................................................................................................ 393
IMAGE ANALYSIS OF REFLECTED LIGHT ................................................................................................................ 394
WOOD ANATOMY ................................................................................................................................................... 395
TROPICAL DENDROCHRONOLOGY .......................................................................................................................... 396
UNIQUE ENVIRONMENTS ........................................................................................................................................ 400
SCLEROCHRONOLOGY ............................................................................................................................................ 401
CONCLUSION .......................................................................................................................................................... 402
REFERENCES ........................................................................................................................................................ 403
APPENDIX A: TREE AND SHRUB SPECIES THAT HAVE BEEN USED BY
DENDROCHRONOLOGISTS .............................................................................................................................. 479
APPENDIX B: AGE OF THE OLDEST TREES PER SPECIES. ..................................................................... 494
APPENDIX C: PITH INDICATORS .................................................................................................................... 504
APPENDIX D: FIELD NOTE CARDS ................................................................................................................. 505
INTRODUCTION ....................................................................................................................................................... 505
APPENDIX E: WEB RESOURCES. ..................................................................................................................... 508

vi
Table of Figures

Figure 1. 1 The Messiah Violin. 6


Figure 1. 2 A cross section of an ash (Fraxinus sp.) tree 13
Figure 2. 1 Crossdating. 24
Figure 2. 2 A) Photograph of a single skeleton plot showing the beginning and end arrows that represent the
inner most date and bark dates respectively. 26
Figure 2. 3 A picture of marker rings recorded using the list method. 28
Figure 2. 4 Limiting Factors. 30
Figure 2. 7 Needle retention in bristlecone pine at upper treeline related to summer temperature 36
Figure 2. 8 Distribution of black gum. 38
Figure 2. 9 Site selection. 40
Figure 2. 10 Sampling scale. 42
Figure 2. 11 Standardization. 43
Figure 2. 12 Standardization with a negative exponential curve and cubic smoothing splines. 45
Figure 3. 1 Leonardo da Vinci (1452-1519) 50
Figure 3. 2 Alexander Catlin Twining (1801-1884) 53
Figure 3. 3 Charles Babbage (1791-1871) 56
Figure 3. 4 A.E. Douglass (1867-1962) 64
Figure 3. 5 A.E. Douglass and his students in 1946. 67
Figure 4. 1 A juniper from Jordan showing lobate growth demonstrating poor circuit uniformity where the
rings pinch out around the circumference of the section. 77
Figure 4. 2 Pith characteristics 78
Figure 4. 3 Planes of wood structure. 80
Figure 4. 4 Gymnosperm wood examples. 82
Figure 4. 5 Gymnosperm cells types. 83
Figure 4. 6 Gymnosperm wood structure. 84
Figure 4. 7 Resin ducts in a gymnosperm. 86
Figure 4. 8 Angiosperm cells types. Angiosperms have more complex cell types which are vessel elements,
tracheids, fibers, parenchyma, ray cells, and pits ( 87
Figure 4. 9 Dicotyledonous angiosperm wood structure. 88
Figure 4. 10 Wood rays in an oak. 89
Figure 4. 11 Gymnosperm versus Angiosperm wood types. 91
Figure 4. 12 Classification of ring porosity. 92
Figure 4. 13 Relative size of hardwood cells and wall thickness in ring porous species. 93
Figure 4. 14 Pore arrangement in angiosperms. 94
Figure 4. 15 Examples of ring porous woods. 95
Figure 4. 16 Examples of semi-ring porous woods. 96

vii
Figure 4. 17 Examples of diffuse porous woods. 97
Figure 4. 18 Reaction wood. 99
Figure 4. 19 Microscopic cross sectional view of compression wood in a conifer (right image) 100
Figure 4. 20 The Auxin model of tree growth 102
Figure 4. 21 Three dimensional ring production. 103
Figure 4. 23 Ring anomalies in Pinus occidentalis. 106
Figure 4. 24 Suppressed ring porous wood growth. 110
Figure 4. 25 Offset of wood growth across rays. 111
Figure 4. 26 Frost Ring. 113
Figure 4. 27 Aphid damage to cells of a red maple (Acer rubrum) tree. 114
Figure 4. 28 Fire scar in ponderosa pine. 115
Figure 5. 1 Starting a borer. 128
Figure 5. 2 Measuring for compressed wood in an increment borer. 131
Figure 5. 3 Coring a tree. 132
Figure 5. 4 Tip of an increment borer. 133
Figure 5. 5 Extracting a core. 134
Figure 5. 6 Extracting the increment borer. 136
Figure 5. 7 Spanish windlass. 141
Figure 5. 9 Core orientation. 145
Figure 5. 10 Mounting cores. 146
Figure 5. 11 Untwisting cores. 148
Figure 5. 12 Sanding belts. 149
Figure 5. 13 Sanding cores. 151
Figure 5. 14 Cleaning a sander belt. 154
Figure 5. 15 Marking the wood. 155
Figure 5. 16 Making a skeleton plot from a sample of wood. 158
Figure 5. 17 The Velmex Measuring System. 166
Figure 6. 1 An example of calculating the Gleichläufigkeit value 176
Figure 6. 2 A) Running rbar and B) Running EPS analysis for the Newberry Crater Lava Flow Ponderosa Pine
Chronology 178
Figure 6. 3 Initializing a new series in MeasureJ2X. 181
Figure 6. 4 Measuring view in MeasureJ2X. 181
Figure 6. 5 The Dendrochronology Program Library (DPL) version 1.24p 184
Figure 6. 6 The Dendrochronology Program Library (DPL) version 6.07p 184
Figure 6. 7 Formatting options in FMT. 185
Figure 6. 8 Twenty three separate functions that can be performed on data in the FMT program. 187
Figure 6. 9 Introductory screen of COFECHA in a DOS command line box. 191

viii
Figure 6. 10 Command line driven DOS box for COFECHA 193
Figure 6. 11 The end of the information that flashes on the screen while COFECHA runs. 195
Figure 6. 12 COFECHA output page 1. 197
Figure 6. 13 Part 2 of COFECHA lists and graphically depicts the length of each core. 199
Figure 6. 14 COFECHA Part 3 shows the index values and sample depth for the master chronology. 200
Figure 6. 15 COFECHA Part 4 provides a graphical representation of the master chronology. 202
Figure 6. 16 COFECHA Part 5 shows the correlation of each 50-year segment to the master. 203
Figure 6. 17 COFECHA Part 6 provides core-level analysis of how well each core dates against the master 205
Figure 6. 18 A second page from COFECHA Part 6 showing when a core has a missing ring. 207
Figure 6. 19 COFECHA Part 7 provides a table of the descriptive statistics for each core 209
Figure 6. 20 EDRM showing the options for editing a file. 211
Figure 6. 21 Standardization and tree-level index series. 213
Figure 6. 22 Examples of four tree ring chronologies that have been standardized using a 15-year cubic
smoothing spline. 214
Figure 6. 23 Comparison of standardization with a negative exponential curve (A) versus a 100 year cubic
smoothing spline (B) on a 600-year chronology. 217
Figure 6. 24 Main menu for ARSTAN. 219
Figure 6. 25 Master chronologies for the Mokst Butte Lava Flow Ponderosa Pine 221
Figure 6. 27 The opening page to PRECON. 226
Figure 6. 28 Correlation results comparing tree rings to climate in DENDROCLIM2002 227
Figure 6. 29 The opening page of the program OUTBREAK. 229
Figure 6. 30 A superposed epoch analysis showing the growth departure in pin oak (Quercus palustris) ring
growth associated with periodical cicada emergences. 233
Figure 7. 1 A.E. Douglass (1867-1962) coring a ponderosa pine. 235
Figure 7. 2 Clark Wissler ( 236
Figure 7. 3 Neil Merton Judd (1887-1976). 237
Figure 7. 4 Earl Halstead Morris (1889-1957). 239
Figure 7. 5 Lyndon Lane Hargrave ( 240
Figure 7. 6 Emil W. Haury (1904-1992) examining a buried beam at Pinedale Ruin, Arizona during the Third
Beam Expedition, 1929 241
Figure 7. 7 Sampling HH-39. 242
Figure 7. 8 An Archaeological borer. 244
Figure 7. 9 Cores can be taken from window lintels. 246
Figure 7. 10 Primary and secondary beams. 248
Figure 7. 11 Cross section of a primary beam. 249
Figure 7. 12 Charcoal samples can also be used for archaeological dating. 250
Figure 7. 13 A log cabin from the southern Appalachian Mountains. 257
Figure 7. 14 Cross section of a beam from a log cabin. 258

ix
Figure 7. 15 Rings on the face of a cello. 261
Figure 7. 16 The Karr-Koussevitzky double bass marked up for measurement 262
Figure 7. 17 Peel bark tree. 266
Figure 7. 18 White oak (Quercus alba) regional mast reconstruction from the southern Appalachian Region. 268
Figure 8. 1 Tree-ring climate reconstruction for the past 1,000 years. 271
Figure 8. 2 Old preserved wood on a lava flow in Oregon. 274
Figure 8. 3 Bristlecone pine trees in Methuselah Grove. 287
Figure 8. 4 Old Pinus occidentalis growing on a high elevation site in the Dominican Republic. 289
Figure 8. 5 The age of a delta or any sedimentary deposit can be determined from trees growing on that
sediment. 290
Figure 8. 6 Flood events can damage trees in many ways, providing dendrochronologists with different
approaches to reconstruct flood activity 292
Figure 9. 1 A ponderosa pine stand in Oregon that has received multiple thinning and prescribed fire
treatments. 307
Figure 9. 2 A catface can be a huge scar when it occurs in giant sequoia. 308
Figure 9. 3 Stand replacing fire in Pinus sylvestris. 309
Figure 9. 4 A catface scar on a living ponderosa pine tree with a partial section removed from the left base of
the tree. 311
Figure 9. 5 We take partial sections from living trees to get a complete history of fire through the modern era,
but leave the tree standing and healthy. 313
Figure 9. 6 A partial section from ponderosa pine with a close up showing the fire scar dates. 314
Figure 9. 7 A fire scar is a three dimensional wound where the living cambium meets the dead cambium. 317
Figure 9. 8 Fire history data can be collected on multiple spatial scales to understand the driving factors of this
natural disturbance. 319
Figure 9. 9 A fire history chart for a network of 55 site-level chronologies 320
Figure 9. 10 A 622-year pandora moth reconstruction from south-central Oregon. 328
Figure 9. 11 A ponderosa pine forest denuded of needles by pandora moth 329
Figure 9. 12 A tree ring signature has been identified for pandora moth 330
Figure 9. 13 Insect outbreaks often affect many trees both on an individual site and on multiple sites. 332
Figure 9. 14 Here, the author is taking samples from a 600 year old ponderosa pine tree for a stem analysis to
examine the wood volume lost due to pandora moth defoliation. 334
Figure 9. 15 Diagram showing how samples taken every 3 meters up a tree can be used to calculate wood
volume for the whole tree 335
Figure 9. 16 Examples of four trees showing changes in wood volume with height of the tree. 336
Figure 10. 1 Coarse woody debris is composed of logs that fall in the forest. 350
Figure 10. 2 Wood structure for roots of Larix decidua at four different depths in the soil 354
Figure 10. 3 A glacier may kill trees as it advances and incorporate those trees in the till. 357
Figure 10. 4 Frequent rockfall down a landslide shoot may accumulate a large amount of sediment 359
Figure 10. 5 These trees around Yellowstone Lake have been subject to gradual soil erosion; adventitious roots
grew as the soil was slowly removed from the site, enabling many of the trees to survive 361

x
Figure 11. 1 A dendrochemistry sample in a core clamp 367
Figure 11. 2 Controls on the isotopic signature in plants (from Anderson et al. 2003). 378
Figure 11. 3 An example of fractionation of 18O/16O from sea water to an inland site with two rain events. 384
Figure 11. 4 Factors that control fractionation in a pine tree 385
Figure 11. 5 Feedback mechanisms affecting fractionation in an oak tree 387
Figure 12. 1 The National Climatic Data Center (NCDC) runs the World Data Center (WDC) for
paleoclimatology which houses the tree-ring chronologies of the International Tree Ring Databank
(ITRDB) 398 

xi
Table of Tables
TABLE 1. 1 SUBFIELD OF DENDROCHRONOLOGY ......................................................................................................... 11
TABLE 2. 1 THE FREQUENCY OF THE VARIANCE THAT REMAINS USING DIFFERENT CUBIC SMOOTHING SPLINES. .......... 46
TABLE 3. 1 THE EARLY DENDROCHRONOLOGISTS SORTED BY THE DATES THAT THEY USED TREE RINGS. .................... 49
TABLE 5. 1 BASIC CHECKLIST OF GEAR NEEDED FOR DENDROCHRONOLOGICAL SAMPLING. ....................................... 119
TABLE 5. 2 AVERAGE WORK TIME IN HOURS TO COLLECT, PROCESS, AND BUILD A CHRONOLOGY THAT IS FROM
200-400 YEARS IN LENGTH FROM 20 TREES (MODIFIED FROM FRITTS 1976). ............................................ 170
TABLE 7.1 SYMBOLS USED TO MARK ARCHAEOLOGICAL SAMPLES ............................................................................. 252
TABLE 7. 2 LONG-TERM CHRONOLOGIES FROM AROUND THE WORLD. ....................................................................... 254
TABLE 10. 1 GEOMORPHIC EVENTS AND HOW THEY CAN BE RECONSTRUCTED USING TREE RINGS. ............................ 345
TABLE 11.1 A SUMMARY OF THE TREE-RING ISOTOPE STUDIES FOR PALEOENVIRONMENTAL RESEARCH ................... 374
 

xii
Acknowledgements

I would like to thank the many dendrochronologists that have given input on this text. Without

their help, this work would be much diminished. Rex Adams, Ed Cook, Lori Daniels, Jeff Dean,

Dieter Eckstein, Esther Fichtler, Hal Fritts, Holger Gaertner, Henri Grissino-Mayer, Richard

Guyette, Tom Harlan, Steve Leavitt, Kathy Lewis, Dave Meko, Steve Nash, Bill Patterson, Fritts

Schweingruber, Greg Wiles, Connie Woodhouse, Tom Yanosky, and Qi-bin Zhang. I would

especially like to thank my wife Karla Hansen-Speer who edited this text multiple times,

provided line drawings for some of the graphics, helped with all of the issues involved in

bringing a book to print, and was also my model for some of the photographs. The National

Science Foundation (NSF) provided some monetary support for the North American

Dendroecological Fieldweek through grant GRS # 0549997.

1
Proloque

I envisioned writing this book since I was an undergraduate student at the University of Arizona

in the Laboratory of Tree Ring Research, when in my introductory dendrochronology class

(taught by Tom Swetnam), we used a photo copy of a book written by Hal Fritts that was no

longer in print. That book was the famous “Tree Ring and Climate” but during the 1990s, the

only way to acquire a new copy of the book was from a photo copy directly from the author.

Thankfully, Blackburn Press later reprinted the book in its entirety in 2001, and that book is now

again available to new students in dendrochronology. Another book that was often suggested for

the novice dendrochronologist was “An Introduction to Tree-Ring Dating” Marvin Stokes and

Terah Smiley. The first book was written at a high level with the explicit focus of

dendroclimatology while the second book was a basic introduction that specifically used a

dendroarchaeological project to explain the techniques of dendrochronology. I have written this

book as a basic introduction to the breadth of dendrochronology including some of the principles

and physiological background that one needs to conduct dendrochronology. The second half of

the book has individual chapters dedicated to each of the sub-disciplines of dendrochronology,

providing a basic bibliography for one to start dendrochronological research in any of the

applications within the field.

I further saw a need for this book in teaching and organizing the North American

Dendroecological Fieldweek (NADEF) for the past nine years along with teaching my own

dendrochronology class at the university level for the past seven years. I found that the students

consistently asked the same questions about the field and I hope that this book answers those

questions. In teaching at NADEF, I realized that every laboratory has its own perspective on the

2
field, although the Tucson Lab has wide-reaching influence because so many people spend some

time at that laboratory for aspects of their training in the field. My own training in

dendrochronology came from four years at the Laboratory of Tree-Ring Research and therefore

my perspective is from the Tucson Lab in Arizona. I have contacted researchers in other

countries to try to ensure the inclusion of the best literature from around the world, but

admittedly my perspective is based on my own research in North America.

I hope that you find this book useful as a reference and as a primary starting point for work in

dendrochronology. It is the result of four years of dedicated reading and writing along with the

input from 30 other dendrochronologists through reviews of different stages of the book.

3
Chapter 1: Introduction

Dendrochronology is one of the most important environmental recording techniques for a variety

of natural environmental processes and a monitor for human caused changes to the environment

such as pollution and contamination. The word dendrochronology has its roots in Greek:

“dendro” means tree and “chronology” means the study of time. Dendrochronology examines

events through time that are recorded in the tree-ring structure or can be dated by tree rings.

Because the tree becomes the instrument for environmental monitoring, it provides a long-term

bioindicator that extends for the lifetime of the tree. Dendrochronology can be applied to very

old trees to provide long-term records of past temperature, rainfall, fire, insect outbreaks,

landslides, hurricanes, and ice storms to name only a few applications. Wood from dead trees

can also be used to extend the chronology of tree rings further back in time. Trees record any

environmental factor that directly or indirectly limits a process that affects the growth of ring

structures from one season to the next making them a useful monitor for a variety of events.

Some Interesting Applications of Dendrochronology

The discipline of dendrochronology is used to mark time or record environmental variability in

the structure of the wood from trees growing in seasonal climates, such as in the mid-latitudes,

high latitudes, and some tropical trees growing in environments with a pronounced wet or dry

season. Because many different environmental variables can affect tree growth, different records

can be gained from a variety of tree species on a site and on a variety of sites in a region.

Dendrochronologists have been able to develop interesting records that contribute too many

areas of modern culture, from forensic science to boundary disputes. For example, tree rings

4
were used as forensic evidence in a murder case by dating the age of a root that grew over a

buried corpse (Thomas Harlan personal communication). Sellards et al. (1923) used tree rings to

settle a boundary dispute between the states of Oklahoma and Texas along the Red River. A

carving on an aspen (Populus tremuloides) tree stem supposedly made by Ted Bundy (the

infamous serial killer) was dated to 1976, a time he was reported to be in that region, thus

causing investigators to intensify their search for Ted Bundy’s victims in this area (Thomas

Swetnam personal communication). This was an unusual use of dendrochronology because the

carving was only in the bark of the tree and not in the wood. Aspens produce rings in the bark

(although not as continuously as in the wood) which were counted to establish the date.

Tree rings have been used to help elucidate strange atmospheric events such as the Tunguska

Event that occurred on June 30th, 1908 (Vaganov et al. 2004). One of the main hypotheses of

what caused this event that flattened 80 million trees over 2,150 square kilometers in Siberia is

the arrival of a large meteoroid that disintegrated in the atmosphere from 5-10 kilometers above

the surface of the Earth. This is the largest impact event in the written history of the Earth.

Vaganov et al. (2004) examined tree rings at the time of the impact and found that the cells

growing during the end of 1908 were deformed. They could conclude that this was likely caused

by a forceful impact on the trees and suggested an upward adjustment of previous estimates to

the amount of force that was exerted on the trees from this event.

One of the more remarkable stories in dendrochronology comes from attempts to date the

Stradivari violin called the Messiah which has a label date of 1716 (Figure 1.1). The Messiah

violin is considered one of Antonio Stradivari’s crowning achievements. It is a well-preserved

5
Figure 1. 1 The Messiah Violin. Any object that is made of wood and has enough rings in a
sensitive series can potentially be dated using tree rings. In an unusual example, the Stradivarius
Messiah Violin was dated using dendrochronology (photo from Topham and McCormick 2001).
instrument that has a distinct red hue to its finish. This instrument was valued between 10 and

6
20 million dollars if it could be authenticated as the true “Messiah” violin, but if it was made by

a copyist in the 1800s, it would be worth far less. This instrument is housed in the Ashmolean

Museum in Oxford, England. Initial tree-ring dating on the Messiah suggested that the

instrument was the original Stradivari (Topham and McCormick 1997, 1998, 2001), but others

(including one dendrochronologist) claimed the violin was made after Stradivari died in 1737

(Pollens 1999, 2001). To settle the controversy, Dr. Henri Grissino-Mayer of the University of

Tennessee was asked to assemble a team of experts and examine the rings a second time in an

attempt to date the violin.

It should be noted that dendrochronology cannot conclusively demonstrate that the instrument

was made by Stradivari. Dendrochronology, however, can be used to disprove the possibility of

it being a Stradivari instrument by finding growth rings in the wood of the instrument that post-

date Stradivari’s death in 1737. The “Messiah” violin was made from a spruce tree and had 120

rings showing on the top of the instrument. Grissino-Mayer et al. (2002, 2003, 2004) worked at

dating this violin against European reference tree-ring chronologies and against chronologies

developed from other known Stradivari instruments. They demonstrated that the last rings in the

instrument dated to A.D. 1687, which was consistent with two other instruments made by

Stradivari: the “Archinto” (dating to 1686) and the “Kux/Castelbarco” violas (dating to 1684).

Some Basic Principles and Definitions in Dendrochronology

Many proxy records (alternate sources of information from natural phenomena) of climate and

the environment exist, such as pollen, ice cores, lake varves (annually layered sediment), coral

layers, and speleothems (calcium carbonate dripstone from caves)(Bradley 1999), but

7
dendrochronology provides the most reliable dating with the highest accuracy and precision of

any of these paleorecords. The practice of crossdating (matching the pattern of wide and narrow

rings to demonstrate dating between trees), which was developed by A.E. Douglass in the early

1900s (Douglass 1909, 1917, 1920, 1921, 1929, 1941), is now being used for some of the other

proxy records that form regular (sometimes annual) increments, such as ice cores, corals, rings in

clam shells, and otoliths (the bony structure in the ears of fish) as a check on the dating of those

records (Black et al. 2005).

The science of dendrochronology has a few basic principles and concepts that have been

repeatedly demonstrated by scientific evidence from multiple disciplines and experimentally

supported by dendrochronological research. These principles are the subject of Chapter 2, but I

summarize some of them here because it is hard to discuss dendrochronology without the use of

some of these terms. The main principle of crossdating suggests that variation in ring width is

driven by limited environmental factors needed for growth; matching these narrow rings

provides the quality control required by dendrochronologists, allowing the assertion of annual

resolution and being able to provide exact calendar years for every tree ring in a sample. This

principle is strengthened by the concept of replication which states that reliable dates must be

supported by enough samples to assure the probability of being in error is sufficiently minute.

For example, for most sites in the southwestern United States, 20 overlapping tree records

(sample depth) are usually sufficient for a reliable chronology (a site-level representation of

tree growth). Good chronologies have been developed with as few as 10 trees sampled on sites

with a consistent site-level chronology and many chronologies have been developed with more

than 100 tree samples. By sampling two cores from each tree, statistics can be used to calculate

8
the amount of year-to-year agreement within trees as well as between trees. Two cores per tree

also enable the researcher to start the crossdating process within a tree and to better represent

overall tree growth. If the researcher can take a cross section from a tree, then they have the

opportunity to examine as many radii as they want around the section. This sampling protocol of

20 trees with two cores per tree results in 40 cores represented in a site-level chronology; a site is

defined as a spatially proximal group of trees with similar environmental conditions such as

slope, aspect, and climate.

Tree-ring width responds to a similar set of environmental factors that limit tree growth. This is

the well-known biological principle of limiting factors. Because ring width is influenced by

anything that limits tree growth, the dendrochronologist must consider what problem is to be

addressed, then find the particular sites and trees that provide the necessary information. This

procedure is one of the most important principles called site selection. Dendrochronologists

label tree-ring chronologies sensitive when their ring-width patterns varied markedly from year-

to-year, while chronologies that had similar amounts of growth every year are called complacent

series. For climate studies, sensitive trees are targeted because their ring variation is likely to

better reflect climate than complacent trees. Climate is essentially the primary limiting factor

that imparts the year-to-year variability that makes crossdating possible. The history of

dendrochronology in Chapter 3 introduces the pioneers in the field that were the first to

recognize and apply many of these basic principles.

9
Subfields of Dendrochronology

Because tree-ring width can vary with anything that affects tree growth, annual records of many

natural phenomena can be developed. The term dendrochronology refers to the science of dating

tree rings and studying their structure to interpret information about environmental and historical

events and processes (sensu Kaennel and Schweingruber 1995). Many subfields within

dendrochronology have been developed and subsequently named by keeping the base of the

word “dendro” and adding a secondary prefix to describe the specific field being studied. For

example, dendroclimatology uses the variation in tree-ring structure and width to infer

information about past climate, while dendroarchaeology uses the date of the outside tree ring

from a beam to study the timing and process of archaeological construction. (see Table 1.1 for a

brief synopsis of the many subfields of dendrochronology, which are discussed in greater detail

in Chapters 7-11).

Limitations of Dendrochronology

As with all research, dendrochronology has certain limitations that must be acknowledged (Table

1.2). Annual tree rings (Figure 1.2) can form in any forest that has one yearly growth period

followed by a dormant period, but some locations, such as many tropical areas, do not have the

seasonality to allow the formation of annual rings. Crossdating must be used to verify the dates

of every ring in a sample. This technique is time consuming, takes special skills, and requires

much patience to learn and apply properly.

The development of wood through cambial activity that forms xylem and phloem in the trees is a

very complex process which has been the topic of an entire book (Larson 1994). Tree

physiologists have not been able to explain the exact biochemical processes that occur from

10
Table 1. 1 Subfield of dendrochronology. Many of the subfields also have subheadings (written
in bold) that more specifically describe the discipline. Dendrochronologists will often identify
themselves as practitioners of one or more of these subfields.
Subfield Description
Dendroarchaeology Tree-ring samples from beams and posts in archaeological
See Chapter 7 dwellings are dated to provide construction dates for the
dwellings. The position of these beams in the dwelling can be
used to study the timing of construction and expansion of
dwellings and to start to understand human behavior in these
cultures. Correlation with regional master chronologies can also
help to dendro-provenance archaeological and historical wood
object.
Dendroclimatology Samples from trees can provide short- or long-term records of
See Chapter 8 past climatic variability for the life-time of the trees. Most often
temperature, precipitation, and drought indices are reconstructed,
although anything that affects the processes of tree growth such
as number of cloudy days, relative humidity, or wind strength can
be reconstructed as well if they have limited growth. Information
from dendroclimatology has provided important information
about past climate change and help us understand what the future
climate might be like. This subfield also includes
Dendrohydrology which is the reconstruction of water level or
streamflow, although this is often sperated out as its own
subdiscipline.
Dendroecology Because trees are an important functional feature of many
See Chapter 9 ecosystems, they can be used as a natural record of ecological
processes, such as tree-line movement, successional processes
through the establishment and death of trees, fire occurrence
(Dendropyrochronology), insect outbreaks
(Dendroentomology), synchronous fruiting (masting) in trees
(Dendromastecology), or movement of invasive tree species.
Dendrogeomorphology The vertical structure of a tree enables it to gather the most light
See Chapter 10 while standing up straight so that land movement can be
reconstructed by the tilting of a tree and the resultant reaction
wood (thicker growth rings produced to straighten the stem of a
tree). Also tree death or establishment can be used to date
geologic phenomena such as landslides, mudflows, seismic
activity along faults (Dendroseismology), glacial activity
(Dendroglaciology), or volcanic events (Dendrovolcanology).
Dendrochemistry Trees absorb chemicals along with the water that they absorb
See Chapter 11 from soil and the gases that they take in from the atmosphere.
These chemicals are deposited in the wood in the trees’ stem,
roots, and branches and can be used as a record of contamination,
nutrient availability, and pollution. Stable isotopes can also be
measured in wood structure to reconstruct past temperature,
humidity, and the source of water or growing conditions of the
trees.

11
Table 1. 2 List of the general limitations to dendrochronology along with solutions to those
limitations.
Limitation Solution
Young trees Finding old trees in unique sites, buried wood, or
archaeological samples to extend a chronology.
Calibration datasets Conduct studies close to available climate or
ecological datasets or establish monitoring stations
for future calibration data sets in remote areas.
Regionalized climate data are also being used to
examine broad scale climate response in areas
without local climate stations.
Lack of ring formation in the Looking for a chemical or stable isotopic signal in
tropics tropical woods. Examine wood anatomy of many
species to find some with annual ring formation.
Lack of a physiological More wood anatomy, tree physiology, and
understanding of how tree rings biochemical studies need to be conducted to better
form understand the growth of tree rings.
 
 
 

12
 

Figure 1. 2 A cross section of an ash (Fraxinus sp.) tree. Note the dark colored heartwood near
the center of the tree and the suppressed rings in the 1930s and the 1950s (at arrow). The split in
the top left section of the wood is a natural break in the wood due to the cross section drying out
after it was sampled (photo by Jim Speer).

13
photosynthesis to the formation of tree rings. Some physiologists have argued that because we

do not completely understand the mechanisms that occur from the assimilation of abiotic

elements from the environment to the formation of the tree ring, we should not be conducting

dendrochronological studies. However, countless analyses of the correlation between tree

growth and environment variables have demonstrated consistent predictable results

demonstrating that despite our lack of understanding of the exact mechanisms, we know that tree

growth does reflect environmental variables. Over 100 years of productive research in

dendrochronology supports its validity and importance in environmental reconstruction.

The development of well-dated tree-ring chronologies for the reconstruction of environmental

variables is restricted by the location of sensitive tree-ring series (Fritts 1976). Trees are not

ubiquitous on the landscape and even when they are present, dendrochronologists must choose

specific sites that are likely to record the environmental variable that they wish to study.

Furthermore, many reconstructions depend upon calibration and verification datasets such as

modern temperature and precipitation records to create a statistical model to reconstruct past

climatic variations (Stahle et al. 1998a). These constraints geographically limit

dendrochronology to areas where the trees produce datable annual rings and where local paired

climate or ecological data exist for calibration of the trees’ response. This calibration data enable

a scientifically meaningful reconstruction and independent verification data allows the researcher

to determine the validity of the reconstructions. Still, tree-ring data can be useful even when

calibration data is not available. For example, the variability of rings in petrified wood

demonstrates variability in climate millions of years ago, and long-term growth suppressions and

releases in tree rings can demonstrate stand dynamics processes.

14
We can use trees to interpret past environmental phenomena, but trees are biological entities that

are driven by their own physiology and biochemistry that creates filters on the climatic or

ecological processes that they record in their annual rings. Trees, therefore, are not strict

monitors of the environment and these biological factors must be taken into consideration when

interpreting tree-ring data.

In comparison with many other proxy data, dendrochronology provides a relatively short record

with only three tree-ring chronologies in the world that extend back 10,000 years or longer.

These are developed from bristlecone pine (Pinus longaeva) in California (Ferguson et al. 1985),

oak (Quercus sp.) in Ireland (Pilcher et al. 1984), and oak in Germany (Becker 1993, Friedrich et

al. 2004). Dendrochronological records for any particular area are further constrained by the

need for well-preserved wood samples that represent a range of time scales. For example, most

wood in the eastern United States will decay on the forest floor in 20 to 50 years, while wood

found on a lava flow in the western United States may last for a thousand years without much

decay of the heartwood (Grissino-Mayer 1995). Very little wood can survive decay for longer

than hundreds or thousands of years thus limiting the length of our tree-ring records. Some

researchers have begun to use subfossil (buried but not permineralized) wood that may extend

their chronologies back 15,000 years or more (Roig et al. 2001, Guyette and Stambaugh 2003).

Petrified wood also provides a possible source of information on climatic variability from

millions of years in the past as long as the rings are well preserved (Chaloner and Creber 1973,

Falcon-Lang 1999) although it is important to not over-interpret the climate information that can

be gleaned from fossilized wood (Falcon-Lang 2005).

15
Objective

The objective of this book is to introduce the fundamental principles, concepts, and methods of

dendrochronology and provide the basic instruction, theoretical framework, and biological and

ecological background for the practitioner of tree-ring research. While portions of this

information are presented elsewhere (Stokes and Smiley 1968 [reprinted as Stokes and Smiley

1996], Fritts 1976 [reprinted as Fritts 2001], Phipps 1985, Schweingruber 1996), I hope to

compile this knowledge into a single basic user manual and easy reference book that covers the

breadth of the field. Whether you are a graduate student incorporating tree-ring chronologies

into your thesis or dissertation, a professional land manager who is looking for environmental

information, or a layperson who has heard about dendrochronology and wants to learn more, I

attempt to provide a practical resource that will provide you a strong start in the field of

dendrochronology.

No previous volume provides a comprehensive history of dendrochronology that incorporates

old world and new world pioneers in dendrochronology. I strive to provide a more complete

history of dendrochronology in chapter three that draws from European, American, Russian, and

Asian dendrochronologists up to the 1950s. An understanding of wood anatomy is becoming

more important in dendrochronology as we push the geographical bounds of past research and

start to study tree species growing in moist environments such as the Eastern Deciduous Forest

or the Tropics. I attempt to provide a quick primer on the aspects of wood growth and structure

that underlie the study of tree rings in chapter four. The core of this book is the field and

laboratory methods that are incorporated in chapters five and six. I try to provide a basic

16
founding in field practices and provide some greater depth in working with the programs and

statistics that are important to dendrochronology. I hope to provide a broad overview and a

useful starting point for all of the major subfields in dendrochronology in chapters 7-11. Each

chapter describes some specialized methods in each subfield and provides a bibliography as a

starting point for research into each area. Finally, chapter twelve describes what I see as some

frontiers in dendrochronology where researchers are gaining the most ground.

Dendrochronology is still a young science and there are many exciting frontiers yet to be

explored.

Summary

I have heard many discussions about the status of dendrochronology. People ask if it is a

discipline, a tool, or an application. I respond that the answer depends upon who is doing the

research and how they approach their work. I, among others, see dendrochronology as a thriving

discipline with its own governing body of principles, theoretical advancements, and areas of

important contributions to society. Others may simply use it as a tool to obtain dates or longer-

term records of past phenomena. Other researchers may work mainly on advancing theory in

different fields but call on the techniques of dendrochronology to advance their understanding

within their discipline. Through my experience teaching dendrochronology classes in the

university setting and coordinating and teaching the North American Dendroecological

Fieldweek (NADEF), I have found that that a basic set of knowledge exists that is new to and

important for the starting practitioner of dendrochronology.

17
The graphics that are collected in this book are the ones that visually represent the state of

knowledge and the theoretical basis of this firmly established field. The text is meant to present

the principles and methods that you will need to work through basic research projects in

dendrochronology on your own. This book is not intended as the final word in

dendrochronology and any practitioner of these methods should delve deeply into the primary

literature and other resource books available in the field. I strive to cite most of the pertinent

literature throughout the text and to lead the reader to useful internet resources that are available

in dendrochronology. I hope that you will find this book useful, whatever your intended

application.

18
Chapter 2: Some Basic Principles and Concepts in Dendrochronology

Introduction

Dendrochronologists follow some basic principles and concepts that describe sampling

protocols, model our concepts of how environmental factors are incorporated in tree growth, and

form our basic procedures of how to date tree rings and build chronologies. We also make some

basic assumptions about the natural world in the way that we conduct research. In this chapter, I

will discuss some of the assumptions, guiding principles, and core concepts in the field of

dendrochronology.

Some basic terminology associated with dendrochronology will be introduced first. The signal

to noise ratio is an important measure of the amount of desired information recorded in the

chronology versus the amount of unwanted information and random variation also included in

the tree-ring record. The noise can come from environmental factors not of interest to the

researcher. For example, growth releases due to mortality of neighboring trees (processes

involved in gap dynamics) are considered noise to a dendroclimatologist, whereas they are the

signal of interest for the dendroecologists interested in reconstructing forest succession.

Calibration is the process of comparing a known record of some environmental variable to the

tree-ring chronology for the purpose of determining tree growth response to that variable. We

use meteorological data (for example monthly temperature, precipitation, or Palmer Drought

Severity Index) as a calibration data set for climate reconstruction. Similarly, we use a record of

19
past fruiting of trees as a calibration data set for mast reconstruction (synchronous fruiting in

trees) or the historical records of insect populations to identify the growth pattern associated with

insect outbreaks. Part of this independent data can be withheld from the original model and used

to verify the reconstruction. This step is important to determine how accurate reconstructions

may be.

Principle of Uniformitarianism

The Principle of Uniformitarianism is a basic assumption of geology and most other natural

sciences. It can be succinctly stated as:

The present is the key to the past.

This means that the processes occurring today are the same processes that occurred in the past.

The classic example states that by collecting the sediment washing down a stream over a certain

time period one could extrapolate how long it will take the entire mountain to erode away,

because the Principle of Uniformitarianism assumes that the processes that determine the rate of

erosion remain the same through time. This basic assumption enables estimates of the rate of

change in natural systems. Dendrochronologists use the Principle of Uniformitarianism when we

reconstruct past climate. Researchers realize that the climate is changing through time and that

this change (e.g. the availability of CO2) may alter how a tree responds to climate, but this is the

best estimate that can be provided until further information is added to the model.

20
Dendrochronologists use calibration data sets such as meteorological data, historical records of

insect outbreaks, or masting to build mathematical models of how trees respond to these

environmental factors. Once this model is developed (usually with regression analysis) it

provides an understanding of how the trees respond to the variable of interest. The model can

then be inverted to reconstruct that variable into the past for the lifetime of the trees. For this

reconstruction to be possible, dendrochronologists have to assume that the processes affecting

the tree’s response to these environmental factors have not changed from the calibration time

period to the period of reconstruction. This is a common assumption made in the natural

sciences, but it has some drawbacks of which the researcher should be aware.

The trees’ response to the environment does vary with age. Seedlings are more sensitive to

environmental factors and are more likely to perish because of limitations in moisture

availability or temperature extremes. Young trees often go through a period of juvenile growth

during which they produce larger than average growth rings. Dendrochronologists should be

aware of and control for these tree responses as they build chronologies and reconstruct

environmental factors over the lifetime of the trees.

Humans have changed the environment, which may change how a tree will respond to climate

variations. We live in a world of elevated CO2 in the atmosphere; prior to the industrial

revolution the normal level of CO2 was 280 ppm while the current level is close to 380 ppm.

This amount of CO2 in the atmosphere is outside the natural range of variability recorded over

the past 100,000 years by ice cores. Most of the calibration climate data has been recorded

during a time of elevated CO2 and tree response to variability in temperature and precipitation

21
may be moderated by the amount of CO2 in the atmosphere. This may affect our climate

calibrations and the resulting climate reconstructions.

There are some ways to reduce the risk of violating the assumptions of the Uniformitarianism

Principle. For example, tree-ring series can be truncated to remove juvenile growth. This

shortens the resultant chronology and reduces sample depth further back in time, but it results in

more reliable reconstructions. Series can also be detrended by fitting curves such as a negative

exponential or a cubic smoothing spline (described later) to the ring-width measurements to

remove trends through time. But none of these treatments deals with the calibration problem of

living in a time of an altered climate. We know that our assumption that present processes have

not changed through time is not always correct, but Uniformitarianism is a productive starting

point in the analysis of past climates and environmental variability. The researcher must be

aware of these assumptions and work to overcome such limitations to understand the natural

world.

Principle of Crossdating

The principle of crossdating is the basic tenet of dendrochronology. It is the main tool by which

the exact year of growth of every annual ring is determined. Without crossdating, a simple ring

count is likely to produce error due to locally absent or false rings. Crossdating is imperative

when ring-width measurements are compared to annual phenomena such as meteorological data.

Without exact annual dating of the tree rings, accurate calibration is impossible because the

chronology will be misdated by one or more years. For example, the temperature data from 1973

22
may be erroneously compared to the annual ring grown in 1972 or 1974, and the result is a

degraded or non-existent climate signal.

The history of the concept of crossdating is relatively long. The French naturalists Duhamel and

Bufon first used crossdating to identify the 1709 frost ring in a series of samples collected in

1737. Twining rediscovered the process in 1827 and Babbage spoke about it in great length in

1838. But it was not until 1904, when Douglass laid out the basic methodology of skeleton

plotting and the refined technique of the memorization method that crossdating was really tested

and documented.

Crossdating matches the pattern of wide and narrow rings in a tree to determine the location of

real ring boundaries based on anatomical wood structure, providing a check of the actual date of

a specimen (Figure 2.1). In this sense, a tree core is like a bar code with varying widths of lines

representing each year. The patterns from one tree can be matched with those of other trees to

determine if all of the rings are represented on a sample. This technique shows where rings are

missing from a sample or where a tree might have formed two or more rings in one year.

Crossdating results in accurate dates for every single ring in the tree-ring record.

There are many ways to date tree rings, but the most repeatable and tested techniques are those

originally developed by Douglass. The method of skeleton plotting assigns each year of growth

to a vertical line on a piece of graph paper (usually graph paper with five lines per centimeter is

used). The length of the line represents the importance of the ring to the signal of the

chronology. Narrow rings are more important for recording limiting environmental factors, so

23
 

Figure 2. 1 Crossdating. Crossdating is the basic principle of dendrochronology and provides the
annual resolution of the dated rings (modified from Stallings 1949).

24
more attention is usually given to the rings that are below average in width. Therefore, the more

narrow a ring, the longer the line marked on the skeleton plot (Figure 2.2). Because of the age-

related growth trend, which will be discussed below, and the individual variability of growth in

the tree through time, the dendrochronologist uses a process of mental standardization in which

the relative width of the ring is determined by comparing the ring of interest to three rings on

either side. Only seven rings are compared at a time and the narrowest rings are noted on the

skeleton plot. This mental standardization keeps long-term trends or short-term suppressions

from dominating the signal in the chronology. The skeleton plot allows for a range of line

lengths from zero indicating an average or larger than average ring width to 10 which is usually

reserved for a ring found to be absent through crossdating. The sample about to be dated should

be visually scanned to determine the size of the smallest and largest rings across the entire cross

section to set the overall scale of the skeleton plot. The smallest rings in the sample will have a

line length of nine boxes and the entire range from 1-9 should be used for all samples. The

resultant plots illustrate the interannual ring- width variability within the wood sample whether

the wood is complacent with very similar ring widths or sensitive containing much variability in

ring width. Many beginning dendrochronologists have a difficult time with this apparently

arbitrary determination of the length of the line on the skeleton plot, but after some practice,

most researchers and students will produce very similar skeleton plots. The process can be

duplicated by a computer program, showing that it is not a purely subjective process (Cropper

1979). However, dating should always be performed visually and can be second checked with

various computer methods. Skeleton plotting allows two different trees growing at vastly

different rates to be compared to determine if all of the rings are represented (see Chapter 5 for

specific details in marking the wood and making a skeleton plot). A master chronology (a

25
 

Figure 2. 2 A) Photograph of a single skeleton plot showing the beginning and end arrows that
represent the inner most date and bark dates respectively. Notice on the left side, the sample ID,
Dendrochronologists name, species, and date are recorded. B) Once multiple skeleton plots are
completed on a site, they can be taped together with scotch tape and a master chronology can be
drawn from them. For a ring to be represented on the master chronology it has to appear on 50
percent of the plots then the length of the lines are averaged together (usually only counting the
trees that represent that ring). The master chronology is drawn upside down from a regular plot
so that it is easy to check the date of subsequent plots against it. You can see in the stack of
skeleton pots, the rings that are represented on most of the samples. These are the rings that
become marker rings on the master chronology.

26
record of ring widths representing the stand level signal) can also be developed from the

individual tree skeleton plots. This is usually done by lining up the skeleton plots so that they

share the same time axis along the bottom and any ring that is consistenly marked on half of the

samples in a given year will be averaged onto the master skeleton plot. Dead wood can then be

dated against this master chronology. One clear advantage of skeleton plotting over the list

method is that samples with unkown outside dates can be dated using skeleton plots.

The list method is a technique used to develop the chronology of marker rings without the added

steps of plotting them on graph paper (Yamaguchi 1991). It should be noted that the skeleton

plot provides more data and a clear graphic representation of the samples, making the dating of

difficult samples more probable. Also, the list method can only date complete samples from

living trees as the outer ring provides the starting point for the list. The list method, therefore, is

not of any use in dating archaeological samples or fire scars from dead wood. The list method is

a faster procedure and can be more efficiently used in wood with a clear pattern of rings. When

developing the list, the researcher starts at an anchor in time, which is the outside of the sample

with the known coring date. Care should be taken to develop the master list only from good

quality cores in which the samples are complete and the tree was living. Note the date of each

small ring on a piece of paper while counting back from the bark of the tree so that a list of

marker rings is generated (Figure 2.3). Those rings that are consistently noted between samples

will be the reliable marker rings that can be used to date other cores.

27
Figure 2. 3 A picture of marker rings recorded using the list method. Five rings (2001, 1995,
1989, 1981, and 1964) all appear as important marker rings that occur between all of the samples
that are recording growth at the time. Note that the marker rings from sample CCP22 stop at
1970 because this tree does not extend earlier than this time. One can list the inside ring date in
a box at the beginning of the list to indicate when the sample started recording.

28
The memorization method is generally used once the master chronology is known for a set of

samples (Douglass 1941). The master chronology may have been produced from skeleton plots,

the list method, or a published chronology. The marker rings in the chronology are memorized

(sometimes with a written aid) and the tree rings are counted back from the bark to the inside of

the core. Each time a narrow ring is encountered it is mentally checked with the list of

previously derived marker rings. If the small ring is a marker ring and should be small, then

continue dating the sample. If the ring is not a marker ring, it is usually best to count to other

marker rings to check the dates across a couple marker rings. The chronology of marker rings

would be consistently off from the master in the case of a missing or false ring. If the whole

chronology appears to be shifted forward in time by one year, then the wood representing the

period in time where that pattern started to diverge from the master should be examined to

identify the missing ring.

Principle of Limiting Factors

The Princple of Limiting Factors states that the most limiting factor will control the growth of

the organism. This is based on Liebig’s Law of the Minimum which is a simplification of the

actual physiological response of a tree to environmental forcing, but it can be used as a first

approximation of the environmental factor that is most likely to be recorded in a given tree-ring

chronology (Figure 2.4). For example, trees growing in the semi-arid environment of southern

Arizona are normally limited by the amount of rainfall each year and actually stop growth in the

middle of the summer when the soil moisture is depleted and very little rain falls to sustain the

trees. But those same trees will start to grow again when the monsoon rains come in late

summer and replenish the water. Trees growing at high elevation tend to be limited by

29
Figure 2. 4 Limiting Factors. Liebig’s Law of the Minimum states that whatever factor is most
limiting to growth will control the rate of growth for that organism. In this case, the slat labeled
PDSI (Palmer Drought Severity Index) would be the most limiting factor for plant growth,
therefore availability of moisture to the plant will control the ring width. It should be noted that
this limiting factor may change through time.

30
temperature, whereas defoliated trees are limited by the reduction in their photosynthetic

potential. Tree growth can also be limited because of a lack of access to nutrients in the soil.

Gardeners often experience the benefits of fertilizing plants with nitrogen to increase growth. A

limiting factor will dominate the growth for each year and will be the main variable recorded in

ring width creating a series of rings that vary in width from one year to the next (Figure 2.5). It

is possible, however, that this limiting factor will change through a plant’s life making

reconstructions of environmental factors more tenuous. When one variable was limiting but then

occurs in abundance, another limiting factor is likely to control growth. Also, trees may be

limited by multiple factors at one time, complicating the physiological response of the tree.

Principle of the Aggregate Tree Growth Model

The principle of Aggregate Tree Growth suggests that trees record everything that effects their

growth and provides a conceptual model for how to envision these effects, ultimately providing a

tool to tease apart the disparate effects of the environment on tree growth. Trees can be severely

limited by one factor, but most likely, they are recording multiple factors that limit their growth.

The Aggregate Tree Growth Model (Equation I; Cook 1985, Cook 1992) is used to conceptualize

this response and to try to understand the different variables that can affect tree growth.

31
Phloem

Xylem

Figure 2. 5 Pine crosssection. Conifer trees in temperate areas produce one ring per year. These
rings can be broken into the earlywood portion (open cells with thin cell walls) and the latewood
portion (cells with thick cell walls and a smaller lumen). Other features that are present are the
pith, resin ducts, cambium, and the bark. The variation in ring width is generally driven by
climate and results in the pattern of wide and narrow rings that we use to cross date the wood
samples. Sometimes a false ring may be present, as in this sample, where the tree growth slows
because of a reduction in the limiting factor for growth of the tree, such as drought. When that
environmental factor limiting growth returns (e.g. when it rains), the tree resumes growth and the
cells grade back to earlywood structure with thinner cell walls (from Fritts 1976).

32
Rt = f(Gt, Ct, D1t, D2t, Et) [I]

Where:

Rt is ring width at year t.

Gt is the age (or size)-related growth trend.

Ct is climate at year t.

D1t is the endogenous disturbance within the stand.

D2t is the exogenous disturbance from outside the stand.

Et is the error term incorporating all of the signal that is not

controlled for by the above variables.

This conceptual model demonstrates that ring width for each year is dependent upon a complex

array of variables that contribute to growth. Trees have an intrinsic age-related growth trend,

respond to current climate conditions as well as reflect the previous year’s climate, and are

affected by disturbances from within and outside of the stand. The age-related (also known as

size-related) growth trend results from a tree putting the same volume of wood on an ever

increasing cylinder. When a radius of the tree is examined, ring-width often decreases in size

with age of the tree. In an open-grown pine tree, this trend can be modeled with a negative

exponential curve, while the ring-width pattern from trees grown in a dense forest may be

dominated by competitive effects from neighboring trees more than this age-related growth

trend. Finally, some variability always remains that cannot be explained which is incorporated in

the model by the error term. Recent research has begun to explore the information that is

contained in the error term by looking at new variables such as biological constraints to growth

33
(see Speer 2001 for an example with mast reconstruction). Calling this the error term is not the

most accurate wording as it incorporates all things not explicitly identified in the model not just

errors in measurement.

The Aggregate Tree Growth model demonstrates the complexity incorporated in each year’s

growth, but it can also be used as a tool to explore the different layers of response. The age-

related growth trend can be removed from the chronology through basic standardization

techniques which will be discussed later. The climate data can be removed by running a

regression analysis between ring widths and the climate variables to which the tree responds.

The residuals from that analysis (the variability not accounted for by the regression model) can

be analyzed to determine what environmental factors are present beneath the age-related growth

trend and climate variables that have been removed, assuming the appropriate calibration data set

is available to build such a model.

Concept of Autocorrelation

All biological organisms are subject to autocorrelation because of the continuity and

unidirectional flow of the progression of time and the development of growth (Figure 2.6). The

needles of conifers produced in one year because of a favorable climate are maintained on a tree

the following years, adding to the photosynthetic potential of that tree (Figure 2.7; LaMarche

1974, LaMarche and Stockton 1974). Therefore, previous year’s climate affects current year’s

growth. This is the most obvious example of autocorrelation, but any biological organism

produces cells, proteins, and sugars which can be used in subsequent years, creating

autocorrelation in the response of that organism to environmental variables. Autocorrelation can

34
Figure 2.6 Autocorrelation. Tree growth often times includes autocorrelation which is the
statistical characteristic that the current year’s growth is affected by the previous year’s growth.
Autocorrelation can be driven by the biological activities of the tree in that the current year’s
climate will affect the heat, rainfall, and CO2 levels for this year’s growth but it also effects the
following years growth through development of new buds, sugars, and hormones. Finally the
climate from that same year will affect growth even further in the future by the development of
leaves, roots, and fruits.

35
Figure 2. 7 Needle retention in bristlecone pine at upper treeline related to summer temperature
(from LaMarche and Stockton 1974). Bristlecone pine trees retain their needles for many years,
in this case 16 years, resulting in good growing conditions in the past affecting current year’s
photosynthetic potential.

36
become a problem because most statistical analysis (such as regression analysis) assumes that the

data is not autocorrelated because it can artificially increase correlation statistics. Fortunately,

this component can be described and removed by determining the variance of the current year’s

growth that is explained by the previous year’s growth. This one-year lag is described by a first

order autoregressive model. The correlation to growth two years prior is described by a second

order autoregressive model, and this continues back in time until no significant autoregressive

signal can be detected.

Concept of the Ecological Amplitude

Ecological amplitude is the pattern of vegetation on the landscape which is controlled by the

range of climate variables to which a species responds (Lomolino et al. 2006). The microclimate

of a site is also modified by topography, slope, and aspect which affects the local distribution of

species. Based on this idea, a tree species should be less stressed at the center of its range and

more stressed near the margins of the range where the climate might be harsher for that species.

Therefore, for climate related research it is often better to sample a species near the edge of its

range to find trees that are more likely to record the climate variable of interest. These ecotones,

or edge regions, are also the areas where change is more likely to occur in the face of a changing

climate, making ecotones important sample sites. For example, black gum (Nyssa sylvatica) is at

the northern limit of its range distribution in New Hampshire (Figure 2.8) so samples taken from

this species at this location are likely to be more sensitive to temperature if that is the limiting

climatic factor at the northern edge of its ecological amplitude. Samples taken from just north of

the Everglades in Florida, at the southern end of black gum’s ecological amplitude, are not likely

37
Figure 2. 8 Distribution of black gum. The shaded area represents the range of black gum (from
Little 1971). (http://esp.cr.usgs.gov/data/atlas/little/nysssylv.pdf).

38
to be limited by lower temperatures but may be limited by competition with other species or by

an excess of precipitation and soil moisture.

Principle of Site Selection

Given that trees will record all of the variables that affect their growth, dendrochronologists use

the concept of site selection to maximize the signal recorded in the trees they sample. Sites

should be located where the trees are most likely to be stressed by the variable that the researcher

is interested in reconstructing (Figure 2.9). For example, if a precipitation reconstruction is

needed then trees located in an arid environment should be sampled. Trees that are growing on

the edge of their ecological amplitude, such as the extreme lower elevation of species in a semi-

arid environment or the southern boundary of the species in North America are more likely to

record drought events. A temperature signal is likely to be recorded at high elevation or high

latitude where low temperatures during the growing season can limit growth. Similarly,

historical documentation of insect outbreaks or fire history is used to guide initial sampling of

trees to demonstrate the tree-ring response to these phenomena.

Trees that are growing at the center of their ecological amplitude with favorable climate year

round are likely to produce complacent growth in which each ring is a similar width. If a tree is

complacent, it does not record much environmental variability that could be used for crossdating.

The tree is likely to be a poor environmental recorder; however it may be limited by its

biological ability to grow, which could be genetic or could be driven by other physiological

factors that might be of interest to a dendrochronologist, such as masting. The opposite of

complacent growth is sensitive growth, where the tree demonstrates considerable variability in

39
Figure 2. 9 Site selection. Specific sites are chosen from which trees are likely to record the
variable that we are interested in reconstructing. A tree with the same size rings is considered to
have a complacent ring-width series (tree on the left) and a tree with a lot of variability in growth
is considered to have a sensitive tree-ring series (tree on the right) (from Stokes and Smiley
1968).

40
year to year growth and is thus recording some environmental variable (Figure 2.9). All of these

factors must be taken into consideration when choosing a site to sample for a given research

project.

Principle of Replication

Replication is the use of multiple samples to develop an accurate stand-level chronology or the

use of many samples back in time to provide good sample depth throughout the chronology.

This principle was recognized by many of the early researchers in the field such as Twining

(1833) and Babbage (1838). By taking multiple samples on a site and matching the ring width

pattern between these trees, valid crossdating for the stand can be demonstrated. Averaging the

growth between two cores from one tree and between 20 trees on a site can remove individual

tree variability and yield a stand-level signal. Averaging across many samples enables

dendrochronologists to change the spatial scale over which a study is conducted; essentially

stepping up from the individual tree to the stand and even regional level (Figure 2.10).

Replication provides the basis for crossdating and contributes to the robustness of environmental

reconstructions. The appropriate sample depth can be determined by using the expressed

population signal (EPS) statistic described in chapter 6.

Concept of Standardization

Standardization removes age-related growth trends and other long-term variability that can be

considered noise by fitting curves to trends in ring series (Figure 2.11). Note that the long-term

trend could be the signal that climatologists are interested in when they examine long-term

41
 

Figure 2. 10 Sampling scale. By taking replicate samples, we can average out individual tree
variability and change our analysis level to higher spatial scales (such as the stand, watershed, or
regional levels) (modified by Bharath Ganesh-Babu from Swetnam and Baisan 1996).

42
 

Figure 2. 11 Standardization. The juvenile growth effect results in a higher level of growth when
a tree is young. If cores from three trees (A through C) are averaged together without regard for
this age-related growth trend, the resultant chronology (D) will mainly record when these
younger trees are incorporated into the chronology. If, however, the series are standardized with
a negative exponential curve and index chronologies are generated by dividing the measured ring
width by the model curve fit value, then these index chronologies (E through G) can be averaged
together to generate a master chronology that maintains its interannual variability and enhances
the stand level signal (H; modified from Fritts 1976).

43
changes in climate, but it is noise to an ecologist that is studying shorter-term variability in forest

dynamics. This process also removes differences in growth rates between samples and produces

a series mean equal to 1.0. The most conservative technique of standardization is the negative

exponential curve that is common in the growth of many trees and is geometrically mandated by

adding the same volume of wood on the surface of an ever increasing cylinder. The negative

exponential curve is deterministic meaning that it follows a model of tree growth. Other

standardization techniques are empirical, meaning they are chosen through experimentation to

find the best fit to a series of data. A cubic smoothing spline is an example of an empirical

model that uses a flexible curve that is allowed to adjust at a regular interval (Cook 1985; Figure

2.12). When using these forms of standardization, the researcher should be cognizant of the

signal that is being removed from the record. A 40-year cubic smoothing spline, for example,

removes 50 percent of the variance at 40 years, leaves 99 percent of the variance at 12.67 years,

and removes 99 percent of the variance at 126.17 years so that very little century length signal is

left in the resultant chronology (see table 2.1 for data on other splines). Splines are a more

organic fit to the data than straight line or exponential fits, but they do remove different amounts

of variance at different temporal scales.

Standardization is a powerful technique that can be used to minimize noise in a chronology and

increase the signal of interest, but it is also a complex issue and probably one of the more

controversial steps in dendrochronological analysis. New techniques are being developed for the

standardization of tree-ring series and will be further discussed in chapter 6.

44
Figure 2. 12 Standardization with a negative exponential curve and cubic smoothing splines.
Each curve represents the same set of ring width measurements back to A.D. 1500. A negative
exponential curve is fit to the ring widths followed by a series of shorter cubic smoothing splines
(300, 150, 80, 40, and 20 years) and the resultant curve fit is shown as the smooth black line.
You can see that shorter and shorter wavelengths are removed from the master chronology.

45
Table 2. 1 The frequency of the variance that remains using different cubic smoothing splines.
Spline Length Leaves 99 percent Leaves 50 percent Leaves 1 percent of
of Variance at of Variances at the Variance at
20 6.24 years 20 years 63.09 years
40 12.67 years 40 years 125.17 years
60 19.01 years 60 years 189.26 years
80 25.35 years 80 years 252.35 years
100 31.69 years 100 years 315.43 years
150 47.53 years 150 years 473.15 years
200 63.37 years 200 years 630.87 years
300 95.06 years 300 years 946.20 years
400 126.75 years 400 years 1261.73 years
500 158.44 years 500 years 1577.17 years

46
Summary

The principles and concepts described in this chapter help dendrochronologists to work through

the process of developing valid dendrochronological reconstructions. Some studies are starting

to show that our perception of concepts such as ecological amplitude and what we expect a site

should be recording are not necessarily as we suppose (Tardif et al. 2006, Speer unpublished

data). The patterns that we observe on the landscape are more complex than some of the basic

models presented in this chapter, but these principles help to guide our sampling and chronology

development and are a good first approximation of the factors that drive our sampling methods.

The next chapter explores the history of dendrochronology and discusses the people who

developed many of these principles.

47
 
Chapter 3: History of Dendrochronology

Introduction and the Early Years

Dendrochronology is a young discipline in the realm of the sciences with many new frontiers left

to be investigated. The first Laboratory of Tree Ring Research was founded at the University of

Arizona in 1937 by A.E. Douglass. The subfield of dendroecology became a major area of

research only in the 1970s and today new research is being conducted using stable isotopes from

tree rings to examine trees’ physiological responses to climate change. Despite these recent

beginnings, the idea that trees produce annual rings had been suggested since the time of

Theophrastus in 322 B.C. (Studhalter 1956).

In this chapter, I will describe the development of the field and many of the basic concepts that

we still use today (Table 3.1). In the 1400s and 1500s, some famous naturalists recognized the

annual character of tree rings and started to look to the environment for causes of variation in

ring growth. In the late 1400s, Leonardo da Vinci (Figure 3.1) described annual ring formation

and suggested that the growth of tree rings is related to weather (Stallings 1937, Sarton 1954,

Corona 1986). “…the rings in the branches of trees that have been cut off show the number of

its years, and which were damper or drier according to the greater or lesser thickness of these

rings.” (Kemp and Walker 2001; translation of Leonardo da Vinci’s Treatise on Painting).

From 1580 to 1581, Michel de Montaigne traveled through Germany, Switzerland, and Italy

and kept a diary of his journey. While in Italy, he reported on a conversation that he had with an

unnamed carpenter who noted the annual nature of tree rings.

48
Table 3. 1 The early dendrochronologists sorted by the dates that they used tree rings.

Scientist Location Year Contribution Reference


Theophrastus Greece 322 B.C. Noted that trees put on new growth every year Schweingruber
1996
Leonardo da Vinci Italy 1452-1519 Noted the relationship between climate and tree growth Sarton 1954
Wrote the Trattato della Pittura mentioning tree rings
1482-1498
Michel de Montaigne Italy 1581 The idea of counting tree rings to attain tree age was related to Montaigne by an unnamed carpenter Sarton 1954

Duhamel and George France 1737 Counted rings to determine the date of a conspicuous frost ring Webb 1986
Louis Leclerc de Dean 1978
Buffon
Carl Linnaeus Sweden 1707-1778 Counted rings to determine the age of trees Webb 1986
Online Article
A.C. Twinning Connecticut 1827 Used crossdating and noted the common signal across a site. Dean 1978
Theodor Hartig Germany 1837 Set up the ecological basis for dendrochronology in Germany. Schweingruber
1996
Charles Babbage England 1838 Noted the concepts of competition, complacency, replication, reaction wood in trees, the concept of parent Babbage 1838
trees, climate reconstruction, and specific tree-ring patterns from storms and floods.
Heizer 1954
Zeuner 1958
Jacob Kuechler Texas 1859 Used modern principles of site selection by choosing trees on low ridges with good drainage and he used Dean 1978
tables to note ring characteristics demonstrating an early example of crossdating
Robert Hartig Germany 1867 Used tree rings to date events of hail, frost, insect damage, and examined trees that were killed by pollution Schweingruber
1996
A.L. Child Nebraska 1871 Compared tree growth in red maple to meteorological data for spring and summer Webb 1986
A. Stoeckhardt Western 1871 Examined forest damage from air pollution. Eckstein and
Europe Pilcher 1990
Jacobus C. Kapteyn Denmark 1880-1881 Examined the relationship between tree growth in oaks and rainfall Webb 1986
F. Shvedov Ukraine 1892 Examined black locust trees for a dendroclimatic analysis that he used for prediction. Kairiukstis and
Shiyatov 1990
B.E. Fernow New England 1897 Wrote a paper on determining the age of blazing on trees by counting the tree rings Webb 1986
A.E. Douglass Arizona 1904 Developed crossdating as a tool and was persistent in developing chronologies and training future Webb 1983
dendrochronologists
Nash 1999
Bruno Huber Germany 1940 Spent considerable amount of time dating wood samples and worked on new statistical techniques to Schweingruber
quantify the strength of dating of different specimens
1996

49
 

Figure 3. 1 Leonardo da Vinci (1452-1519) mentioned that trees put on annual rings and respond
to local climatic conditions in the Trattato della Pittura (Treatise on Painting), which he worked
on from 1482 to 1498 in Milan at the court of Ludovico Sforza (from the Library of Congress).

50
The artist, a clever man, famous for his ability to make mathematical instruments, taught

me that every tree has inside as many circles and turns (cerchi e giri) as it has years. He

caused me to see it in many kinds of wood which he had in his shop, for he is a carpenter.

The part of the wood turned to the North is the straightest, and the circles there are closer

together than in the other parts. Therefore when a piece of timber is brought to him he is

able, he claims, to tell the age of the tree and its situation (the orientation of the section),

(translation of the original Italian text by Sarton 1954).

The 1700s and the 1709 Frost Ring

Many other early scientists had recognized tree-ring growth and used tests to demonstrate that

trees produce rings annually. French naturalists Henri Louis Duhamel du Monceau and

George Louis Leclerc de Buffon discovered in 1737 that a conspicuous frost-damaged ring

occurred 29 years in from the bark of several newly felled trees in France recording the 1709

marker ring. The winter of 1709 was significant in dendrochronology because of its frequent use

as a marker ring by early dendrochronologists across Europe. Carl Linneaus noted the 1709

frost ring when he examined wood samples in Sweden (Linnaeus 1745, 1751). Two other

scientists also noted this frost ring in other countries demonstrating how important this type of

marker ring is in dating tree rings. Duhamel went on to conduct early experiments in pinning

trees (the process of creating a small wound in the tree that can later be sampled to examine

growth since the wounding) and used aluminum foil as an early type of dendrometer band

around the tree to examine the annual growth of trees in 1751 and 1758 (Studhalter 1956).

51
In 1785, Friedrich August Ludwig von Burgsdorf examined tree growth over a blaze marking

a trail that was cut into a tree in 1767. He found 18 rings had formed since the blaze was cut,

thus demonstrating annual growth in this oak (Quercus sp.) tree in Germany. Burgsdorf also

noted the 1709 frost ring in beech (Fagus sp.) and other trees in Germany (Studhalter 1956).

Alphonse de Candolle, a Swiss botanist, discovered the 1709 frost ring when examining juniper

(Juniperus sp.) tree rings in France in 1839-1840. Candolle went further by suggesting a number

of early methods that could be used to crossdate rings, although much of his work dealt with

average ring widths over a period of time rather than maintaining the annual resolution of the

rings (Studhalter 1956).

The 1800s: Tree Rings Become Common Knowledge

De Witt Clinton in 1811 (while he was mayor of New York City and the year before an

unsuccessful run at the U.S. presidency) counted tree rings from trees growing on the earthen

mounds near Canandaigua in New York State. He estimated that the trees had about 1,000 rings,

and concluded that the mounds were created by prehistoric Native Americans and not by early

Europeans. This was the first recorded archaeological use of tree rings although Clinton did not

exactly date the wood to obtain more reliable bounding dates on the structures and his ring count

of 1,000 rings seems greatly exaggerated considering the tree species in this area he was able to

test if the structure was made prior to or after European colonization (Zeuner 1958).

Alexander Catlin Twining (Figure 3.2) used replicated samples in 1827 from many trees

harvested for the building of a wharf in New Haven, Connecticut. This early account is unique

because it conducts dating across multiple samples and relates the growth to climate.

52
 

Figure 3. 2 Alexander Catlin Twining (1801-1884) noted the application of using tree rings to
examine climate in 1827. Photo from http://www.rootsweb.com/~ctnhvbio/Twining_Alexander.html.

53
In the year 1827, a large lot of hemlock timber was cut from the north eastern slope of

East Rock, near New Haven, for the purpose of forming a foundation for the wharf which

bounds the basin of the Farmington Canal on the East. While inspecting and measuring

that timber, at the time of its delivery, I took particular notice of the successive layers,

each of which constitutes a year’s growth of the tree; and which, in that kind of wood, are

very distinct. These layers were of various breadth, indicating a growth five or six times

as full in some years as in others, preceding or following. Thus every tree had preserved

a record of the seasons, for the whole period of its growth, whether thirty years or two

hundred, -and what is worthy of observation, every tree told the same story. Thus, if you

began at the outer layer of the two trees, one young and the other old, and counted back

twenty years, if the young tree indicated, by a full layer, a growing season for that kind of

timber, the other tree indicated the same (Twining 1833: 391-392; The italics are in the

original text).

Twinning goes on to foreshadow the use of tree rings to reconstruct climate beyond modern

records. He also suggests that different genera will have a different climate response.

It would be interesting … to compare the sections of one kind of tree with that of another

kind from the same locality, - or to compare sections of the same kind of tree from

different parts of the county. Such a comparison would elicit a mass of facts, both with

respect to the progress of the seasons, and their relation to the growth of timber, and

might prove, hereafter, the means of carrying back our knowledge of the seasons, through

a period coeval with the age of the oldest forest trees, and in regions of the country where

54
scientific observation has never yet penetrated, nor a civilized population dwelt (Twining

1833: 393).

Charles Babbage (Figure 3.3) in England wrote at length about the application of tree rings in

determining the age of geological strata in his 1838 paper “On the Age of Strata, as Inferred from

Rings of Trees Embedded in Them” (Heizer 1954). He discusses, with examples, the concept of

crossdating (the process of matching ring widths to obtain exact dates of annual growth) and

goes further to mention that distinctly small rings are due to climatic variability. Babbage (1838)

also discusses competition, complacency, replication, reaction wood in trees, the concept of

parent trees, climate reconstruction, and specific tree-ring patterns from storms and floods. He

also formulates a research agenda using stem analysis to understand the growth throughout an

entire tree and suggests that tree-ring patterns found in the roots should be the same as those

found in the canopy.

In the following quote, Babbage talks about the climate response of trees and their ability to

record climate through time. “These preeminent effects are obvious to our senses; but every

shower that falls, every change of temperature that occurs, and every wind that blows, leaves on

the vegetable world the traces of its passage; slight, indeed, and imperceptible, perhaps, to us, but

not the less permanently recorded in the depths of those woody fabrics” (Babbage 1838: 258).

Babbage described the principle of crossdating with an example of the trees’ response to climate.

If we were to select a number of trees of about the same size, we should probably find

many of them to have been contemporaries. This fact would be rendered probable if we

55
 

Figure 3. 3 Charles Babbage (1791-1871) wrote an article in 1838 about the potential for
dendrochronology applied to buried wood in the Ninth Bridgewater Treatise, although his tree
ring comments were restricted to “Note M: On the Age of Strata, as Inferred from the Rings of
Trees Embedded in Them” (image from http://encyclopedia.laborlawtalk.com/Charles_Babbage .)

56
observed, as we doubtless should do, on examining the annual rings, that some of them

conspicuous for their size occurred at the same distances of years in several trees.… The

nature of the season, whether hot or cold, wet or dry, might be conjectured with some

degree of probability, from the class of tree under consideration (Babbage 1838: 258-

259).

The concept of the effect of local ecology on tree-ring response were also noted by Babbage.

“Some [trees] might have been protected by adjacent large trees, sufficiently near to shelter them

from the ruder gales, but not close enough to obstruct the light and air by which they were

nourished. Such a tree might have a series of large and rather uniform rings; during the period of

its protections by its neighbour; and these might be followed by the destruction of its protector”

(Babbage 1838: 260).

One of the more important principles in dendrochronology is the concept of replication which

Babbage also realized could be used to examine broad scale climate patterns. “But the effect of

all these local and peculiar circumstances would disappear, if a sufficient number of sections

could be procured from fossil trees, spread over considerable extent of country” (Babbage 1838:

260-261).

From 1837 to 1877, Theodor Hartig taught botany at the University of Braunschweig in

Germany where he laid the ecological groundwork for later dendrochronological research in

Germany done by Robert Hartig (his son) and Bruno Huber (Schweingruber 1988,

Schweingruber 1996). Theodor Hartig started a tradition of ecological examination in Germany

57
that included the use of tree rings. Much of the present day dendroecological research in Europe

has grown from this foundation.

Jacob Kuechler, a German immigrant to the U.S. with an interest in weather, used crossdating

to examine three post oak (Quercus stellata) trees from Texas in 1859. Kuechler’s (1859) own

publication is in German, but an editorial by Cleveland Abbe (1893) relates an investigation

made by Kuechler and reported on by Colonel William W. Haupt. Kuechler used modern

principles such as site selection by choosing trees on low ridges with good drainage as well as

tables to note ring characteristics, demonstrating an early example of crossdating (Stallings 1937,

Glock 1941).

Many of the early researchers mentioned above noted the climatic application of tree rings.

Other researchers in the late 1800s noted additional applications that could be studied using tree

rings. In 1866, the German botanist Julius Ratzeburg was probably the first to document an

insect outbreak due to the effect of defoliation by a caterpillar on tree rings (Ratzeburg 1866).

He was able to assign absolute dates to the outbreak events by examining tree rings (Studhalter

1956). In 1882, Franklin Hough also discussed the possibility of dating insect outbreaks from

damage the insects caused to trees (Studhalter 1956). Some of the first work in examining forest

damage from air pollution was conducted by Adolph Stoeckhardt in Western Europe

(Stoeckhardt 1871). This early investigation provided the lead for current researchers to examine

and quantify the effects of air pollution (Eckstein and Pilcher 1990). Elias Lewis (1873) would

frequently count the number of rings on fallen trees or stumps to determine the local growth rate

of a species and then use that number to estimate the age of living trees based on their diameter

58
and this age/diameter relationship. This was an early use of tree rings to estimate age structure in

a forest stand although today we realize that diameter is not always a good predictor of age

(Studhalter 1956).

John Muir, the famous American naturalist, noted the annual nature of tree rings and that they

could be used for geomorphic reconstructions, specifically determining the age of glacier-carved

structures in the Sierra Nevada Mountains of California. He also expressed interest in having the

time to study such phenomena with tree rings in his writing My First Summer in the Sierra which

were from his journals written during his first visit to the Sierra Nevada in 1869.

Have been sketching a silver fir that stands on a granite ridge a few hundred yards to the

eastward of camp – a fine tree with a particular snow-storm story to tell. It is about one

hundred feet high, growing on bare rock, thrusting its roots into a weathered joint less

than an inch wide, and bulging out to form a base to bear its weight. The storm came

from the north while it was young and broke it down nearly to the ground, as is shown by

the old, dead, weather-beaten top leaning out from the living trunk built up from a new

shoot below the break. The annual rings of the trunk that have overgrown the dead

sapling tell the year of the storm. Wonderful that a side branch forming a portion of one

of the level collars that encircle the trunk of this species (Abies magnifica) should bend

upward, grow erect, and take the place of the lost axis to form a new tree. (Muir 1911

compiled in Cronon 1997: 235-236).

Young pines, mostly the two-leaved and white-barked, are already springing up in these

cleared gaps [avalanche tracks in the Sierra Nevada of California]. It would be

interesting to ascertain the age of these saplings, for thus we should gain a fair

59
approximation to the year that the great avalanches occurred. Perhaps most or all of them

occurred the same winter. How glad I should be if free to pursue such studies! (Muir

1911 compiled in Cronon 1997: 280).

The recognition of the formation of annual rings in trees, their record of the climate and insect

outbreaks, and their dependence upon microsite differences was common knowledge to foresters

in the 1880s as evidenced by repeated discussion of annual rings in a forestry textbook called

The Elements of Forestry by Franklin Hough (1882).

In cross sections made years afterwards, the record of the seasons for a long period may

be determined, at least in effect, by the width of the rings of annual growth. We

sometimes find, at recurring intervals, a narrow ring, perhaps in every third year, that

may have been caused by the loss of leaves from worms that appear at that interval, and

that have thus left their record when every other proof of their presence has perished. We

have seen sections of trees in the museums of Schools of Forestry, in which these proofs

were recorded through a century or more of time, and the years could be definitely fixed

by counting inward from the year when the tree was felled (Hough 1882: 70).

Hough (1882) goes on to discuss the calculation of basal area increment for the purpose of

quantifying the amount of growth in each year. He also discusses the microsite variations that

affect tree-ring growth. “The rate of growth in wood differs very greatly, according to the soil,

elevation, aspect, climate, humidity, temperature, prevailing winds, and other causes” (Hough

1882; 75). The growth rings of multiple species are also shown in many figures throughout the

text. From this textbook, it must be concluded that knowledge of the annual growth rings of

trees and their response to the environment was common knowledge at this time. Hough (1882),

60
however, describes counting rings for this record (a practice still common in forestry today)

rather than the process of crossdating.

In 1881 Arthur Freiherr von Seckendorff-Gudent collected tree samples from 6,410 Austrian

black pine (Pinus nigra) throughout Austria, Hungary, and Slovenia and used many of the basic

principles in modern dendrochronology including crossdating and replication (Wimmer 2001).

He took the analysis further by noting the climate response of the trees to local climate data

(Seckendorff 1881).

When counting the tree rings on the disks, particular sequences of tree rings were

repeatedly found in most of the trees. As an example, on most disks we found the 1871

ring showing a wide latewood, and the narrow 1802 ring. The tree rings of 1862 and 1863

were very close and significant due to their obvious difference in the strength of the

latewood.

These significant tree-ring formations, which I named “characteristic tree-rings”, were an

excellent tool to determine the age even on trees grown on very poor sites. From the

many discovered “characteristic tree-rings” we always found at least a few on each disk.

…This method of age determination also helped to avoid counting false-rings.

A comparison of tree-ring characteristics with temperature and precipitation for the dated

years shows the relationship between tree growth and climate. Although, local site

conditions are the major factor for tree-ring formations, the effect of particularly warm

and cold years with low and high rainfall cannot be neglected. This influence (climate)

may be smaller or bigger in a growing region, whether the climate is more of local or

more of regional character.

61
Very hot and wet summers, such as the hot summer in 1811 (a good vine year) and the

hot and also wet year of 1846 are characterized with extreme tree-ring formations. While

the 1811 ring is distinct because of its weak (small) latewood, the year 1846 is significant

because of its wide latewood.

… For now it is sufficient to state that climate has an effect on the formation of tree-rings

and this effect can be softened by local site conditions but not revoked completely

(Seckendorff 1881 as translated in Wimmer 2001).

Robert Hartig (son of Theodor Hartig) used tree rings to date events of hail, frost, and insect

damage, publishing 34 papers from 1869 to 1901 on the anatomy and ecology of tree rings while

he was a professor at the University of Munich (Studhalter 1956, Schweingruber 1988). Robert

Hartig conducted a great deal of work in wood anatomy and was one of the first to look at the

physiological basis for ring formation. Hartig (1888) categorized the rings of conifers into three

sections of the spring zone, summer zone, and autumn zone. These divisions were later made

into earlywood and latewood that we still use today (Studhalter 1956).

He also examined trees that were killed by pollution and noted their long-term growth decline

before their death that made ring identification on the outside of the sections impossible. He was

able to find the rings represented on the stem near the canopy of the tree and used these samples

to crossdate the samples at the base and determine which rings were missing. Further work on

hail, frost, and insect outbreaks was conducted by K. Rubner (1910) and I. W. Bailey (1925a,

1925b).

62
F. Shvedov worked on an early precipitation analysis using two black locust (Robinia

pseudoacacia) trees and found a three to nine year cyclicity in the data so that he could correctly

predict upcoming droughts in 1882 and 1891 (Shvedov 1892). His early work in

dendroclimatology in Odessa in the Ukraine sets him apart as one of the early founders of

dendroclimatology (Kairiukstis and Shiyatov 1990).

Jacobus C. Kapteyn, a Dutch astronomer, used crossdating on over 50 oaks collected from

Holland and Germany to examine climatic patterns that might be recorded in those trees. His

work was completed in 1880, but was not published until 1914 (Kapteyn 1914). He concluded

that spring and summer rains were the most important climatic variables that affect tree growth

in this area of Europe. Kapteyn was ahead of his time, employing modern practices such as

crossdating, replication, and standardization. He used a 15-year running average to smooth his

data and noted that he was removing any cycles greater than 15 years that might have been

included in the wood. Kapteyn tested for missing rings, using crossdating and also identified

precipitation cycles of about 12.5 years in his final chronologies (Schulman 1937).

The Early 1900s, Douglass, and Huber

A.E. Douglass (Figure 3.4) was the first researcher to use crossdating “...persistently and

extensively…” (Studhalter 1956) from 1901 through the 1960s and has been named the

“…undisputed…father of dendrochronology” (Schweingruber 1988). He developed the

repeatable process of crossdating that is the cornerstone of dendrochronology today. Douglass

was an astronomer by training and assisted Percival Lowell in finding sites with clear skies for

observatories in the southwestern U.S. (Webb 1983). Douglass later had a disagreement with

63
 

Figure 3. 4 A.E. Douglass (1867-1962) in the storage room of the Laboratory of Tree Ring
Research underneath the football stadium at the University of Arizona in 1940 (from Webb
1983).

64
Lowell because Douglass would not publicly support Lowell’s hypothesis that patterns on the

surface of Mars were man-made canals.

On a horse–drawn carriage trip in 1901 near Flagstaff Arizona, Douglass noticed that a cross

section displaying rings from a ponderosa pine (Pinus ponderosa) tree showed a variation in

width of the rings. In 1904, he had the opportunity to examine a number of pine cross sections

and found a distinct pattern of small rings on the first, third, sixth, ninth, eleventh, and fourteenth

rings in from the bark. At that time, he documented what is now called the Flagstaff Signature,

consisting of small rings in 1899, 1902, and 1904. During later research he noticed that rings

from a tree in Prescott, Arizona (50 miles southwest of Flagstaff) had a similar pattern of small

rings. This pattern was repeated on numerous logs in the area and after years of work, Douglass

found that the pattern was repeated throughout the southwestern United States (Webb 1983).

Douglass moved to Tucson, Arizona in 1906 and took up a position as an Assistant Professor of

Physics and Geography at the University of Arizona. While at the University he continued his

work with tree rings, while teaching physics, and continuing his astronomical pursuits including

acquiring funding for Steward Observatory. Douglass was interested in reconstructing a long-

term record of sunspots. Knowing that sunspots were related to energy fluctuations in the sun

and that the sun provides energy for the climate system, Douglass hypothesized that one could

measure variations in solar intensity recorded in tree rings. He was later able to demonstrate that

the trees could be recording cycles driven by climatic parameters (Douglass 1909) and that they

were recording rainfall (Douglass 1914).

65
Douglass collected many species of trees from locations around the U.S. in California, Oregon,

South Dakota, New Mexico, as well as Arizona from his base at the University of Arizona. He

later collected more samples from England, Germany, Austria, Norway, and Sweden in the fall

of 1912 during his sabbatical. In 1915, Douglass collected his first giant sequoia

(Sequoiadendron giganteum) from a grove near Hume, California, in the same location that was

sampled by Ellsworth Huntington in 1911. The sequoia collections yielded a 3,000-year

chronology that extended back to 1305 B.C. By 1919, Douglass had collected 230 tree samples

from the U.S. and Europe and he had measured 75,000 rings (Webb 1983).

Douglass had the right combination of skills and talent for dendrochronology. He was

painstakingly meticulous, and he had a memory for dates. He memorized the entire chronology

for the southwest during his efforts to date the archaeological structures in this region. Douglass

also developed techniques that are still used today to facilitate dating, such as skeleton plotting

which will be introduced later.

Douglass formed the world’s first tree-ring laboratory in 1937, at the University of Arizona when

space underneath the football stadium bleachers was allocated as a temporary housing for the

lab. The Laboratory of Tree-Ring Research is still located there today. Douglass trained a

number of students (Figure 3.5), most notably Edmund Schulman, Ted Smiley, Florence

Hawley, James Giddings, and Emil Haury, who sustained the field of dendrochronology and it is

due to all of their efforts along with their European counterparts of Bruno Huber, Walter Liese,

Bernd Becker, Dieter Eckstein, and Fritz Schweingruber that dendrochronology is a highly

regarded field of research today. I will discuss these later contributions to the field of

66
 

Figure 3. 5 A.E. Douglass and his students in 1946. From left to right: Fred Scantling, Sid
Stallings, A.E. Douglass, Edmund Schulman, and James Louis Giddings (image from Nash
1999).

67
dendrochronology in chapters 7-11 which describe the methods and analyses applicable to each

subdiscipline.

Bruno Huber was one of the main researchers in Europe to spend considerable time and energy

in dating tree samples publishing over 39 papers from 1938 to 1970 on dendrochronology. From

1899 to 1969, he was a professor of forest botany in Germany at the Technical University of

Dresden and the University of Munich (Schweingruber 1996). Huber was aware of Douglass’

work and concluded the more complacent growth in trees from central Europe was due to the

more temperate and humid climate of the region. Despite these difficulties, Huber was able to

produce accurate chronologies and worked on new statistical techniques to quantify the strength

of dating of different specimens (Liese 1978). Samples taken from old structures that contained

wooden beams and posts were used by early European researchers, including Huber, to extend

their chronologies back in time. Huber used oak beams from medieval buildings in Franconia to

extend his chronology back to A.D. 1000 (Zeuner 1958). Huber (1935) also examined wood

anatomy and determined that fluid from the roots in ring porous trees (wood types with large

pores at the beginning of eaech ring such as oak trees) travels up the stem ten times faster than in

diffuse porous trees (wood types with disbursed pores such as maple trees) even though ring

porous trees only use their earlywood pores in the present year to transport fluid while diffuse

porous trees use several years of scattered pores to transport fluid. Around the same time, K.

Brehme, another German researcher, developed a chronology from larch (Larix sp.) trees in the

Bavarian Alps extending back to A.D. 1300 and Wellenhofer and Jazewitsch used oak (Quercus

sp.) trees from the Spessart Mountains in western Germany to build a chronology back to A.D.

1391 (Zeuner 1958).

68
Edmund Schulman was one of Douglass’ early students who made his own contributions to

dendrochronology. He conducted early work in dendroclimatology and expended much of his

energy in finding old trees to produce long chronologies for climate reconstructions. These

explorations lead him to find the bristlecone pines (Pinus longaeva and Pinus aristata) that are

now considered to be the oldest living organisms that are not cloneal (Schulman 1954).

Schulman (1956) also published the first chronologies from South America with the Chilean

incense cedar (Austrocedrus chilensis) and the Chile pine (Araucaria araucana).

Florence M. Hawley was another of Douglass’ students in the 1930s and she completed her PhD

with the University of Chicago in 1934. She extended Douglass’ work to the southeastern

United States, trying to date moundbuilder artifacts from Tennessee and Mississippi. She

developed some of the first chronologies from the southeastern United States under much

scrutiny because of the belief at the time that trees in the eastern deciduous forest would not

produce datable tree rings. She continued this research from a professorship in anthropology at

the University of New Mexico.

The Modern Era and International Organization

The previous record brings the history of dendrochronology up to the 1950s and 1960s with the

careers of Douglass, Schulman, and Huber. Since that time, many prominent

dendrochronologists have made great strides in the science including Dieter Eckstein, Fritz

Schweingruber, Bernd Becker, Mike Baillie, Gordon Jacoby, Hal Fritts, and Ed Cook who each

published more than 100 papers in dendrochronology as determined from the online

69
Bibliography of Dendrochronology through 2006. The number of publications has risen

dramatically so that Grissino-Mayer’s online Bibliography of Dendrochronology now holds

11,202 citations (see the internet references in Appendix E). I will discuss the contributions to

the field by many of the modern practitioners in the second half of this book.

International organizations that encourage dendrochronological research, teach

dendrochronology, and provide venues for tree-ring research presentations have developed in the

last century with a great influx of members and meetings over the past 20 years. The Tree-Ring

Society (TRS) is the oldest society of dendrochronologists and was founded by A.E. Douglass in

1935. The journal Tree-Ring Bulletin was first published in 1934 and has since changed its

name to Tree-Ring Research. It now has over 200 members from more than 30 countries. The

Association for Tree-Ring Research (ATRR) started in Europe in 2003, providing a network for

European dendrochronologists. A second tree-ring journal called Dendrochronologia was first

published in 1983 and is another major outlet for dendrochronological literature. A new

organization called the Asian Dendrochronology Association (ADA) started in 2006 and

provides a network for Asian dendrochronologists.

A.E. Douglass started annual meetings on dendrochronology in 1934, which continued to run in

1935, 1936, 1937, 1939, and 1941. These first meetings included researchers from Arizona and

New Mexico that had an interest in archaeology and climatology. The Tree-Ring Society

continued to meet at various venues through the intervening years including a notable meeting in

Tucson, Arizona in 1974 where the idea of regular international meetings arose and another

major meeting occurred in Norwich, England in 1980. Regular international meetings on

70
dendrochronology started at the International Conference in Ystad, Sweden in 1990 and

continued with other international conferences since that time. These International Conferences

on Dendrochronology have included the 1994 meeting in Tucson Arizona, which had 207

participants from over 35 countries, and meetings in Mendoza, Argentina (2000), Quebec City,

Canada (2002), and Beijing, China (2006). The next meeting is planned for Finland in 2010.

Smaller conferences have been developed around the world to encourage local research with

EuroDendro being the longest running of these conferences. Their first meeting was in

Lourmarin, France in 1989, followed by meetings in Liège, Belgium (1990), Travemünde,

Germany (1994), Moudon, Switzerland (1996), Savonlinna, Finland (1997), Kaunas, Lithuania

(1998), Malbork, Poland (1999), Gozd Martuljek, Slovenia (2001), Obergurgl, Austria (2003),

Rendsburg, Germany (2004), Viterbo, Italy (2005), and Hallstatt, Austria (2008) (Dieter Eckstein

personal communication). In recent times other regional conferences have been developed such

as the Southeast Asian Dendrochronology Conference (1998), the Asian Dendrochronology

Conference (2007), and the Ameridendro Conference (2008).

Fritz Schweingruber started an International Dendroecological Fieldweek in 1986 and Paul

Krusic started the North American Dendroecological Fieldweek in 1990 (Speer 2006). These

two educational outreach programs are in their 22nd and 18th years, respectively. These

fieldweeks continue to be one of the main educational opportunities for researchers that do not

have access to local dendrochronology courses. The success of the fieldweek model has led to

the development of the South American Dendrochronological Fieldweek (now in its 4th year), a

Southeast Asian fieldweek, and a summer course in Turkey.

71
A more lasting contribution of this international collaboration in dendrochronological research is

the development of the International Tree-Ring Databank (ITRDB; Grissino-Mayer and Fritts

1997). This data archive and computer forum arose from the international meeting in 1974,

during which participants expressed the need for a repository of tree-ring chronologies so that the

work of individual researchers can be passed along and preserved through time. Hal Fritts

founded the ITRDB and was its main proponent in its early years. In 1990, the National

Oceanographic and Atmospheric Administration (NOAA) took over the operation of the ITRDB

and founded the World Data Center – Paleoclimatology A (WDC) program in Boulder,

Colorado. The databank currently holds more than 2000 chronologies from six continents. In

1988 the managers of the ITRDB started a computer forum to enhance communication between

tree-ring researchers around the world. Today this forum has over 600 members from 32

countries that subscribe to the listserve. All of these international organizations, meetings,

fieldweeks, and the ITRDB continue to foster an international tree-ring community.

Summary

Dendrochronology as a field has grown out of the prior work by all of the researchers mentioned

in this chapter. From this work, you can see how we have accumulated knowledge over time

and continue to improve dendrochronology as a science. We keep adding new applications,

longer records, and larger spatial analyses. Chapter 7-11 in this book will describe each subfield

in greater depth and discuss the modern recent history of dendrochronological research. Because

the number of researchers and the amount of research has increased so tremendously since A.E.

Douglass’ time, it is hard to synthesize all of that work into one volume. I hope that the

72
references in the remainder of this book will be a good guide to the varied publications in

dendrochronology.

73
Chapter 4: Growth and Structure of Wood

Introduction

Tree rings are composed of individual cells that constitute the building blocks of the organism of

the tree. One must understand the cellular level of tree growth in order to accurately identify the

individual tree rings. A basic understanding of tree physiology is also important for

comprehending the biological processes that link the environment to ring formation. Tree rings

are the end result of a complex sequence of assimilation of natural resources by the tree. A

cascade of chemical reactions and cell division ultimately produce the annual ring that contains

the information dendrochronologists analyze. Kozlowski and Pallardy (1997) and Salisbury and

Ross (1992) are suggested for further reading.

Tree Physiology

Gymnosperms (plants that produce naked seeds) are more primitive phylogenetically than

angiosperms (flowering plants) and have less-developed and fewer cell types. Gymnosperms

(also known as softwood trees or conifers) transport water from the roots to the leaves through

tracheids (long narrow cells that comprise growth rings) in the outer living part of the xylem in

the area of the sapwood (the region with living parenchyma cells). Angiosperms (also known as

hardwood or deciduous trees) more efficiently transport most of the water and nutrients from the

soil to the leaves in specialized, capillary-like cells called vessels. These vessels are larger in

diameter than tracheids and transport water more efficiently but are more prone to embolism (air

bubble formation during conduction that blocks water movement). Angiosperms are further

devided into monocotyledons (or monocots) and dicotyledons (or dicots). A cotyledon is a seed

74
leaf or a leaf that breaks out of the seed. The monocots (like palms and yucca plants) produce

vascular bundles of xylem and phloem tissue but they do not produce a vascular cambium that

results in growth around the stem of the plant which would otherwise produce annual rings.

Therefore, monocots are not useful for dendrochronology, although researchers may be able to

quantify the age of some monocots through incremental height growth patterns as the tree grows

taller. Dicots, on the other hand, often produce annual rings around the circumference of the tree

from cell division in the vascular cambium.

Most gymnosperms and dicots in seasonal climates produce one ring per year. The ring can be

divided into earlywood and latewood. Earlywood is defined as cells that have large lumen (the

opening in the center of the cell) relative to the cell walls. Latewood cells are always flattened

and have a more compact lumen relative to the cell walls and consequently appear darker (Figure

2.5). Earlywood is usually produced in spring and early summer while latewood is formed in the

late summer. However, this timing varies with species and environmental forces.

Apical meristem (or primary meristem) are located at the tips of branches and roots and are the

origin for elongation of branches, roots, and height growth in a tree. Secondary meristem is

produced in most gymnosperms and dicots and enables a tree to grow in circumference through

time and produce tree rings. Cell division occurs in the vascular cambium (often simply called

the cambium) which is a narrow layer of meristematic cells between bark and wood. During cell

division xylem is produced toward the inside of the tree, becoming the wood structure that

supports the tree, and phloem is produced towards the outside of the tree and becomes the inner

bark. A cork cambium forms the outer bark of most trees. The walls of all woody cells

75
continue to thicken up to a cell-type specific extent within a few days or weeks before the cells

die, lose their protoplasm, and start to function for the tree as conducting or strengthening tissue.

Water transport is driven by transpiration (the evaporation of water) through the stomata in the

leaves of the tree and the cohesion of water molecules throughout a connected column from the

leaves all the way down to the roots. Transpiration is the pump that drives water (and

subsequently nutrient) uptake in the roots. The phloem cells remain alive much longer than the

xylem cells and transport the products of photosynthesis (sugars and hormones) down the tree.

Less frequently, a series of thin-walled ray cells are also produced during cell division in the

cambium resulting in a radial cell component that connects the outside of the tree to the inside.

Heartwood forms in the middle of the stem as a result of an active production of substances,

mainly phenols, which are deposited to close down the structure and guarantee that the wood is

resistance to decay (Figure 4.1). The lighter colored outer wood is called the sapwood and in

conifers is the area that tranports water up the tree from the roots to the canopy.

Visualize a tree as a series of stacked cones representing a complete sheath of wood that is put on

the tree each year. The tree grows upward by cell division at the apical meristem (or shoot tip)

and outward from cell division in the secondary meristem causing the cones to stack upward and

grow outward. When a dendrochronologist cores a tree, a sample is removed from bark to pith,

collecting the full number of rings produced at the height of the core. The pith is the bundle of

cells produced by the upward growth of the apical meristem, allowing trees to reach to greater

heights and creating the cambium initials that start secondary thickening of trees (Figure 4.2).

76
Figure 4. 1 A juniper from Jordan showing lobate growth demonstrating poor circuit uniformity
where the rings pinch out around the circumference of the section. Notice the darker inner wood
called heartwood and the lighter colored outer wood called sapwood. See figure 1.2 for and
example of good circuit uniformity.

77
 

A B

Figure 4. 2 Pith characteristics. Pith is the very center of the tree that is formed by the terminal
leader as the tree extends in height each year. If one hits the pith when coring the tree, that date
will provide the exact age of that tree at that height. A) Longitudinal section of butternut with a
chambered pith, B) beech, C) catalpa, and D) sumac, and E) oak (graphic from Hoadley 1990).

78
Basic Wood Structure

All wood samples can be examined from three distinct views, or planes, which provide a

different perspective on the cells that compose the wood (Figure 4.3). The cross sectional view

(also called the transverse view) is what we see on the surface of a stump when a tree is cut

down. In this view, you can clearly see the cross section of the tracheids in coniferous trees,

which are elongated tube-like cells that make up the majority of the wood and function to

transport fluids and nutrients vertically in the xylem of the tree. This is the view that

dendrochronologists examine most frequently. If you look at a side view along a cut from the

bark to the pith of the tree, you will be examining the radial view of the section (think of the

radius of a circle). In this case, you can see the full length of the tracheids, but the ring

boundaries are often obscured. The last view is a cut down the outside of the tree, basically

parallel to the pith column. This is the tangential view (tangent, or perpendicular, to the radius),

and can often be seen in furniture as the veneer cut from a tree. Each view provides a different

perspective that a wood anatomist can use to identify the type of wood being examined.

Cell Features and Types

Gymnosperms, such as pine (Pinus sp.), spruce (Picea sp.), and juniper (Juniperus sp.)(Figure

4.4), produce simpler wood structure (Figure 4.5) than hardwoods and are mainly composed of

elongated tracheids that are connected by boredered pits between the tracheids and resin ducts

may occasionally occur (Figure 4.6). Tracheids make up most of the cells in conifer wood and

function as structural and conducting elements, transporting nutrients along with water from the

roots. Bordered pits are evident on the tracheids’ cell walls that allow water transport from one

tracheid to another. Parenchyma are another cell type that can be found in gymnosperms. These

79
 

Figure 4. 3 Planes of wood structure. Wood samples can be sectioned to expose three primary
surfaces used in wood identification. The cross sectional (or transverse) view, the radial view,
and the tangential view (from Fritts 1976).

80
.

81
 

Figure 4. 4 Gymnosperm wood examples. Examples of species with coniferous growth such as
pine (Pinus), Douglas-fir (Pseudotsuga menziesii), coast redwood (Sequoia sempervirens), bald
cypress (Taxodium distichum), hemlock (Tsuga), yew (Taxus), and cedar (Thuja) (graphic from
Panshin et al. 1964).

82
 

Figure 4. 5 Gymnosperm cells types. Gymnosperms only have a few cell types which are called
tracheids, parenchyma cells, ray cells, epithelial cells, and resin ducts (graphic from Hoadley
1990).

83
phloem

pit

Gymnosperm Structure

Figure 4. 6 Gymnosperm wood structure. In coniferous wood, the diameter of the cell, cell wall
thickness, and size of the lumen determine the ring structure (from Schweingruber 1996).

84
cells are alive and have a complete protoplast in the lumen of the cell. In a cross sectional view,

they can be identified as a normal cross section of a tracheid cell, except that the vacuole (central

cavity of the cell) is dark with cell material. Finally, resin ducts that transport resin throughout

the tree to seal off wounds can be found on the cross sectional view of many coniferous genera

(Figure 4.7). Because of the function of this resin, conifers are relatively resistant to decay.

Thickening of the cell walls and flattening of the cells are the main indicator of annual ring

structure in coniferous wood (Figure 4.7). Bordered pits in the tracheid walls and resin ducts can

be present, but pits do not occur in all conifers.

Angiosperms have more complex wood structure than gymnosperms (Figure 4.8). Fibers

provide the key structure and support for the tree, but angiosperms also produce vessels that are

used for the main water transport in the tree. Angiosperms may have large and small vessels

along with fibers, tracheids, and parenchyma cells (Figure 4.8), all of which are evident in cross

section. Pits are very evident in hardwood species in the radial section and allow for water

transport between individual vessels. Parenchyma cells are more common in hardwoods, and

actually form the ring boundary in some genera of diffuse porous species. A three-dimensional

wood block of a hardwood sample shows that vessels dominate the view, but fiber cell size is

still important for differentiating ring boundaries in some genera (Figure 4.9). Rays form

perpendicular to the ring boundaries and are very prominent in hardwoods, providing efficient

transport and storage of nutrients, photosynthetic products, and some metabolic wastes to the

heartwood (Figure 4.10).

85
Figure 4. 7 Resin ducts in a gymnosperm. Resin ducts are produced in most conifers. The resin
duct is a large hollow vessel that is surrounded by guard cells. When the guard cells relax, they
allow resin to flow through the resin duct. The tree uses this resin to seal off wounds. Compare
to vessels in hardwood – not the same type of cell or function (photo by Jim Speer).

86
Pit

Figure 4. 8 Angiosperm cells types. Angiosperms have more complex cell types which are vessel
elements, tracheids, fibers, parenchyma, ray cells, and pits (graphic modified from Hoadley
1990).

87
phloem

Figure 4. 9 Dicotyledonous angiosperm wood structure. Angiosperms, such as this diffuse


porous wood, have different cellular structure than gymnosperms and are composed of multiple
vessels and more prominent rays. Ring boundaries, however, may still be defined by the size of
the fibers, size of the lumen, and the thickness of their cell walls (from Schweingruber 1996).

88
Figure 4. 10 Wood rays in an oak. Rays are wood cell structures that transport materials
horizontally (radially) through the tree and are most evident in ring porous and diffuse porous
tree species (photo by Jim Speer).

89
Forms of Wood Structure

Two main wood structures can be identified, which are the non-porous woods of the

gymnosperms and the porous wood types of the dicotyledon group of the angiosperms (Figure

4.11). The dicots are further broken into ring porous, semi-ring porous, and diffuse porous wood

types (Figure 4.12). In non-porous wood structure the ring boundaries can be identified by

examining the size and cell wall thickness of the tracheids. The presence of vessels differentiates

the gymnosperm wood from angiosperm wood. Vessels are an advanced evolutionary trait of

angiosperms that enable the trees to more efficiently transport water up the tree. They can be

large or small (Figure 4.13) and their distribution in the ring can help a dendrochronologist

identify the ring boundaries. Large vessels occur at the beginning of the growth ring in ring

porous genera. Small vessels may form as solitary individuals, as vessel multiples, vessel chains,

nested vessels, or as wavy bands and can occur anywhere in the ring (Figure 4.14). Ring porous

wood structure is defined by a row of vessels that are produced at the beginning of the growing

season before leaf-out. Because these vessels are formed early in the growing season when

photosynthates have yet to be produced, the tree uses stored reserves from the previous growing

season. Ring porous genera, e.g. oak (Quercus sp.) and ash (Fraxinus sp.), are the most distinct

of the angiosperms with an obvious row of vessels occurring at the beginning of the growth ring

(known as the earlywood zone) (Figure 4.15). Semi-ring porous genera, e.g. hickory (Carya

sp.) and elm (Ulmus sp.), have some vessels that form at the beginning of the ring, but also have

smaller vessels distributed throughout the ring (Figure 4.16). The earlywood zone is not as

consistent and distinct as with ring porous genera. Diffuse porous genera, e.g. maple (Acer sp.),

birch (Betula sp.), and aspen (Populus sp.), have small vessels distributed throughout the ring

that have no relationship with the ring boundaries (Figure 4.17). This varied distribution of

90
 

Gymnosperm

Angiosperm

Angiosperm

Figure 4. 11 Gymnosperm versus Angiosperm wood types. Gymnosperms do not have pores and
angiosperms have either ring porous with a row of vessels at the beginning of each ring or
diffuse porous woods that have vessels distributed throughout the ring. All of these tree samples
are growing from right to left in these images (photos by Jim Speer).

91
Figure 4. 12 Classification of ring porosity. Gradation from ring porous wood to diffuse porous
wood with semi-ring porous as an intermediate stage (graphic from Hoadley 1990).

92
Figure 4. 13 Relative size of hardwood cells and wall thickness in ring porous species. In diffuse
porous species, the disparity in size of the cells may be smaller (graphic from Hoadley 1990).

93
A B C

D E F

Figure 4. 14 Pore arrangement in angiosperms. Pores can be arranged in many different patterns
in angiosperm wood. This arrangement helps with wood identification. A) Acer with solitary
pores, B) Populus with pore multiples, C) Dyera with pore multiples, D) Ilex with pore chains,
E) Gymnocladus with nested pore clusters, and F) Ulmus with wavy bands. All images are at
15x magnification (graphic from Hoadley 1990).

94
 

Figure 4. 15 Examples of ring porous woods. Ring porous genera such as oak, (Quercus sp.),
ash (Fraxinus sp.), chestnut (Castanea sp.), and sassafrass (Sassafrass sp.) (graphics from
Panshin et al. 1964).

95
Figure 4. 16 Examples of semi-ring porous woods. Semi-ring porous genera include hickory
(Carya sp.) and elm (Ulmus sp.) (graphics from Panshin et al. 1964).

96
Figure 4. 17 Examples of diffuse porous woods. Diffuse porous genera showing no association
between the pores and the ring boundaries. Genera include maple (Acer sp.), alder (Alnus sp.),
birch (Betula sp.), and musclewood (Carpinus sp.) (graphic from Panshin et al. 1964).

97
vessels often obscures the ring boundaries, making ring identification in diffuse porous genera

particularly difficult. These trees also produce a large number of vessels, but the vessels have

nothing to do with the annual ring structure, are generally smaller, and can be randomly

distributed throughout the rings. In some cases, 80% of the field of view is taken up with these

vessels. Because the vessels are not associated with the ring boundaries, the cell wall thickness

of the fibers should be examined to determine the ring boundaries, similar to analysis of non-

porous species.

Reaction Wood

Trees growing on a slope or that are tilted will produce reaction wood to maintain or re-obtain

their vertical orientation. Gymnosperms produce compression wood on the downhill side of the

tree which is composed of thick walled and rounded tracheids. Angiosperms produce tension

wood on the uphill side of the tree with reinforced cell walls acting to pull the tree up straight

(Figure 4.18). The differing response causes the pith to be displaced upslope from center in a

conifer and downslope from center in a hardwood tree. In cross section, the cell walls may be

obviously thickened in compression wood (Figure 4.19), while in the radial view, spiral

thickening along the outer surface of the tracheids may be evident in tension wood only under

high magnification. I will demonstrate in Chapter 10 how reaction wood can be used to

determine the date of mass movements such as landslides.

98
Figure 4. 18 Reaction wood. Angiosperms and gymnosperms react differently to the pull of
gravity on a steep slope. Gymnosperms will produce compression wood on the downhill side of
the tree to effectively push the tree back up straight, while angiosperms will put tension wood on
the uphill side of the tree to pull the tree back up straight. Both of these types of reaction wood
produce larger rings while smaller rings are produced on the opposite side of the tree. This
reaction can be used to determine the date when a slope shifted causing the tree to react (from
Fritts 1976).

99
Figure 4. 19 Microscopic cross sectional view of compression wood in a conifer (right image).
Note the cell-wall thickening of this pine compared to the normal cells in the image to the left
(graphic from Hoadley 1990).

100
Growth Initiation and Absent Rings

Growth hormones (such as auxin and cytokinin) trigger cell division, cell elongation, and fruit

development. During years of good environmental conditions, growth hormones are produced in

abundance at the apical meristem and are transported down the stem in the phloem of the tree,

initiating growth all along the cambium (Figure 4.20). In stressful years, however, insufficient

growth hormone production may fail to initiate growth for some parts of the stem, especially

near the base of the tree (Figure 4.21). The results of this phenomenon are locally absent rings

that are only present in certain regions of the stem. Growth hormones tend to move from the tip

of the branches to the tips of the roots so that a tree is more likely to be missing rings near its

base.

Ring porous genera often produce vessels at the beginning of the growing season before leaf out,

suggesting that these vessels develop from cambial derivatives which overwintered in an

undifferentiated state. This phenomenon likely results in the observation that ring porous trees

usually do not produce locally absent rings. However, ring porous trees can produce rings so

closely packed together that it can be difficult (if not impossible) to differentiate the ring

boundaries.

Growth Throughout the Year


Trees continue to growth throughout the year, with different parts of the tree developing at

different times of the year (Figure 4.22). Krueger and Trappe (1967) examined growth in three

different parts of Douglas fir trees. Most stem diameter increase occured from March through

November, although some of that activity can be due to water draw up. Shoot elongation will

101
 

Figure 4. 20 The Auxin model of tree growth: the darker area at the top of the modeled stem is
new auxin production whereas the lighter shades of gray represent past years of auxin
production. Auxin is a growth hormone produced in the canopy of the tree. Auxin triggers cell
division in the canopy driving the production of tree rings through secondary growth. In
stressful years, not enough auxin is produced, resulting in a lack of secondary growth initiation,
causing areas around the stem to not form a ring during some years. This also explains pinched
rings around the circumference of a cross section (from Nogler 1981 as cited in Schweingruber
1996).

102
Figure 4. 21 Three dimensional ring production. It is important to think of tree ring production
in three dimensions. Each ring is formed on the trees like a sheath wrapping around the stem.
Based on the environmental conditions and the growth hormones produced in each year, a ring
may be absent around a cross section or vertically from one section to another. The absence of
rings is why crossdating is important for determining the complete chronology for each tree and
stand (from Stokes and Smiley 1968).

103
Figure 4.22 Tree growth throughout the year. Trees have the capacity to grow some part of the
organism in just about any month out of the year. The measurements were made on growth of
the stem, shoots, and roots on Douglas-fir trees (data from Krueger and Trappe 1967 graphic
from Fritts 2001).

104
frequently occur over a shorter period of time (May through August in this example), but root

growth can occur in just about any month of the year. This activity throughout the year, gives

trees the potential to record climate from many different times of the year.

Ring Anomalies

Rings are produced in many different forms that may confuse a dendrochronologist, but close

examination of a full cross section usually enables the appropriate identification of these problem

rings (Speer et al. 2004). Micro rings can be produced that are only two cells wide, with one

cell of earlywood and one cell of latewood in gymnosperms (Figure 4.23a). Micro rings are

difficult to find on a cross section, but a well sanded surface and the aid of crossdating can help

the dendrochronologists to locate them.

False rings occur when the limiting factors reduce growth rates and cause the tree to shut down

during some part of the year; but then that limiting resource returns and the tree continues to

grow. False rings can be used to record various environmental events such as severity of

monsoon events in precipitation-limited climates. In northern Michigan, false rings have been

documented as being more frequent in trees that are growing quickly and located in co-dominant

or intermediate canopy positions (Copenheaver et al. 2004). In most cases, these false rings can

be identified because the cell walls gradually thicken into a pseudo latewood, but then they

gradually thin back out (Figure 4.23b). If you follow an individual radial file (a row of cells

radiating out from the center of the tree that originated from an individual cambium cell), you

will be able to observe this cell wall thickening into the false ring and then gradually the cell

walls will thin back to earlywood cell widths. This contrasts with the abrupt transition in cell

105
   

C D

Figure 4. 23 Ring anomalies in Pinus occidentalis. Trees can produce a whole series of
anomalous ring forms that need to be properly identified for successful cross dating. A) Micro
rings may be only a few cells wide. B) False rings form when tree growth begins to shut down
because of limited environmental resources, but starts again because of the return of input from
the limiting factor. False rings can usually be identified in coniferous and diffuse porous ring
structure, because the cell walls gradually thicken, appearing to be latewood, but then gradually

106
return to normal earlywood cells. It is useful to follow a radial file of cells through this false ring
boundary. If you find a single radial file that does not complete the latewood but remains
earlywood cells through this boundary, then it is likely to be a false ring. C) Diffuse ring
boundaries arise when the normal process causing trees to go dormant for part of the season do
not occur. The pictured sample is from the Dominican Republic at 19.5 ºN Latitude where trees
are dormant during the January through March dry season, in other words, they stop growing
because of lack of moisture. If a year has unusually high precipitation during the dry season, the
tree is not forced to become dormant and continues to grow, producing a diffuse ring boundary
and no clear distinction from one year to the next. D) Pinching rings are produced when the
tree is damaged or nutrients are limiting so that growth is not initiated all the way around the
stem. E) Five normal size rings pinch to very small size and some disappear completely. You
can see that coring this tree at different locations will produce widely varying chronologies,
making the use of cross sections necessary instead of cores (photos by Jim Speer).

107
wall thickness associated with a true annual ring. In conifers, if you can find one radial file that

does not shut down completely, this is likely to be a false ring boundary.

Some trees (especially in tropical climates) form diffuse ring boundaries when growing

conditions are optimal and the tree is never forced to cease growing for part of the year.

Therefore, annual boundaries are diffuse without any real change in cell wall thickness between

years (Figure 4.23c). Often the ring in the second year is composed largely of latewood cells

because the tree never enters dormancy and it continues to produce thick cell walls in

anticipation of the end of the growing season.

Dendrochronologists prefer to have circuit uniformity in the cross sections of trees that they

examine. A well formed tree will have the same amount of growth around the circumference of

the cross section that is cut from a stump of a tree, so that taking a core sample from any place

around the stem will yield the same number and width of rings (See Chapter 5 for more details

on sampling methods). When viewing the cross section, it should appear as a bull’s eye target

with a regular pattern of rings around the center (like Figure 1.2). Many tree species do not have

circuit uniformity so care must be taken to collect measurements of what would be the average

amount of growth for each year. Teak (Tectona grandis) wood and some juniper species

(Juniperus sp.) produce a lobed growth pattern around its circumference so that it is difficult but

not impossible for the researcher to determine normal growth on the stem of the tree (Figure 4.1).

Some trees that do not exhibit circuit uniformity have pinching rings around the circumference

of the cross section. In this case, one or more rings will pinch out so that two different cores

from the same tree at the same height will yield vastly different ring counts (Figures 4.23d and

108
4.22e). If too many rings pinch out too frequently, it is very difficult, if not impossible, for the

dendrochronologists to locate these missing rings, even with crossdating.

These ring anomalies can occur in any of the wood types (non-porous, ring porous, semi-ring

porous, and diffuse porous), although I have never experienced absent rings in ring porous wood.

Trees with ring porous wood structure conduct most of their water in the vessels in that single

year of growth, although some water is transported in small-sized latewood vessels and tracheids

in a relatively (compared to gymnosperms) reduced area of sapwood. Rings can be very small

with little more than earlywood vessels produced over a series of years, making ring

identification difficult (Figure 4.24).

Cell division is continuous along a radial file, but can occur at different rates around the

circumference of the tree. The different rates of growth are not a problem when the ring is

continuous around the circumference of a cross section, but when rays interrupt the rings it is

possible for them to become misaligned. A dendrochronologist examining a series of rings in an

oak tree, for example, should match the width of the rings on either side of the ray before

visually crossing the ray to follow the rings on the other side (Figure 4.25).

Other ring anomalies may be found in the wood as well. Some of these are caused by

environmental conditions and others are caused by interactions with other organisms. In the mid

and high latitudes, Frost rings occur when the air temperature drops well below freezing during

the growing season. There are two competing hypotheses about how frost rings form. Some

suggest that frost rings can form when water freezes in the lumen of a cell and explodes the cell.

109
 

Figure 4. 24 Suppressed ring porous wood growth. Ring porous trees (in this case an oak)
always produce pores at the beginning of the growing season, but those pores may be packed so
tightly that the determination of ring boundaries is difficult. This section shows about 36 rings
(graphic from Baillie 1982).

110
 

Figure 4. 25 Offset of wood growth across rays. Radial files produce cells at their own rate.
When crossing a ray, it is possible for the ring not to be strictly aligned. It is, therefore,
important for dendrochronologists to make sure that they follow the same ring when crossing a
ray. This is done by a quick mental crossdating check to make sure that the ring widths are the
same size on either side of the ray for a number of the surrounding rings (graphic from Baillie
1982).

111
Later, as the cambium differentiates, these crumpled tracheids get crushed producing a

distinctive frost ring (Bailey 1925, Glock 1951) (Figure 4.26). Another explanation is that the

water in the stem near the ground or in the ground itself freezes, but transpiration from the

canopy continues to draw water, collapsing the outermost conducting cells that are not yet

lignified similar to the collapse of a straw when drawing on a thick milkshake. These distinct

frost rings can be observed in the wood of high elevation trees and, as we saw in the history of

dendrochronology, can become important marker rings.

Some aphids suck sap from sieve cells damaging the cambium and producing pith flecks. They

can be observed in the cross sectional view as a cluster of bubbly-textured wood (Figure 4.27).

This aphid damage cannot be used as a marker ring unless it results from an unusual outbreak of

the insect so that the damage makes a distinct marker ring that is synchronous between trees.

Fire scars are another distinctive anatomical feature that is caused by localized cambial

mortality due to the high temperature of the fire. Charcoal is not necessarily a part of the fire

scar as the scar is formed where the cambium was killed off, because of the high temperature of

the fire. The bark will often slough off after the cambium has been killed leaving exposed wood.

The living cambium on either side of the scar will then grow quickly, completing the scar

structure in the tree. Fire scars can be identified based on distinct cellular characteristics (see

Smith and Sutherland 2001). Living cambium cells on the edge of the dead cambium will

differentiate at an accelerated rate to cover the injured area and seal off the damaged wood

(Figure 4.28). This results in a distinctive growth curl after the fire scar occurred. Based on a

close examination of the area where the scar occurs in the tree ring, dendrochronologists can

112
 

Figure 4. 26 Frost Ring. Frost rings occur when freezing temperatures are reached during the
growing season. The cold temperatures make the water in the cell lumen expand and destroy the
integrity of the cell walls so that the cells become crushed (photo by Jim Speer).

113
Figure 4. 27 Aphid damage to cells of a red maple (Acer rubrum) tree. The aphids are active in
the cambium layer and move vertically up and down the tree, feeding on the newly developing
wood and damaging the meristematic tissue (photo by Jim Speer).

114
Figure 4. 28 Fire scar in ponderosa pine. From the left of the picture, the area where dead
cambium meets the living cambium is visible. The accelerated growth of the living cambium
cells creates the growth curl, healing over the injured area to the left (photo by Jim Speer).

115
determine the season of the fire, based on how much of the ring was developed before the injury

(See Chapter 8 for more details). Multiple fire scares in conifers usually occur on one aspect of

the tree because the wound caused by the first fire event makes the tree more susceptible to

scarring by subsequent fires.

Summary

With this basic knowledge of wood anatomy, cell structure, and tree growth, you can start to

analyze tree rings and to differentiate ring boundaries. We will explore sample collection and

preparation in the next chapter. You will find that the most important process for identifying

ring boundaries is to have a good polished cross sectional surface with which to work. The ring

boundaries can only be identified if you are able to see the structure of each individual cell under

the microscope. Therefore, before you analyze a sample you have to prepare the surface to the

best of your ability.

116
Chapter 5: Field and Laboratory Methods

Introduction

Good research starts with well-planned and executed field practices. Two important components

of field work addressed in this chapter are basic field methods for sampling dendrochronological

projects, and designing your sampling scheme so that you can accurately describe the patterns

observed in the environment. Field work will always bring up surprising circumstances and

unexpected situations that you will have to adapt to in the field. While this chapter describes

some basic field practices, modification of your sampling plans may be needed due to the field

area and its specific challenges.

Gear

Before going into the field, you must assemble the gear that will be needed for sampling. The

basic tools for dendrochronological fieldwork include at least two increment borers, straws in

which cores are stored, map tube for holding cores and straws, diameter tape, permanent markers

(like Sharpie© pens), golf tees or chop sticks for clearing wood stuck in a borer, a field notebook,

rope, compact drill kit-rifle cleaning kit, digital camera, global positioning system (GPS), and a

hand saw. Along with this basic gear, it is good to have a field vest to keep all of the gear

organized. With these tools, you can sample most basic dendrochronological projects and travel

relatively lightly. Your field equipment will change depending upon the project; for example, a

stand-age structure study will require two 100m measuring tapes to lay out plots, while fire

117
history may require a chainsaw to take cross sections. Table 5.1 gives a list of recommended

field gear.

A golf tee is the perfect tool for removing small pieces of wood that may remain in the borer tip

or for widening paper straws that might have been crushed. A bamboo skewer, chopstick, or

dowel can serve the same purpose. A rope may be necessary to remove a stuck increment borer

(see the Spanish windlass technique below). A digital camera is an important piece of equipment

to record the site characteristics, tree characteristics, and field methods. WD 40 should be used

to clean the increment borer, unless you are sampling for a chemical or isotope analysis. A

sharpening kit (with a small wedge stone and a cone sharpening stone) should be on hand for

sharpening dull or chipped increment borers. A cruiser pack (an empty backpack frame often

used while hunting) with bungee cords is an excellent piece of equipment to carry out a large

number of cross sections. Beeswax can be used to lubricate the increment borer when coring

hardwood trees. When the borer is just removed from a tree, it is warm from friction. The

beeswax can easily be melted onto the warm borer tip at this time. WD 40 can also be used as a

lubricant and to break down excess sap when coring pitchy pine trees but should only be used

when necessary. Fingerless gloves are recommended to obtain a better grip on the borer while

leaving your fingers free for the delicate work of packaging a core in a straw. Finally, always

carry lots of water, a first aid kit, and a two way radio for unforeseen issues that might arise in

the field.

It is always important to take enough field notes to provide a complete site description for later
publications that will come out of your work (see Appendix D for some sample field notecards).

118
Table 5. 1 Basic checklist of gear needed for dendrochronological sampling. Portions of this
equipment will be needed for different field projects, but this list should provide a good
foundation of the equipment that you might need.
Equipment Project Notes
Increment borer All Bring duplicates in case of breakage or jamming.
Map tube All For storage of samples
Straws All Paper or plastic
Permanent markers All Fine point and ultrafine point Sharpie© work very well
Diameter tape All Diameter at Breast Height (DBH) is a standard forestry measure that
we often take as part of our tree description
Masking tape All For joining plastic straws or making minor repairs
Hand lens All To examine rings in the field
WD40 All For cleaning increment borers and for lubrication
Beeswax All For lubrication and water barrier on borers
Hand saw All For taking small wood sections
Golf Tee, chopstick, or dowel All To remove pieces of wood from the tip of the increment borer
Long 9/16 inch drill bit with All For drilling out jammed wood. Do not use this from the cutting end
handle and be careful of the cutting tip
Rifle cleaning kit for a 22 All For cleaning increment borers. Paper towels and the spoon can also be
with cloth pads used to clean the borer shaft.
Weight lifting or bicycling All Fingerless gloves for hand protection while preserving the dexterity of
gloves your fingers
Rope All For starting a borer or to remove a stuck borer from a tree with the
Spanish windlass technique
Backpack All To hold all gear
Field vest All To provide easy access to the frequently used gear
Compass All For orienteering and field measurement
Knife All Always helpful, but don’t use on the increment borer tip
Nylon climbing rope All Useful for the Spanish Windlass technique or to help with laying out
plots.
Sharpening kit All For resharpening increment borer bits. Mainly used in camp on long
field trips or in the lab between trips.
First aid kit All Always have one on hand for minor injuries and possible broken limbs
Camera All Very important for documenting field sites and field techniques
Topographic maps All Important for mapping sample locations and for orienteering
GPS unit All Important for mapping locations of field samples
Chainsaw Fire History For taking larger cross sections
Chaps Fire History Safety protection for the legs
Helmet with ear protection Fire History Safety protection for the head
and face shield
Gloves Fire History Safety protection for the hands
Plastic wedges Fire History For keeping chainsaw cuts open when cutting whole sections
Scrench chainsaw tool Fire History Tool for work on the chainsaw
Round sharpening file Fire History For sharpening dull chainsaw blades
2 in1 fuel and oil can Fire History For carrying extra fuel and oil
Small pry bar Fire History To pry out cut cross sections
Plastic wrap or fiber tape Fire History To securely wrap fire history samples so that no pieces are lost and to
protect delicate samples
Cruiser pack with bungee Fire History Empty frame pack for taking out many cross section samples
cords
50-100m tape measures Stand-Age For setting up plots
Structure
Boomerang increment borer Optional For coring on trees with deep fissures in the bark. This is a homemade
handle item of a bent increment borer handle.
Increment borer starter Optional Helps you push the borer into the tree with your body mass

119
Much time and effort is often taken to get into the field and to locate a field site, so you should

collect as much information as possible while you are in the field. The vegetation should be

noted for the canopy as well as the shrub and even the herb layer as this understory vegetation

can often give more information about the long-term moisture conditions on the site. Slope,

aspect, and the location of trees relative to each other and prominent landmarkes are also

important pieces of information that should be recorded. A global positioning system (GPS) is a

good tool to locate and later map the locations of specific trees that have been sampled.

Occasionally the importance of indivudal samples will require that the samples be tagged with

permenant marking such as an aluminum tag that has been stampled with a specific sample

identification number. In this case, it is good to carry a lightweight hammer and nails. A digital

camera can also be used to collect data in the field and is a great way to record the appearance of

the sampled trees.

A good case study of proper field techniques can be observed with the current effort to extend

the bristlecone pine (Pinus longaeva) chronology further back in time (Tom Harlan personal

communication). Individual trees in the White Mountains of California can live to be over 4,000

years old. Many previous sampling trips have provided a very long chronology from these

amazing trees with such noted historical figures as Edmund Schulman, Val LaMarche, and Wes

Ferguson having taken samples from this area. Today, Tom Harlan is trying to extend that

chronology further back in time and is completing an exhaustive sampling protocol throughout

the high elevation zones of the White Mountains. To locate the oldest samples, the researchers

have documented the locations of past and current samples and mapped out the locations of old

versus young samples. Currently they are trying to increase the sample depth between 6,000

120
B.C. and 10,000 B.C. By relocating previously sampled trees of great age and locating remnant

wood in the correct time period, Tom Harlan has been able to continue to collect very old wood

and extend the chronology back in time. While working on this project the researchers have

found it difficult to relocate old sample due to poor field notes and proper archiving of those

notes and the samples. Because of these difficulties the current project is very aware of the need

for good notes. Every sample that is collected today is permanently tagged with a metal tree

identification tag that is nailed into the wood. A photograph of the tree or log is taken with a

white board stating the tree ID, its location in Latitude and Longitude (from a GPS

measurement), the date of the photograph, and the initials of the field team collecting the

samples. This quality of documentation ensures that the samples can be relocated in the future

and that this chronology can continue to be developed.

Paper or plastic straws may be used to protect the core. Paper straws are hard to find and do not

work very well under extremely wet conditions, but they allow the core to dry without molding.

Plastic straws are convenient because they can be found at any fast food restaurant, but care

should be taken to slit the straw so that the air can circulate. Masking tape may be used to hold

plastic straws together or longer clear plastic straws can be used for longer cores. The clear

plastic straws also allow cursory examination of the cores to see broad ring patterns or if the core

is broken in many pieces. Plastic straws can also be sealed with a stapler or melted shut with a

lighter depending upon your own preferences. Paper straws can be joined by pinching the paper

straw against the core and then sliding, with a twisting motion, a second straw over the first.

Plastic straws should be slit or a hole punch can be used to ventilate the cores so that mold does

not form. Whenever a core is packaged in a straw, the straw should be labeled with a site

121
designation (usually three letters) a tree number (usually two, sometimes three numbers), and an

A or B for the first and second core taken from a tree. For example, a second core taken from

the third tree sampled from Shakamac Park might be labeled SHA 03 B. The date, your initials,

the tree species, and any other relevant field notes can also be recorded on the straw. Following

the United States convention, the tree genus and species is noted with the first two letters of the

genus and the first two letters of the species, for example Pinus ponderosa is PIPO.

Site Selection

Site selection is the first important consideration in choosing where to sample (see the Principle

of Site Selection in Chapter 2). Often times the study area will be outlined by local land

managers or by the goals of the research. Once the study area is determined, specific sites need

to be chosen that will adequately represent the area and topic that is being examined. Individual

sites can be chosen through a random selection technique to represent the broader landscape or

targeted sampling can be used to explore specific signals.

Random Versus Targeted Sampling

When in the field, remember to use the principle of site selection and observe how the

environment is likely to affect the site on which you are working. Most science consists of

observing patterns, and from that, determining the process that drives that pattern. It is important,

therefore, to observe the patterns on the landscape and to document those patterns in your field

sampling. The sampling protocol may control what can be observed on the landscape, so

researchers should be explicit about their sampling regime. Often, random sampling is used to

facilitate extrapolation of conclusions to the broader landscape. Square or circular plots are

122
randomly located to sample a representative area of the forest type. Random sampling locations

in a field area can be determined either before you go into the field by using geographic

information system (GIS) or Excel, or while in the field using a compass bearing and random

number generator. Randomly choosing plots before you enter the field is useful because it

removes the bias of the observer who gravitates, however unconsciously, toward “good” trees.

Another advantage of choosing plots before going into the field is the ability to develop a

stratified random sampling regime so that samples are spread out over different vegetation types.

This method requires time spent in the field to locate the pre-chosen plots with a global

positioning system (GPS). It is also possible to generate random samples in the field by finding a

stand that you want to quantify, then randomly selecting a compass bearing and distance, using a

random number generator for a number between 1 and 360 degrees and then from 1-100m

distant. Once you locate the randomly generated spot, you can start your transect or plot at that

point.

In many of the applications of dendrochronology, targeted sampling instead of random sampling

is necessary. If your purpose is a climate reconstruction, the oldest trees located in the most

climatically stressful areas should be targeted. This is because not all trees and all landscape

positions record the same climate signal. We need to select trees that will be sensitive to climate,

record a coherent stand level signal, and have the longest record available. Although a few

young trees can also be sampled to make sure that the outer rings are well represented, because

older trees may be suppressed on the outside. For surface-fire-regime fire history reconstruction,

the specific trees recording the longest and most complete fire histories also need to be targeted.

In this example, a general reconnaissance should be conducted so that the researcher knows the

123
samples that are available in the field site. The trees that will yield the longest and most

complete fire history based on a count of externally visible fire scars and wood preservation,

should be sampled. Fire history in a stand-replacing-fire-regime can be sampled following the

methods of a stand-age structure in which the establishment date approximates the age since the

last fire (Heinselman 1973). Finally, if you are interested in gap dynamics in a dense forest, the

gap making trees need to be targeted to acquire death dates and trees immediately within and

responding to that gap should be sampled to record the date of gap occurrence.

Plots, Transects, or Targeted Sampling

Some basic decisions have to be made about how to sample the trees on the landscape. This

decision varies based on the research goal. Circular and square plots work well for sampling a

given area for stand-age structure. Circular plots are easy to set up from a given center point and

a known radius and require fewer decisions about whether a tree is considered in the plot or out

of it. Square plots are a little harder to lay out with tape measures and a compass, but result in

plots which have a well defined sampling area. Transects functionally become long rectangular

plots and allow you to sample across gradients (such as an elevation, aspect, or moisture

gradient). A nested band transect is useful for sampling stand-age structure. For this type of

transect, you can run a tape measure out 50m. Everything within 1m of either side of the tape

should be cored at ground level. To increase the sample depth in the older age classes, all trees

greater than 20cm diameter at breast height (DBH) within 2m of either side of the tape and all

trees greater than 30cm DBH within 3m of either side of the tape should be sampled as well.

These size categories will change depending upon the forest type being sampled and the purpose

of the study.

124
Coring a Tree

The height at which you core a tree is dependent upon the question that you are asking. If you

are interested in the exact age of the trees for examination of successional processes in a stand-

age structure, the trees should be cored at the base so that the sample is taken as close to the

point of germination as possible. This will yield the most accurate age of the tree. A number of

problems exist with sampling this close to the root collar of the tree. Many trees have lobate

growth at the base of the tree that is associated with root activity just under the soil. This

irregular growth could confound a climate reconstruction. Also, it is harder to core a tree at the

base where you are restricted to using your upper body strength to take a core. The increment

borer handle can also hit the ground while trying to core at the base of the trees. To avoid this,

one will often excavate an area at the base of a tree so that the handle can turn freely. Shorter

borer handles can also be used to get closer to the base of the tree or a borer handle can be bent

to create the “Brown Banana Boomerang Borer” handle (also known as the Quad B) that bends

back towards the operator and allows the person coring to get closer to the ground or to core

deeper in between large fissures in the bark. A power borer is also an option which uses a large

chainsaw engine connected to a drilling attachment that converts the motion of the chain into

torque like a drill. These machines can be dangerous as they create a lot of torque and they are

not sold in most stores.

Heart rot due to root disease, basal injury, or browsing by animals is more likely to be

encountered at the base of the tree than at breast height. Unless tree establishment dates are

needed, we usually take cores at approximately breast height (1.4 m) even though the initial

years of tree growth will not be represented because the tree would not have grown to breast

125
level in its first years. Coring at breast height is advantageous because the whole body can be

used to build momentum for coring, and the most common forestry measure in North America is

diameter at breast height, so that samples taken at this height can tie into the extensive data and

literature compiled by forest researchers.

The first question that most lay people and forest managers ask is whether coring the tree causes

damage to the tree. The simple answer is yes, coring the tree opens up the tree to pathogens that

can cause rot and discoloration in the tree, but the tree has natural defenses to combat injuries

where the bark is broken. Many conifer trees will exude pitch into the core hole, sometimes

even within a few hours, effectively sealing off the core hole. Angiosperm trees

compartmentalize the wound by creating a barrier that stops the spread of fungus once it comes

into the tree. The main issues associated with coring are that the researcher leaves behind a

whole in the tree and is likely to cause some local discoloration of the wood around the bore

hole. If the trees are of great economic importance, such as orchard trees, one can spray a

fungicide in the bore hole until bark grows over the opening, but this is costly and takes a lot of

time. It is possible to go back and find some of the original trees that A.E. Douglass cored in the

1920s and they are doing fine today.

Two cores should be taken from all trees sampled so that crossdating can begin at the tree level,

in other words, the two cores from the same tree can be dated and compared with one another.

More cores can be taken to obtain a solid core or to try to get older rings in a tree. When these

two cores are averaged together, we have a better estimate of overall tree growth. If the tree is

growing on a slope, the cores should be taken parallel to contour to avoid reaction wood in the

126
tree. Conifer trees will produce larger rings (compression wood) on the downhill side of the tree

to keep the tree growing upright. In hardwood trees, the larger rings (tension wood) develop on

the uphill side of the tree, therefore, a core taken parallel to contour avoids the larger rings of

reaction wood and represents the average ring growth at that height in the stem. Be careful when

two people take cores from the same tree at the same time. The cores should be taken at

different heights so that the increment borers do not meet inside the tree and damage each other.

To start an increment borer, push the bit of the increment borer into a fissure in the bark of the

tree as you turn the handle in a clockwise direction (Figure 5.1). The fissure in the bark gives

you a starting place and allows you to avoid coring through a thicker area of bark. Although, on

trees located near roads, grit may accumulate in the fissures which could dull the increment

borer. Starting a borer is an easy process in softwoods but can be exceedingly difficult in trees

with smooth bark such as sugar maple and beech or in hardwood trees such as oak or hickory

trees. To aid in starting a borer in hard trees, you can also use an increment borer starter, which

consists of a metal plate that can be positioned against your chest and a shaft that fits into the

opening on the increment borer bit at the handle. This allows you to push with your chest as you

turn the borer by hand. The starter also helps you to make sure that you are coring straight into

the tree, perpendicular to the stem. If you wobble as you core into the tree, you will cut an

irregular core until the shaft of the borer is solidly seated inside of the tree.

Once the borer is started, the borer handle simply needs to be turned in a clockwise direction

until the the tip of the borer has passed the center of the tree. You can measure how far you have

127
 

Figure 5. 1 Starting a borer. When starting an increment borer, push the borer into the tree with
equal pressure on the shaft of the borer as you turn it into the tree. Starting an increment borer is
especially hard on smooth barked hardwood trees. A starter may be used when coring a hard
tree. It is made of a metal plate that can be placed against the chest and a shaft that is inserted
into the increment borer bit at the handle. This allows you to push with your chest as you turn
the handle of the borer (from Jozsa 1988).

128
cored into the tree by holding up the spoon so that the knob on the spoon is at the borer handle

and see how far into the tree the borer has penetrated. With larger trees, this may take a second

person standing back to observe if the spoon makes it to the half way point into the tree. You

should always keep two hands on the borer handle and make sure that you provide even pressure

along the shaft. Do not bend the increment borer shaft. As you core into the tree, feel the

resistance to turning the increment borer. If the borer starts to turn easily, you may have cored

into a pocket of rot in the tree and are at risk of getting the borer stuck. At this time stop and

remove your core and then the borer from the tree. If, as you core into the tree, it becomes very

difficult to turn the borer, more than you would expect from the friction of having more of the

borer shaft in the tree, the core may be twisting up inside the shaft and you are at risk of a

jammed increment borer. Stop and remove your core. A jammed borer usually is the result of a

poorly sharpened bit but can be exacerbated by rot in the tree. It takes a lot of time to clear a

jammed increment borer, so it is better not to get to that point.

Testing for a Compressed Core

It is possible to check the depth of your core in the shaft of the increment borer to see if it is

jamming up. This procedure is only recommended when coring softwoods such as pine. If you

stop coring for any period of time in the hardwoods, the wood fibers relax back on the borer and

you risk breaking the borer when you start to core again. On a conifer, you can stop and push the

spoon into the shaft of the borer until you feel resistance on the spoon, which means that you

have hit the bark of the core in the shaft. Hold your thumb on the spoon at the opening of the

shaft, marking the depth of the spoon in the shaft. Then carefully extract the spoon, making sure

that you are not taking part of the bark with you. Put the spoon up to the tree bark along the

129
increment borer shaft (Figure 5.2). The distance from the bark to the handle of the borer should

be the same distance that you measured inside of the shaft for the depth to the core. If your

marking thumb is one or more inches from the handle of the borer, then you have a jammed core

and should remove the core and the borer immediately.

Taking and packaging a core

The increment borer bit cuts and pushes away the wood surrounding a pencil size core of wood

inside the tree, so that once the borer is completely turned into the trunk, the only area of the core

still connected to the tree is the inner disc of material just at the cutting tip (Figure 5.3). The

inside shaft of the increment borer is tapered to a smaller diameter at the tip, so that the spoon,

inserted into the shaft, is forced to pinch into the end of the core (Figure 5.4). When the

increment borer is then turned in the counter-clockwise direction, the core is broken off inside

the tree and can be extracted by pulling the spoon out of the borer shaft. At this point, the core is

placed into a straw to protect it and maintain the proper order of any wood fragments that come

out of the increment borer. The shaft of the increment borer should be used as a third set of

hands holding your core while you package it in a straw (Figure 5.5). Remove the spoon only far

enough out of the shaft to slide the exposed core into the straw. Pinch down the end of the straw

to seal the core in the paper straw or use masking tape to close plastic straws. Do not simply fold

over the end of the paper straw or leave the tape as a flagged end. That flagged end will become

stuck on other cores in the map tube and take up more space than is needed. The straw is then

labeled with the site designation, the tree number, and the side of the tree that the core is taken

from (usually coded as an A or B core). Once the cores are neatly packaged in a straw they

should be placed in a map tube to protect them from breakage or getting lost. Remember that

130
 

Figure 5. 2 Measuring for compressed wood in an increment borer. In pine trees, you can use the
spoon to measure if the wood of the core is binding up inside of the increment borer shaft. In the
image, you can see that the thumb marking the depth of the core in the shaft is about 2 inches
from the handle, when measured against the tree in image C. This means that the core has
twisted and jammed up inside the borer and should be removed immediately (photographs from
Henri Grissino-Mayer; Grissino-Mayer 2003).

131
Figure 5. 3 Coring a tree. As the borer is turned into the tree, it cuts away the wood around the
increment borer and compresses the wood away from the borer. The core stays in its original
orientation and fills the borer as you core into the tree so that when you extract the core, you
have a full sample of rings from the bark to the pith (from Jozsa 1988).

132
Figure 5. 4 Tip of an increment borer. The tip of the increment borer shaft is tapered so that the
diameter of the core is smaller than the diameter of the inside of the increment borer. This
allows the spoon to pass by the core and to pinch into the core near the pith of the tree (from
Jozsa 1988).
.

133
Figure 5. 5 Extracting a core. To extract the core from the tree, turn the borer a half turn in the
counterclockwise direction, which breaks the wood off inside the tree, allowing you to pull the core out with the
spoon. Once you extract the spoon with the core on it, slide paper or plastic straws over the core from the bark end.
It is easier if you leave the core in the borer so that only a few inches of the core can be seen while you slide the
straw over the core. The shaft of the borer acts as a third set of hands and reduces the possibility of losing pieces of
the sample (photo by Jim Speer).

134
plastic straws should be slit to ventilate the cores and to keep mold from forming. These cores

can be removed from the map tube at the end of the day and bundled together in newspaper or

with string so that they can air dry. Once the cores have been bundled in newspaper, I have

found that cores can be placed on the dashboard of the field vehicle to help them dry out.

Removing an increment borer from the tree

When removing the increment borer from the tree, turn the borer in a counterclockwise direction.

The borer should gradually come out of the tree as you turn it. A borer may become stuck in a

tree if left too long because the wood that was pushed out of the way will relax back on the shaft

of the borer. If this occurs, apply a sharp backward jerking force on the borer as you turn the

borer counterclockwise so that the spiral threads of the borer bite back into the wood of the tree

(Figure 5.6). As a last resort, a rope can be used to create a Spanish windlass to remove a stuck

borer. Note that the clip that keeps the handle of the increment borer connected to the shaft can

vibrate loose or come undone. Be very careful when pulling back on the increment borer handle

to make sure that this clasp is engaged or the handle can come off in your hands which could be

dangerous when coring on steep slopes. Some tape or a rubber O-ring can be used to make sure

that this clasp does not come loose.

Cleaning an Increment Borer

Increment borers are made out of metal and are susceptible to rusting. If the borer shaft becomes

rusty, the metal will be weakened and the cutting edge can be pock-marked from the break down

of the metal by the rust. To avoid these problems, increment borers should be cleaned with a

135
 

Figure 5. 6 Extracting the increment borer. Normally you can extract an increment borer by
turning the handle in a counterclockwise direction with equal pressure along the shaft of the
borer. When the borer becomes stuck in a pocket of rot, you have to pull and turn at the same
time while being careful that the handle clip on the borer shaft does not release resulting in a
backwards fall (from Jozsa 1988).

136
dewatering agent (such as WD-40©) and paper towels or steel wool. A .22 gun cleaning kit can

be used to clean inside the shaft of the borer, but one should take care not to push the brush to far

past the cutting edge or it may be damaged. I prefer using the spoon of the increment borer with

a postage stamp sized piece of paper towel wrapped around the teeth of the spoon. The cleaning

agent should be sprayed along the outside and inside of the shaft with special attention paid to

the tip of the borer. Then the paper towels can be run up and down the inside of the shaft. The

papertowel usually comes off of the teeth of the spoon and then can be pushed out the cutting

end of the borer (as long as the piece of towel is not too large). I repeast this process until the

paper towel coming out is clean. This cleaning process should be done at the end of each field

trip as a minimum and could be done at the end of each field day, especially if you are coring

trees that tend to be moist such as in the eastern deciduous forest in the spring.

Sharpening an Increment Borer

The tip of increment borers become dulled through regular use and could become chipped if they

encounted a rock or some metal. This cutting edge is the most important component in getting a

good straight core without many breaks or twisting. The tip should be inspected regularly

through a microscope or with a hand lense to check its condition. Increment borers can be sent

for professional sharpening, but they remove so much metal that this can only be done 2-3 times

before so much metal is removed from the tip of the borer that the borer can no longer be used.

A sharpening kit can be purchased from the forestry catelogs or at some better hardware stores.

The kit should include three whet stones (a rectangular, wedge, and conical stone) that are about

three inches long and some honing oil. Increment borers can be sharpened at the microscope so

that you can closely watch how you are affecting the cutting edfe of the borer or you can work

137
on the borer in your lap and check the cutting edge periodically in the microscope. Sharpening

the tip of an increment borer is a delicate process and it is best if you can get some instruction

from someone who is skilled at it, but I will describe the general procedure below.

The rectangular stone is only used in dire circumstances where the tip of the increment borer is

chipped. The tip of the borer must be worked back below that chip. Remember that the tip of an

increment borer is tapered so that the core that is cut is a smaller diameter than the inside of the

borer shaft. This allows the spoon to pass by the core and attach at the tip of the core so that it

can be removed. If you sharpen the tip of the borer back too far, you remove this taper and ruin

the borer. Put some honing oil on the whet stone and holding the stone perpendicular to the

shaft, rub the stone back and forth to grind down the tip past the bottom of the chip in the cutting

edge. If the chip is more than 1/8th of an inch deep, then the borer will not be able to be fixed.

Under normal circumstances with a dull borer that does not have chips out of the tip I start with

the wedge stone to sharpen the tip and the threads of an increment borer. The stone should be

help at a 37-45 degree angle to the axis of the shaft. A 37° angle will make a sharper borer, but it

will not hold its edge as long. A 45° angle will hold its edge longer, but it may be harder to start

in some wood types. The wedge stone should be used to sharpen the outer bevel of the cutting

tip while the borer shaft is continually rotated. You do not want to hold the borer steady and

work back and forth on one part of the cutting tip, because the cutting tip is a circle and you will

wear down one side. So through constant rotation of the borer shaft and swiping the wedge stone

in the opposite direction of that rotation, you can sharpen that rounded edge. You can apply

138
some pressure as long as you are careful to be consistent in the amount of metal removed around

the circumference of the cutting tip.

The conical stone is only used to remove the tiny metal burs that are bent into the shaft during

this sharpening procedure. Insert the conical stone into the end of the cutting tip and gently

rotate it to remove these burs. You are not really sharpening the inside of the cutting tip; only

the outer edge does the cutting and you do not want the conical stone touching all of the inside

edge of the cutting tip at any time. If you put too much pressure on the cutting tip and force the

conical stone into the tip, it will flare out the tip of the borer, ruining it. Only a little work with

the conical stone is needed and then the cutting edge can be checked for sharpness.

You can look at the cutting tip through a microscope and you should see the shiny metal surface

where you just sharpened around the cutting tip. To check the sharpness of the tip, you can use

many layers of paper towel and turn the tip of the borer on the towels. It should cut out a series

of small disks of the paper towel. If it does not, then the borer is not sharp enough. This test

should be done carfully, because if the borer is sharp and you don’t use enough paper towels,

you will cut little disks out of your finger which is not a pleasant experience.

The threads of the borers can also be sharpened in similar maner to the tip of the increment

borer. Use the wedge stone and constantly rotate the borer while running the wedge stone in the

opposite direction along the edge of the threads. You will need to do this on both sides of the

threads and remembers that increment borers have either two or three threads that will need to be

sharpened.

139
When you are done, both the cutting edge and threads should be sharp and you should take care

in handling the increment borer. It is easy to cut you finder tips on the sharpened threads. It is

also easy to damage you newly sharpened tip, so be carefull as you return the increment borer

shaft to the handle of the borer for storage and always be carefull when you are coring in the

field to make sure that grit and stones do not come in contact with the tip of the increment borer.

Spanish Windlass Technique for Retrieving a Stuck Borer

A borer may become stuck in a tree if you encounter a pocket of rot or if you leave the borer in

the tree for too long. If you cannot get the borer unstuck by pulling and turning the handle, you

can use a Spanish windlass to remove the stuck borer. I should note that this technique can be

very dangerous as much tension is put on the rope and the increment borer handle during this

procedure, so the utmost caution should be exercised. For this procedure, you need to have a tree

directly behind you and a rope. Take the rope and wrap it around the handle and clip of the

borer, making sure that the clip will not release prematurely. Then take the rope and wrap it

around the tree directly behind the borer. Bring the end of the rope back and tie the rope to itself

(Figure 5.7). You have now made one continuous loop of rope linking the tree and the borer. As

you turn the borer, the rope will twist and shorten, eventually providing the backward pressure

needed to remove the borer from the tree. There is a high amount of pressure pulling the borer

out of the tree, so when the borer bit gets into the bark, the borer will be forcefully pulled from

the tree. Hold onto the borer handle and be careful you do not get hit by the handle or the bit of

the borer. Make sure that you do not let go of the handle of the borer because the borer can fly

through the air in an uncontrolled fashion and possibly hurt someone or damage the increment

140
 

Figure 5. 7 Spanish windlass. The Spanish Windlass technique uses the force generated from a
twisting rope to provide backward pull on the increment borer, enabling you to remove a stuck
borer from a tree. You need to connect the rope around the handle of the increment borer and a
tree directly in line with the increment borer, then twist the handle of the borer counterclockwise.
The increment borer will come out of the tree quickly when most of it has been rotated out of the
tree, so be careful with this technique. You should keep a tight grip on the borer to control it as
it comes out so that you or the borer does not get hurt.

141
borer bit. Some researchers will release the tension on the windlass before the borer leaves the

wood, but be certain that the borer bit is in solid wood at this point so that you do not have to go

through the process of retying the windlass.

Laboratory Methods

Once the cores are brought in from the field in their paper or plastic straws, the laboratory work

begins. This consists of preparing the wood by drying, mounting, and sanding the cores, and then

analyzing the cores through such methods as skeleton plotting, and measuring. Chapter 6

discusses the analysis of wood using computer and statistical methods.

Preparing Core Samples

While the cores are still in their straws, they can be dried in a drying oven for 24 hours at 60°C,

for a week in a fume hood with continuous airflow, or for 20-25 seconds on high in a microwave

oven. If you are lucky enough to live in a dry climate, cores can also be air dried as long as the

plastic straws are well ventilated. If the core is immediately glued to a mount when it is still wet,
通⻛风的
it will develop cracks as it dries and shrinks, making it hard to be certain that no wood was lost

in the field. Drying the cores on too high of a temperature may cause some wood types to twist.

Also if the research project is examining wood chemistry or isotopic analysis, a high drying

temperature may volatilize some chemicals in the wood.

Once the core is dried, it is mounted on a prefabricated wooden core mount (see Stokes and

Smiley 1968; Phipps 1985 for a review of laboratory techniques). The best mounts are narrow

142
enough to view two mounted cores side-by-side in a stereozoom microscope at 20X

magnification. Professionally manufactured core mounts can be purchased that are made from

poplar wood and measure 1.25cm X 0.75cm X 1.2 m in size. The mounts have a half circular

groove routed into them to take the 4.3 and 5.15 mm cores that are the standard dimension of

increment cores. The cores should be mounted using water soluble white glue so that the cores

can be removed from the mount and remounted, if necessary, by soaking them overnight in a

water bath.

Mounting cores. Before the core is mounted, all of the information from the straw should be

copied on to the core mount. This should include the sample ID, the tree species, the date the

sample was taken, and the initials of the person taking the core. Once the mount is prepared a

line of glue can be extruded into the core mount groove and the core can be carefully mounted in

the groove. The core has two cross sectional views and two radial views. Imagine the circular

core squared on four sides: two opposite sides are cross sectional views and two are radial views

(Figure 5.8; see chapter 3 for a description of these wood sections). Care must be taken to mount

a cross sectional view facing up, otherwise the ring boundaries may not be evident after sanding.

The radial view of the core is often seen as being coarse because of the torn tracheids, or shiny

because of the side view of the long tracheids (Figure 5.9). Also, the tangential view can be

examined to align the tracheids so that their long axis is mounted vertically. String, binder clips,

masking tape, or heavy weights can be used to hold the cores in place as the glue dries (Figure

5.10). If you do not restrain the core, it will soak up moisture from the glue and curl out of the

core mount. The glue will usually dry in about two hours. I personally prefer using string as it

allows you to pull the core tightly into the core mount and is flexible enough to provide pressure

143
 

Cross sectional view

Radial view
Radial view

Cross sectional view

Figure 5. 8 Schematic of the different sides of a core. The circle represents looking at the end of
a core, where the core has two cross sectional views and two radial views. The researcher must
take care to mount a cross sectional view facing up in the core mount, or the rings will not be
clear (drawing by Karla Hansen-Speer).

144
A

Figure 5. 9 Core orientation. The core should be mounted with the cross sectional view facing
up. A) A core correctly mounted with rings facing up. B) An unmounted core. C) The arrow
points to an example of torn tracheid. The tracheids are torn on the radial view so that the
surface looks rough or shiny. The vertical fibers on the end of the core can also be used to
determine proper orientation of the core (photo taken by Tony Campbell).

145
A B

C D

E F

Figure 5. 10 Mounting cores. Once the cores have been dried, they are mounted on
prefabricated wooden core mounts with water soluable white glue. The information from the
straw is written on the side of the core mount (A). A thin bead of glue is extruded into the
groove (B). The core is delicately removed from the straw and pressed into the groove (C). The
core is then secured to the core mount to keep it from curling as it absorbs the moisture from the
glue (D). The string can be tied off and the mount should sit for at least two hours before it can
be sanded (E). Once the string is removed all of the cores are ready for sanding (F). String, tape,
binder clips, or weights can be used to hold the cores in place while the glue dries (photos by Jim
Speer).

146
wherever the core is broken. The string can also be reused many times. One drawback of using

string is it leaves fibers behind that can be observed under the microscope when the core is being

examined. For drying the glue quickly, some researchersa at the University of Arizona will

microwave their cores for 20-25 seconds which drives off the moisture in the glue causing it to

dry quickly.

Untwisting cores. Cores may become twisted from a dull or nicked tip of an increment borer. If
钝或 带切⼝口的
a twisted core is mounted without treatment, then the core will vary between the cross sectional

view and the radial view as you move along the core mount. You can use a Low Pressure Steam

Jet Generator (or a tea kettle with a molded aluminum foil spout to direct the steam) to moisten

and heat the core so that it can be untwisted (Figure 5.11). Gentle continuous pressure should be

applied to the core counter to the direction of twist while the core is moved back and forth

through the jet of steam. This process usually takes about 30 seconds for each twist. Be careful

not to burn your fingertips. The technique is especially difficult with short lengths of core.

Cores can also be microwaves with a wet paper towel for a short period of time to get the same

effect of moistening and heating the core so that it can be untwisted.

Sanding cores. Once the glue is dry, the cores needs to be sanded with progressively finer

sandpaper from ANSI 80 grit (177-210 µm) (mainly used for hardwoods), 120 grit (105-125

µm), 220 grit (53-74 µm), 320 grit (32.5-36.0 µm), and 400 grit (20.6-23.6 µm) (Orvis and

Grissino 2002) (Figure 5.12). The first sanding grit is used to flatten the core surface for

subsequent polishing and takes the longest. The progressive sequence of finer belts allows you

to efficiently remove the striations created from the previous sanding belt as you polish the core

147
 

Figure 5. 11 Untwisting cores. If a core is twisted (usually due to a nicked or dull increment
borer), then the cores can be straightened over a jet of steam (photo by Jim Speer).

148
80 Grit
120 Grit

220 Grit

320 Grit

400 Grit

Safety Goggles

Figure 5. 12 Sanding belts. Cores and cross sections should be surfaced using progressively finer
grades of sandpaper from 80-grit to 400-grit. You should also use a dusk mask, ear plugs, and
eye protection while using an electric sander (photo by Michael Glenn).

149
to a better finish. The final surface should be polished so that each individual cell of the cross

sectional view can be clearly seen under a microscope with 7-40 X magnification. A 4” X 24”

belt sander with a flat top for sanding cores is often used although some researchers use an

orbital sander and others even use a drill press with a sanding disc attached to it. The drill press

technique has merits because you can look down on the core surface as you sand it to determine

when it has been sanded enough. It is also possible to use a razor blade to surface the cores. A

sharp razor blade with a steady hand and polishing with superfine steel wool can create a clean

suface. This razor blade technique is particularly useful for dendrochemistry and isotopic

analysis where contamination from saw dust should be avoided. I personally use the belt sander

and invert it so that the belt faces up and clamp the sander handle to the table. This creates a flat,

stable surface on which you can sand the cores. It is a good practice to change the angle of the

core between sanding grits (sanding along the length of the sander then switching to a 45 degree

angle from the axis of the sander) so that you can see the striations from the previous belt (Figure

5.13). Once those striations are removed, you can move on to the next finer level of a sanding

belt. Hand sanding film at 30, 15, or 9 μm can be used to provide a finer polish to the finished

surface. Sandpaper with the ANSI grit rating is a general value of the roughness of the surface

even though many different sized particles may be used in the sandpaper. Sanding film with a

micron rating is made of particles with a specific size as determined by a geologic sieve, and

therefore, the sanding film provides a better surface than the equivalent sandpaper grit. You

should be carful not to sand the surface of a core down too far. About half of the core should be

left when all sanding is done, so that you have the largest area of wood to look at under the

microscope. The final polish on the cross sectional view is most important for allowing the

proper identification of ring boundaries. Some researchers use ethanol or isopropel alcohol on

150
 

Figure 5. 13 Sanding cores. Cores can be sanded on a belt sander (4 X 24” belt sander
recommended for the surface area that it provides) with progressively finer grits (50, 120, 220,
320, and 400 grits) to polish the cross sectional view until the individual cells are apparent under
a microscope at 40X magnification. In this picture the operator is sanding the core at a 45 degree
angle from the axis of the sander so that he can see the striations made from the previous grit.
Once those striations are sanded off of the core, it is time to move on to the next finer grit.
Repeated visual examination of the core helps determine when the core has been sanded enough.
Of course, the final test is to examine the core under the microscope (photo by Jim Speer).

151
pine trees to remove excess resin. The surface can also be buffed with suede leather but if the

surface is too polished, then it may be hard to make pencil marks on the surface. Many

researchers have experimented with wood dyes to bring out the ring boundaries, but in my

experience, a well polished surface (even on maple wood) is superior to any dye for the

identification of ring boundaries.

Preparing Cross Sections

Cross sections can be sanded with the same belt sander that is used on cores. First, the sample

needs to be securely mounted to the table. You can use four layers of masonite peg board as a

working surface. Cut dowels into pieces about one inch long to fit into the peg holes. Put

multiple cross sections on the peg board and use the short dowels to securely fasten each sample

to the board. These dowels, placed around the circumference of the section, will enable you to

sand the sample while it stays in place. You can also use a friction pad (rubberized pad) to hold

the sample in place. The same series of sandpaper used on cores is also used on cross sections,

but if the surface of the cross section is particularly uneven from the original chainsaw cut, you

may start with a coarse 50 grit (125-149 µm) sandpaper or cut a clean surface with a band saw to

remove the saw cuts. Once a flat surface is obtained on the cross section with a band saw or 50

grit sandpaper, continue to work through the finer grits of sandpaper in the same way you would

prepare cores. While sanding, always keep the sander flat on the sample and keep it in continual

motion. Do not start or stop the sander on the section because this will cause gouging in the

wood. The sander should be running when it is placed on the sample and running when it is

removed from the sample.

152
A thick gum eraser can be used to clean sandpaper belts, greatly extending their usable life

(Figure 5.14). These large erasers can be purchased at wood working supply stores. Hold the

eraser against the belt surface while it is running. The gum from the eraser clumps up the saw

dust and resin, removing them from the spaces between the sanding medium on the belts. This

cleaning should be done after the use of any grit belt before the belt is removed from the sander.

Analysis of Cores and Cross Sections

At this stage, the cores and cross sections have been prepared and are ready for visual analysis

and crossdating. The goal of crossdating is to assign calendar dates to each annual ring, and one

way to start is to mark a visual ring count of the decades on the wood. Use a #2 pencil to initially

denote the decades because you will probably need to erase and change them as you continue

analysis. You can start your inspection from the outside of the tree (bark side) if you know the

date of death or cutting, or the inside (pith) if the sample you are working with has an unknown

death date. Starting from the pith, cores can be marked from zero as the innermost ring of the

tree with every tenth ring marked with a single pencil dot to designate the decade year. Every

fiftieth ring receives two dots, a hundredth ring receives three dots, and a millennial ring gets

four dots (Figure 5.15). When the core is briefly scanned it is easy to count up the total number

of rings, and the dendrochronologist can refer to this relative time scale if there is any question

on the dating of the core. This is a conservative technique that does not assume an accurate date

of the wood until those marks are erased and real calendar years are marked on the samples.

Another technique is to start from the outermost, bark side ring and use it as an anchor in time

for when the tree was cored, counting back from that outermost ring and assuming calendar years

as you work backwards in time from the bark of the tree. This is a faster technique because it

153
 

Figure 5. 14 Cleaning a sander belt. Sander belts lose efficiency before the grit is worn down as
they become covered in a layer of resin and dust from the wood. You can clean sander belts
using a rubber gum eraser that can be purchased at most woodworking stores. The eraser pulls
the saw dust from within the grit of the sandpaper and clumps it together, shooting it from the
sander. If cleaned, sander belts can be re-used some 20 or more times before the grits wear down
(photo by Jim Speer).

154
 

Figure 5. 15 Marking the wood. A systematic method of marking the wood provides a temporal
frame of reference so that you do not lose count of the rings. One dot is used on every decade
year, two dots every 50 years, three dots every 100 years, and four dots every millennium. This
allows you to quickly scan the wood and determine the date anywhere along the sample. Micro
and missing rings are indicated with dots across ring boundaries (from Stokes and Smiley 1968).

155
does not require re-marking the wood, but it can be misleading because these are not truly dated

rings until the process of crossdating is complete. In this process, one dot is still used for decade

rings, two dots for 50 years, three dots for centuries, and four dots for millennia, but in this case

the millennium mark coincides with A.D. 2000 and the first century mark coincides with A.D.

1900. If you are working with dead wood with an unknown outer date, you still have to use the

relative marks from the inside of the core until the core is crossdated against a master

chronology.

Skeleton Plotting

Skeleton plotting is the basic technique invented by A.E. Douglass described in the early 1900s

and used by many dendrochronologists around the world for the first attempt at dating a sample

(Stokes and Smiley 1968). Most dendrochronologists today use the same plotting paper of five

squares to a centimeter as Douglass originally used. When two samples of wood are compared,

they may be growing at different rates which would prevent a productive comparison of these

two cores. Time can be put on a standard scale by using graph paper where each vertical line

represents one calendar year and the cores can be compared against each other. This can also be

thought of as a two dimensional plot with time on the x-axis going from old on the left to the

present on the right and an inverse scale of the narrowness of the ring on the y-axis and ranging

from 0 (for average) to 10 (for an absent ring). This also reduces the bulky sample to a concise

record of the narrow marker rings that can be compared between samples, stored for future

reference, and compiled into a master chronology. A marker ring is a ring that is consistently

narrow or has identifiable characteristics and is consistent between different trees. Graph paper

is used as the standard scale for comparisons between trees where each line represents one year.

156
A line is drawn for the years representing narrow rings that are responding to a limiting

environmental factor. Extraordinarily big rings can be marked with a “b” on the skeleton plot

and may be as reliable for crossdating as narrow rings. Dating by use of skeleton plots and other

methods to be mentioned later are much more efficient and quicker than measuring the rings and

relying on statistics to find the match. Visual dating allows the dendrochronologist to use all

aspects of the wood, such as color, latewood thickness, and marker rings, to determine the dating

of the sample of wood. The computer program COFECHA (discussed more in Chapter 6) was

created by Richard Holmes as a second check of the dating developed by a dendrochronologist.

Historically (up until the 1980s), most dates were independently checked by a second

dendrochronologist before they were assumed to be accurate and published. Today, the only

second check that we generally use is the COFECHA program as long as visual dating is done

independently of this statistical tool.

When preparing a skeleton plot, cut a sheet of 8 ½ X 11 inch graph paper on the long axis into,

thin strips that are 15 squares high (Figure 5.16). We try to use the same graph paper that

Douglass did in the early 1900s so that all of our plots can be compared to each other. This

graph paper has five lines per centimeter. One hundred ten years can be marked on this sheet

and additional sheets can be glued on to accommodate longer cores. The empty white margin of

the paper on your left should be used to write the sample ID, your name, and current date. The

top third (5 lines) of the paper should be marked with a regular count, either starting at year zero

on the left and marking every tenth year, or starting with the outside date on the right and

marking every decade going back in time. Either way, time is progressing from left to right. A

flag is used to designate the inside date and outside date of the core. These flags are important

157
 

Figure 5. 16 Making a skeleton plot from a sample of wood. The plot illustrates time on a
standard scale of each line representing one year. The more significant smaller marker rings are
represented by longer lines on the plot. Beginning and end flags are drawn to show the inside
and outside dates on the sample. Calendar years or a relative dating scale is marked along the
top of the paper with every 10th ring marked with a date. Begin plotting from left to right, pith to
bark. The sample ID, your name, and the date that the plot was created should be written in the
blank space on the left of the plot (From Stokes and Smiley 1968).

158
because they provide the establishment date (or inside most date) and death date (or sampling

date) respectively.

Because of the age-related growth trend discussed in chapter 2, some standardization process

must take place while making a skeleton plot. Otherwise, all plots would start off with no

narrow marker rings at the beginning and the length of the line (designating more narrow marker

rings) will gradually get longer towards the outside of the plot paper. While this is an accurate

representation of the growth curve of the tree, it does not provide useful interannual variability

for dating. To remove this trend and any possible suppression and release events from forest

dynamics, you should use a mental standardization process. Compare the ring that you are

dating to three rings on either side of it. If the ring on which you are focused is relatively narrow

compared to surrounding rings, it receives a vertical mark on the skeleton plot paper. Another

technique compares the ring in question only to the rings before it, i.e. those closer to the pith,

because this prior growth may affect the current year’s growth making them a more accurate

comparative pool. This also allows for autocorrelation as mentioned in Chapter 2.

The bottom 10 lines of the skeleton plot paper are used for representing the marker rings with a

line that is 10 boxes tall representing the narrowest possible ring in the core and a line that is one

box tall representing a ring that is only marginally narrower than average. Any ring that is of

average width or wider gets no mark on the plot. Rings that are significantly wide, however, can

be marked with a “b” on the plot designating them as a big ring. These wider rings can also be

used as marker rings in dating samples. Dendrochronologists usually concentrate on narrower

rings because wider rings are less noteworthy. The narrower the ring, the more significant that

159
ring is as a marker, resulting in a longer line on the skeleton plot. Many students making their

first skeleton plots are concerned about an absolute scale for the length of the line on the plots.

The length of the line is really an arbitrary designation that has to be determined by the person

making the plots. But after some experience with the range of possible ring widths, most plots

converge to a similar pattern. Cropper (1979) made a computer program that could use the

information from skeleton plots to date cores demonstrating that this is a repeatable process that

can be quantified. More recently, Tom Harlan has commissioned a new crossdating program

called Crossdate that makes electronic skeleton plots just as we do on graph paper and will

compare plots to a master chronology and provide statistics on the best matches. This program

was written for and is particularly useful when working on bristlecone pine samples that are

dated against a 10,000 year master chronology that is over 15 meters long when marked out on

graph paper.

Much of the inter-annual pattern used for dating is in the marker rings, but the spaces between

those rings also represent much of the pattern. It is important (because of the mental

standardization and for the precision of the dating pattern) that there are not any areas that have

four or more rings marked as small in a row. If more than four rings in an area are small, that

area is considered suppressed and only the smallest ring(s) in that area should be used as marker

rings. With four marker rings in a row, shifting the plot back and forth will match up marker

rings in four separate positions, removing the annual resolution needed for crossdating. On the

other hand, double and sometimes triple small years can be important dating markers.

160
When I work on a new site, I usually skeleton plot at least ten cores from that site. I compare

these plots to each other and determine if all of the rings are represented on each core by

matching up the marker rings. I start off by matching the two plots from within the same tree.

We expect dating within trees (from the two cores from the same tree) to be stronger than dating

between trees. If one or more of the plots do not match the others, I go back to the wood to find

where the problem in dating lies in that particular core. Every place where two plots disagree

(for example only one plot shows a narrow ring) I go back to the wood to check both plots. This

variation in plots may not be a dating error, but could be due to the individual ecological

response of the tree or even differences (such as compression wood) on the side of the tree that

you are examining. If many decades of marker rings are consistently off in one direction,

however, it is likely to be a dating error. The quality of the match is difficult to determine and is

something of an arbitrary determination, but repeated attempts to date a sample from trained

dendrochronologists produces the same results. The date of an unknown sample should be

checked through the length of the entire master chronology (the master plot for the site

containing the average widths for the narrow rings). An analyst will recognize some locations

where the marker rings match up better than others. I will often mark these dates on the plot as

possible dates, then go back to them at the end and determine which represents the strongest

match.

The result of these checks is a correctly dated set of cores from which a master chronology can

be built. As a first step in developing a master chronology, I overlap the individual plots, one on

top of the other so that they are very precisely aligned with all of the yearly line marks

corresponding to one another. The master chronology plot is made in a mirror image of the other

161
plots (Figure 2.3b). The blank space on the left side of the plot paper still contains information

about the site and master plot status. The dates, however, are real calendar dates that are listed

along the bottom of the plot paper. The lines representing the narrow marker rings now run from

the top of the plot paper down to a maximum of 10 boxes on the graph paper. The mirror-image

characteristic of the master chronology enables regular skeleton plots to be easily compared with

it in order to identify matching dates. A line is drawn on the master plot each time the ring in

question is represented on at least 50 percent of the individual tree plots. For example, if 1974

appears as a narrow ring in 5 out of 10 plots, that ring should be marked on the master

chronology. The length of the line on the master is calculated by taking the average length of the

lines represented on the individual tree plots. I do not count the cores not showing a narrow ring

for those years in the average, but I will make a slightly longer line if that ring is represented on

nearly every skeleton plot.

The master chronology, then, is a continuous time series containing all marker rings that agree

between trees for the length of the chronology. This is the best tool that you can use to date your

remaining samples.

List Method

The list method is another way to determine marker rings, if the outside date of the sample is

known. The analyst can count back the rings from the bark to the pith, marking calendar years

on the core according to the previously mentioned dot notation. Each time a narrow ring is noted,

the date is written in a vertical list under the sample ID (Figure 2.4). Once the researcher has

done this for five to ten cores, he or she can go back to the lists and determine which rings are

162
consistently narrow between samples. At this early stage, as with skeleton plots, one should be

careful that none of the samples are consistently off from the others which would represent an

initial dating error. Once a list of marker rings is developed, the analyst can use these marker

rings to quickly date the rest of the samples.

Memorization Method

The memorization method starts with known marker rings that may have been developed from

the skeleton plot or list method (Douglass 1941). The narrow marker rings can be memorized or

written down as a list. Newly surfaced cores should be counted back from the known outside

date, and the calendar years should be marked on the core using the dot notation. Every time a

narrow ring is seen, one should check it against the marker rings. If the ring should be narrow,

then the dating is still accurate. Continue this to the inside of the sample. If the dating of the

wood is off by a year or more from the marker rings, then the analyst checks other marker rings

in the sample. If these rings are consistently off in one direction from the master chronology, a

dating error has been located. The time period when the marker rings started to become different

from the master should be examined for possible micro, false, or locally absent rings.

I usually build my master chronology from ten cores using the skeleton plot method, and then

date the remaining cores using the memorization method. When building the master, I start with

the oldest cores in the collection so that I develop the longest possible master chronology

representing the entire length of the chronology. Skeleton plotting takes some time, but it is the

best technique for building a strong working master chronology, permanently recording that

master, and providing a basis for dating. The memorization method allows for quick dating of

163
the subsequent cores but relies on a valid master chronology. After measuring the tree-ring

widths, I second check all of my dates using COFECHA (see Chapter 6). The final check on a

master chronology is to check its dating against other master chronologies from the surrounding

area. The remote possibility exists that every tree in your chronology is missing a ring or that

you assigned the outside date off by a year. Comparison to another master chronology might

also help demonstrate crossdating in sections of your chronology with low sample depth. This

second check of the whole chronology against another master can confirm your dating. At that

point I am confident that my dates have no error and are accurate and precise with annual

resolution.

Measuring Methods

Most dendrochronology projects require ring width measurement for a quantitative analysis for

comparison with climate data or some other calibration data set. Other projects, such as

archaeological dating or fire history, simply require crossdating and do not need the samples to

be measured. One of the benefits of measuring all samples is that the program COFECHA can

provide the validation on the visual crossdating. These measured ring widths can also be

contributed to the International Tree-Ring Databank (see appendix E for web addresses) which is

a worldwide repository of tree-ring chronologies. This is also the location where you can find

other master chronologies to which you can compare your dating.

Measuring Systems: Many measuring systems exist that can be used to obtain accurate

measurements of tree rings. Most of these systems have a moving stage whose location is

determine by rotation of a lead screw or by an optical linear encoder. These systems include the

164
Bannister Measuring Stage, the Measurechron, the Henson Measuring Stage, the Zahn

Measuring Stage, the LinTab Measuring System, and the Velmex Measuring System. All of

these systems are used in conjunction with a stereozoom microscope supported by a boom stand

(Figure 5.17). The Bannister, Measurechron, Henson, and Zahn measuring stages all count the

number of rotations of the lead screw to determine the width of each ring. One drawback from

this type of system is that the screw can wear over time so that if the technician measuring a core

measures past the end of a ring, error can be incorporated in the measurement by turning the

screw back. This can be avoided by backing off from the ring boundary and measuring back up

to it. The Bannister, Henson, and Zahn systems are no longer made and it is hard to find

replacement parts for them. The Velmex Measuring System and the LinTab Measuring System

have a movable stage that is advanced by a lead screw connected to a handle but an optical linear

encoder actually determines the exact location of the stage and measures its position to an

accuracy of 0.01 mm, 0.002 mm, or 0.001 mm depending upon the precision of the instrument.

The microscope should have a crosshair reticle in one of the eye pieces and this crosshair should

be lined up with a ring boundary so that the vertical hair is tangent to the curve of the ring

boundary. Measurements should be made along a core or cross section perpendicular to the ring

boundary or along a radial file (a row of cells that are produced from the same cambial initials).

The average width of the ring should be measured based on the observable ring area. For

example, if a ring pinches across the field of view, the average width of that ring should be

measured. Because it is necessary to measure perpendicular to the ring boundary, the core must

be repositioned to take the curvature of the ring near the pith of a core into consideration (Figure

5.18).

165
 

Figure 5. 17 The Velmex Measuring System.


The Velmex Measuring System is the standard instrument for measuring ring width. It is a
movable stage, rigged with an optical encoder, and working in conjunction with a stereozoom
microscope that has a crosshair reticle in one eye piece. The print button is depressed each time
the crosshair lines up with a ring boundary, sending the measurement from the QuickCheck
device to the computer file. These measurements are retained in the virtual memory of the
computer until the file is saved to the hard disk (photo by Jim Speer).

166
Figure 5. 18 Measuring rings near the pith of a core.
Rings should be measured perpendicular to the current ring boundary at an area of average ring
width for that year. The core needs to be repositioned at the end of each measurement while you
measure rings near the pith that exhibit a distinct curvature (drawing by Karla Hansen-Speer).

167
Other measuring systems use digital images that are produced by scanning the sample on a flat-

bed scanner. WindDendro and LignoVision are two such programs. They both provide

automated measuring options that can speed the time it takes to measure samples. The draw

back for these systems is that the accuracy of the program depends upon the resolution of the

scanned image. Close supervision by the operator is needed to make sure that all of the rings are

accurately measured. Micro rings can easily be missed by these automated processes. Wood

samples that do not have very distinct ring boundaries such as diffuse porous woods are not very

well recognized by these system and the cost of the programs can be prohibitive.

Measuring rings. The best technique for measuring rings is to measure perpendicular to each

ring boundary. As a core approaches the pith, it often shows much curvature at the center, and

the core will have to be adjusted in between each measurement to stay perpendicular to the

previous ring boundary (Figure 5.18). The microscope that is connected to the measuring

machine must have a crosshair reticle in one of its eye pieces. This crosshair is the target that is

used to mark the ring boundary. The vertical crosshair should be tangent to the previous ring

boundary, and the horizontal crosshair should reach the next ring boundary without going off of

the wood sample.

Many researchers prefer to use a video capture system on a trinocular microscope to send the

image from the microscope to a monitor. Crosshairs can be attached to the video monitor by

using fishing line that has been colored black with a permanent marker. The image moves across

the monitor in real time so that measurements can be made on the video screen. This

arrangement reduces the eye strain of continually looking through the microscope for hours at a

168
time. Some resolution is lost between the microscope and the video monitor so this system is not

the best for very narrow rings. Also there can be some parallax between the crosshairs and the

image on the monitor so the technician has to remain still while measuring each ring.

Work Time Distribution

Each step of a project, from collection to analysis, takes a certain amount of time, and it can be

useful when planning a project to have an idea about the time one can reasonably expect to spend

on each part. The data collection process takes much less time than the laboratory procedures

(not counting travel time to the site). Crossdating takes the most amount of time, including the

check of the dating that can be done with the COFECHA program. Measuring the cores also

takes considerable time, but the analysis can progress relatively quickly once these steps are

completed. Table 5.2 lists the amount of person hours that it takes for each stage of a standard

project that is completed by a skilled dendrochronologist with relatively straightforward wood.

169
Table 5. 2 Average work time in hours to collect, process, and build a chronology that is from
200-400 years in length from 20 trees (modified from Fritts 1976).
Task Mean Minimum Mean Mean Maximum Mean Percentage
1. Collection 11 15 23 9
2. Specimen preparation 7 12 17 7
3. Dating 51 72 120 42
4. Measuring 30 39 53 23
5. Dating Check with COFECHA 10 17 28 10
6. Basic climate response analysis 3 5 8 3
7. Project Supervision 8 11 15 6
Totals 120 171 264 100 %

170
Chapter 6: Computer Programs and Statistical Methods

Introduction

Dendrochronology uses a suite of custom computer programs that incorporates both standard and

complex statistical routines and tools that facilitates crossdating, climate analysis and

reconstruction, biological response modeling, and tree-ring data editing. Many of these

programs were written beginning in the 1960s and 1970s and are, therefore, DOS-based

programs that run in a DOS shell in the Windows operating environment. Richard Holmes

rewrote these programs or wrote many new programs for the Macintosh operating system and

created the Mac-compatible Dendro Program Library (DPL), a set of routines that helps

dendrochronologists explore tree-ring data. Some programs have also migrated the other

direction. ARSTAN (a program that conducts autoregressive time series standardization of tree-

ring data) was initially written for the Mac by Edward R. Cook of Columbia University and then

ported to the PC in the late 1980s and early 1990s. Currently, the most up-to-date versions of

ARSTAN are made available to run on Macintosh computers first. Many other programs have

been developed for Macintosh computers, Unix systems, or the SAS statistical package, but I

will not describe those programs in this chapter. More information about these programs and

applications can be obtained through the International Tree-Ring Data Bank (ITRDB) computer

forum archives. Most of the programs mentioned below are free and can be downloaded from

Henri Grissino-Mayer’s Ultimate Tree Ring Web Pages (see appendix E for web addresses).

171
In the following sections, I describe some useful statistics followed by descriptions of the main

dendrochronology programs in the approximate order of their use. I explain the purpose of the

program and, in some cases, provide a keystroke tutorial that walks you through the execution of

the main programs. I also provide some basic interpretations that explain the output for the main

programs. Some of this information is published elsewhere in a different format and by different

authors. I cite these references at the beginning of each section so that the reader can also

examine those publications.

My intent in this chapter is to provide the basic tools needed to conduct analysis, not an

exhaustive description of the programs and their output. See the cited references for more

detailed description of the programs. Many of these programs can run as a black box using the

program’s default settings, where the user does not need to understand the internal (often

statistically complex) operations executed by the program. Please try to educate yourself as

much as possible about how each program functions and the proper parameters for the specific

project in mind. Also, look to some of the classic literature published on dendrochronological

methods, such as Fritts (1976) and Cook and Kairiukstis (1990), for further reading.

Statistics in Dendrochronology

Series Intercorrelation

In the case of dendrochronology a tree-ring series from one core might be correlated against the

master chronology or two cores can be compared to each other. The series intercorrelation can

be the average of every series back to the master chronology and in this case will represent the

172
common stand-level signal recorded for a site (Equation I) . It is calculated between two series,

such as a core (x) and the master chronology (y) using

t n

 ( x  m )( y  m )
t x t y

rxy  t 1
[I]
(n  1) sxsy

where, xt is the index value for a core at year t, yt is the index value for the master chronology at

year t, mx is the mean index value for the core, and my is the mean index value for the master, sx

is the standard deviation for the core, sy is the standard deviation for the master, and n is the

number of years being compared. This equation adjusts for the variance between the core and

the master chronology as well as simply comparing the size of the rings in each year.

Mean Sensitivity

Mean sensitivity is a measurement of the year-to-year variability in tree-ring width ranging from

0 to 1 (Equation II). If every ring were the same width, the series would have a mean sensitivity

of 0 and if every other ring were absent then the mean sensitivity would approach 1. For dating

tree rings, it is possible to have series that are too complacent and other series that are too

sensitive to date accurately. From personal experience, a series with a mean sensitivity around

0.1 is so complacent that it is difficult to date and a mean sensitivity of greater than 0.4 is so

sensitive that it becomes extremely tricky to date due to frequent micro or absent rings next to

very wide rings. Mean sensitivity around 0.2 is generally accepted as series that are sensitive

enough for climate reconstruction. The equation to calculate average mean sensitivity for a

series is

173
1 t n 1 2( Xt  1  Xt )
msx  
n  1 t 1 Xt  1  Xt
[II]

where, Xt is ring width in year t, Xt+1 is ring width in the following year, and n is the number of

years being compared.

Gleichläufigkeit – Sign Test

The Gleichläufigkeit (G) is a measure of similarity between two chronologies based on the first

difference between successive tree rings (Eckstein and Bauch 1969, Schweingruber 1988). In

other words, it tests to see if two chronologies are increasing in growth at the same time or

decreasing in growth at the same time. This examination of annual trends enables the researcher

to compare the trend of cores for dating as well as comparing the ring widths (Equation III and

IV). G-scores have been incorporated into the programs CDendro, CATRAS, and TSAP or they

may be calculated by hand.

 i
 ( xi 1  xi ) [III]

  0 :G
i ix
1
2
when   0 : G
i ix
0

  0 :G
i ix
1
2

1 n 1
then G( x, y )   
n  1 i 1 G ix G iy
[IV]

174
Figure 6.1 shows an example that compares two cores to calculate the G-values. Each core’s

increase or decrease is calculated for every year-to-year change, and then these G-values are

added together for each year-to-year change. This sum on an annual basis is then added up for

the entire length of the core and the result is the G-value between those two cores. For example,

if one tree is increasing in growth in the first year and the second core is also increasing in

growth for that year, the chronologies score a 1 for that year. If one tree is decreasing in annual

trend while the other core is increasing, then those chronologies score a 0 for that year. In the

end all of these annual scores are summed and in the case of the figure 7 out of 10 intervals are

trending in the same direction so the chronologies score a G = 70%.

Rbar

The running rbar is one statistic that can be used to examine the signal strength throughout the

chronology. It is calculated by taking the average correlation between all series in a 100-year

window with 50 years overlap, throughout the entire chronology. Because it is a running

correlation between series, it is a good measure of the common signal strength through time and

is dependent upon the sample depth (Cook et al. 2000).

Expressed Population Signal (EPS)

The Expressed Population Signal (EPS) is a measure of the common variability in a

chronology which is dependent upon sample depth (Equation V; Wigley et al. 1984, Briffa and

Jones 1990). Its formula is

175
 

Figure 6. 1 An example of calculating the Gleichläufigkeit value (from Schweingruber 1988). If


an interval increases for one core it receives a +1/2 value and if the same interval increases on
the second core it also receives a +1/2 value giving a total G-value for that one interval of 1. The
calculation is conducted for each interval with a constant value being equal to 0 and a decreasing
interval is scored as -1/2. All of these intervals are summed over time throughout the chronology
resulting in the overall G-value (from Schweingruber 1988).

176
t * rbt
EPSt  [V]
t * rbt  (1  rbt )

where, t is the average number of tree series using one core per tree and rbt is the mean between-

tree correlation. When the EPS value drops below a predetermined level, the chronology is

starting to be dominated by individual tree-level signal rather than a coherent stand-level signal

(Figure 6.2). The chronology can still be well dated and useful for dating studies such as in

archaeological research, but may produce large confidence limits in a climate reconstruction. A

value of 0.85 has frequently been used as an appropriate cut-off point. This chronology measure

is frequently used by European dendrochronologists and has recently come into use by American

dendrochronologists.

Subsample Signal Strength (SSS)

The subsample signal strength (SSS) is a measure of the amount of signal captured by a

subsample of cores out of some master chronology (Equation VI; Wigley et al. 1984, Briffa and

Jones 1990). This calculation enables the researcher to quantify the variance in common

between a subset of samples and the master chronology, which is particularly important as

sample depth decreases in a climate reconstruction further back in time. It is calculated by

t '[1  (t  1)r ]
SSS  [VI]
t[1  (t '1)r ]

177
A

Figure 6. 2 A) Running rbar and B) Running EPS analysis for the Newberry Crater Lava Flow
Ponderosa Pine Chronology (from Clark and Speer unpublished data).

178
where t′ is the number of cores (if one core per tree) or trees (if two cores per tree) in a subset of

the whole population, t is the number of cores or trees in the full set, and r is the mean interseries

correlation for the chronology.

Measuring Programs

Many programs have been developed to measure tree rings, such as TRIMS, MEDIR, PJK, and

MeasureJ2X. These programs take tree-ring width data that are fed to the computer through a

data recording box such as Accurite or QuikCheck. Each program has some good features and

many have some bothersome quirks, just like all computer programs. TRIMS, MEDIR, and PJK

have been used for decades and do an excellent job as an interface to record tree-ring widths.

They also have some minor data editing features that enable the user to correct measuring

mistakes while still at the measuring system. Currently, MeasureJ2X is the measuring program

that is recommended when buying a new Velmex measuring system, so I will explain its use in

greater detail below.

Measure J2X

MeasureJ2X is written in JAVA language and can, therefore, be used on Mac or PC. It is one of

the few programs that was written professionally so it does have a cost. It has a graphical user

interface (GUI) which presents a program that functions like most of the Microsoft package of

programs. One can open files, save them, start new ones, see statistics on cores already

measured, and do some editing of measurement files. It still has some limitations, such as the

179
inability to rename folders, so it is best to prepare that space before you start to measure (see the

MeasureJ2X User Guide which is available online through the Voortech website).

Keystroke Tutorial for MeasureJ2X. Set the core up on the measuring stage as described in

Chapter 5 under the section labeled “measuring rings.” MeasureJ2X has menu options of File,

Series, Options, Setup, or Help (Figure 6.3). To start a new series, go to the Series menu and

choose New. Figure 6.3 will then appear on the screen and request the series ID and start year

for that core. When entering series IDs, always enter the same length site and tree characters so

that later programs can differentiate between cores from the same tree. For example, it is

standard to have a three character site designation and a two digit tree number followed by an

“a” or “b” for two cores taken from the same tree. The program ARSTAN can then average

these core-level measurements together resulting in a tree-level chronology, but for that to

happen the standard site and tree mask (or code) must be adhered to. Once these data are

entered, click OK and go to the initialization of measurements.

The next window that comes up is the measuring window (Figure 6.4). The sample ID shows up

at the top of that internal window (in this case TES01) and the first year to be measured is

displayed on the screen as well (in this case 1895). The program needs to be initialized at this

time, which entails sending a beginning measurement from which all other measurements will be

calculated. This enables the user to reset the stage at any time and not need to zero out the

measurements. Click on the Measure button. A new window will pop up asking for an initial

measurement. It is good to reset the measurement to zero at this stage and click OK. A new box

will pop up saying what the initial measurement was and click OK on that as well. Now the

180
 

Figure 6. 3 Initializing a new series in MeasureJ2X.

TES01

Figure 6. 4 Measuring view in MeasureJ2X.

181
program is initialized and waiting for measurements. Turn the dial on the measuring machine,

which moves the stage and the sample, until the crosshair is at the next ring boundary. Push the

Print button on the remote, which sends the measurement to the computer screen. The display on

the computer screen of the Display Value will show the cumulative measurement for that core

and the measurement which is the width of the individual ring (usually in millimeters but can be

changed to inches). This procedure is repeated for each ring on the core. All of the

measurements will be displayed on the screen along with the year of each ring. The computer

will beep each time a measurement is entered and a second beep should be heard for each decade

year that is measured. It is important to pay attention to this second beep and to use the decade

years as landmarks so that a mistake in measuring can be identified within the decade being

measured.

Once all measurements are completed, it is important to note that the data are in the computer

memory and have not been saved permanently to the hard drive. The user should click the Done

button on the screen and then close out the measuring window by clicking the small x in the top-

right hand corner of the screen. Be very careful that the measuring window is being closed and

not the entire program window. If the program is closed the measurements are lost. Once the

measuring window is closed, the user can save the file by going to the drop-down menu in File

and clicking Save. At this point another series can be initialized by clicking Series and New (as

above) and when this file is saved, it will append these new measurements to the bottom of the

last file measured.

182
It is possible to Delete or Shift the series in the Measuring Window if a mistake occurs. To

delete one or more rings, highlight the ring width measurements in the Measurement column.

Click Delete and choose to leave the first year or the last year fixed in time. Any other edits

should be conducted in the EDRM or EDT programs (discussed below).

DPL

The Dendrochronology Program Library (DPL) is a compilation of DOS programs that have

been developed by multiple users and provides useful tools for dendrochronology. This program

library has also been ported to the Macintosh operating system. Historically it was a package of

31 programs, some of which have been so useful that they have been taken out and now stand

alone, including COFECHA, EDT (now called EDRM), FMT, and YUX (Figure 6.5). Many old

versions of DPL are being used in research labs today. The modern version of DPL contains 20

programs (Figure 6.6), many of which are useful for filling gaps in data, converting and

displaying meteorological data, or generating a climate reconstruction.

FMT

The FMT program enables the researcher to change the file format as well as do some basic file

reorganization such as putting the series in alphanumeric order. The first set of menus gives

options to change the format of the file between any standard dendrochronological format such

as compact, measurement (in 0.01 or 0.001mm precision), index, one column, or two column

(Figure 6.7). The compact format was developed when the computers that we used to measure

tree rings had very limited hard drive space. This format removes all of the spaces between ring-

183
 

Figure 6. 5 The Dendrochronology Program Library (DPL) version 1.24p contains 31 Fortran
programs that can be accessed by their three letter designation in this command line driven DOS
window. Many of the more commonly used programs have been extracted as stand alone
programs such as EDRM, COFECHA, YUX, and FMT.

Figure 6. 6 The Dendrochronology Program Library (DPL) version 6.07p contains 20 Fortran
programs that can be accessed by their three letter designation in this command line driven DOS
window.

184
Figure 6. 7 Formatting options in FMT.

185
width measurements. The computer can read the file based on space delimitation but the

operator cannot read the measurements in the file because they all run together. The

measurement format has also been called decadal or Tucson format. The file is presented with

each line representing a decade and the first column holds the ring width for that decade year.

The second column is the first year in the decade and this continues through the ninth year of the

decade in the last column. This format is easy for the use to read and determine the ring widths,

however, the decimal places have been removed to conserve space and the end of file marker

designates where the decimal should occur. A -9999 marker means the file has 0.001 mm

precision and the decimal place should be three spaces for the furthest right reported value. An

end of file marker of -999 means that the file has 0.01 mm precision and the decimal place

should be two characters from the right. The index format includes the ring-width index and the

sample depth that went into that calculation. It is also more compact than the measurement or

decadal format and is difficult to read. The second set of menus provides 23 options for

procedures that can be conducted on the series (Figure 6.8). I find option 17 for reordering the

series to be the most useful tool in this package.

COFECHA

Historically, dendrochronologists at the University of Arizona would date a sample of wood with

skeleton plots, remove their marks on the wood, and have a second dendrochronologist skeleton

plot the wood to check the dates. If their two dates differed, they would confer and find the

problem. This quality control and second check on all dates produced from the University of

Arizona helped to establish the reliability of dendrochronology as a dating technique. In today’s

research climate with expectations of high productivity, researchers don’t have the time to

186
Figure 6. 8 Twenty three separate functions that can be performed on data in the FMT program.
Option 17 allows you to reorder the series based on alphanumeric sequence of the core
identifications.

187
completely check each other’s dates, therefore, Richard Holmes developed the quality control

computer program called COFECHA (Holmes 1983).

COFECHA took the place of the second dendrochronologist as a quality control check on dating

of samples but it was never intended to be the only attempt to date a sample of wood or to

replace crossdating. COFECHA provides a statistical match between segments of each core and

the master chronology which is made out of the measurements that are entered into the program.

However, if half of the cores going into the program are not properly dated prior to statistical

analysis, the master chronology will be worthless; therefore it is essential that the researcher

crossdates the wood samples before using COFECHA. The worst part about using COFECHA

as the sole method of dating is that an operator will not know what good dating for that tree

species and site type looks like and will not even discern that there is a problem. The operator of

the program can manipulate the data and produce a chronology with acceptable statistics but the

resulting chronology is not necessarily well dated. The result will be inaccurate dates and poor

correlations with calibration data, such as temperature and precipitation records.

I attended a professional presentation where a researcher claimed that trees in a hardwood forest
阔叶林林

did not have any relationship with climate. When asked about his dating, the person replied that

he had not yet checked the quality of dating with COFECHA and gave no indication that the

samples were dated by any other means. Because of this lack of time spent dating the samples,

the research made an inaccurate conclusion, and extrapolated it to the hardwood forest. Dating

of wood samples is the heart of dendrochronology, and as such, dendrochronologists should do

everything in their power to properly date samples and check their dating quality. Two attempts

188
at dating are necessary to provide the quality control that has been the hallmark of good

dendrochronological research. The first attempt should include a visual dating method during

which the researcher learns the wood; visual dating can include skeleton plots, the list method, or

the memorization method from a known chronology. The second check on the dating can be

done by another research using visual dating or by a statistical check such as with COFECHA.

See Chapter 5 for detailed descriptions of skeleton plots, the list method, and the memorization

method for primary dating of samples.

COFECHA is a DOS program with all of its simplicity and quirks (Holmes 1983, Grissino-

Mayer 2001). It has also been ported to the Macintosh operating system. For those that have not

used DOS programs frequently, that means input file names can be no longer than eight

characters and cannot have spaces or non-alphanumeric characters in a file title, and the program

is command line driven (meaning you have to type your responses). It is best if the program

COFECHA is placed in the same directory as the files to be analyzed so that you do not have to

type in the directory chain each time the program is run. COFECHA leads the operator through

default options with most of the command steps. On the command line, COFECHA will often

provide answer options such as “<Yes>/No”. The option that is in brackets is the default option

and pressing enter will choose that answer. Proper use of these default responses can facilitate

efficient use of this program.

COFECHA works by statistically creating a master chronology with the cores that the operator

enters into the program. This means that if undated series are entered into the program then the

master chronology will be useless. COFECHA takes the ring width measurements that were

189
obtained from a measuring stage and, by default, fits a 32-year cubic smoothing spline to the

cores for standardization. Next, it averages all of the index series for all of the cores together to

create the master chronology. It then removes the core that is about to be analyzed, cuts it into

50-year segments with 25 years of overlap, and statistically correlates each segment against the

master chronology. If the correlation is below the specified confidence level, which is set at 99%

by default, then COFECHA checks from -10 to +10 lag years for a better match. If it finds a

better match it reports a B flag in the output; if it does not find a better match it reports an A flag

for that segment, simply meaning that it has a low correlation. This is the basic concept of how

COFECHA works and I will go through the specific key strokes in running the program in the

following section.

Keystroke Tutorial of COFECHA

To start the program, double-click on COFECHA.EXE in your directory (see Holmes (1983) or

Grissino-Mayer (2001) for more information on COFECHA). A DOS command box for the

program will open (Figure 6.9). The first entry that the program asks for is a five digit identifier

for your program run. This identifier will be tacked on as a prefix on any subsequent file created

by this program and should enable you to later (10 years down the road) understand what the file

contains. I usually use a three letter site designation with possibly one letter for species if I have

sampled multiple species on a site, and then a number at the end that can progress each time a

new run is started (such as MORQ1 for Mogan Ridge Quercus first run). You will find that you

will sometimes run COFECHA many times per site before you are done with the chronology.

The next question will ask you to enter the existing input file name. Remember that the file

name must be eight digits or less and not include any spaces or odd characters. COFECHA can

190
Figure 6. 9 Introductory screen of COFECHA in a DOS command line box.

191
read files in many different formats: compact, measurement, indices, Accurite measurements,

meteorological, spreadsheet, single column of values, two columns of values, or a user defined

protocol. COFECHA will automatically recognize most of these formats and ask you if it has

identified the correct format. Next it will ask for the file name containing samples to run as

undated tree-ring series. COFECHA can attempt to date undated series by breaking the series

into 50-year segments and statistically testing each segment against the master chronology but

not include it in the master chronology. The output from this option will show up as Part 8 in the

COFECHA output near the bottom of the .OUT file. Assuming that you do not want to enter

undated series into COFECHA, simply hit enter. The next option is a title for this run which can

be 36 characters including spaces and odd characters. The title should be an informative

description for each run. It is useful to type out the site name, species, and any other notes for

this run in the title. Remember that you might be looking back at these files in 10 years and not

have any idea what the very short file names mean. This is your chance to mark the file with

needed information to remind you of this research project at a later date. When you are done

entering the title for this run, hit enter to get to the next stage of the program.

The heart of COFECHA is the table that allows you to change the spline length for creation of

the master chronology, change the segment length and overlap, run an autoregressive model,

change the critical level of correlation (which will be based on your segment length or N), decide

whether to save the master dating series, list the ring width measurements in the output, list the

parts of the output to include, and decide whether to calculate absent rings in the master series

(Figure 6.10). The default options in this program are listed on the right side of the screen and

are usually applicable to most purposes. Richard Holmes tested a series of spline lengths in

192
Figure 6. 10 Command line driven DOS box for COFECHA showing the series summary at the
top of the screen, the line for Undated tree-ring series input file, the Title of this run line with
data entered, and the table that offers options to control the analysis.

193
creating the master chronology and found that the 32-year cubic smoothing spline is the most

appropriate spline length for enhancing the interannual variability that leads to strong dating.

The segment length is optimal for providing a high N for statistical tests and providing the

flexibility to pin-point where missing or false rings may occur in the chronology. These

segments are lagged, by default, at 25 years, again making it possible to pin-point dating

problems.

To make changes in the main table in COFECHA, use the program submenus that are keyed by

the number associated with the function. To adjust any of the values in this table, simply type

the number on the left side of the screen that correlates to that object. For example, to make a

change from a 50 years segment with a 25 year overlap to a 30 year segment with a 10 year

overlap, type “2” and hit enter. A submenu then opens and asks for the length of the segment.

Type “30” and hit enter. A prompt comes up asking for the lag between segments. Type “10”

and hit enter. Because the N has been reduced from 50 years to 30 years of comparison, the

critical level automatically adjusts from 0.3281 to 0.4226.

As I mentioned before, the default options in this table work well for most analyses. To run the

program, hit Enter once you have made any changes that you want in the table. The program

will then execute and very quickly display on the screen the progress of the program and finally

the correlation of each core with the master, as shown by a series of brackets where each bracket

represents a 0.05 overall correlation (Figure 6.11). This will flash by on the screen and the

program will exit itself. An output file with the result of the run is placed in the directory where

194
 

Figure 6. 11 The end of the information that flashes on the screen while COFECHA runs. In the
top half of the screen is the summary of how each core correlated with the master chronology
and the graphical representation is in the right column. The box in stars contains the summary
statistics for the chronology that will also be reported on the first page of the COFECHA output.

195
you ran the program. The file will begin with the prefix that you entered at the beginning of this

run, the three letters COF to designate this as a COFECHA file, and the suffix .out.

Reading the Output of COFECHA

The output file contains all of the summary statistics about the master chronology, the correlation

of each core with that master, and some descriptive statistics for each core. The first page of the

output provides the program name and version, the date of the run, the title that was entered, as

well as the file name used in the analysis, the parts included in the output, and the control options

that were selected during the run (Figure 6.12). The bottom half of the page contains the

summary statistics for the chronology, starting with the time span of the master chronology, the

entire continuous time span for the chronology, and the portion of the chronology with a sample

depth of two or more series. Next, COFECHA provides a warning of any rings that are inserted

as absent on only one series.

The table bracketed by stars in Part 1 is the most important summary of the COFECHA run,

much of which should be reported in a research paper. The table presents the number of dated

series, the master chronology length, the total number of rings in all series, and the total number

of dated rings (as in those that overlap with at least one other chronology), the series

intercorrelation, the mean sensitivity, and the segments with possible problems. The series

intercorrelation is a measure of the stand-level signal and mean sensitivity is a measure of the

year-to-year variability in the master chronology. These two statistics should be reported in any

publication as they are the most comparable measure of site-level signal and sensitivity between

196
 

197
Figure 6. 12 COFECHA output page 1.
sites. These statistics are described in greater detail in the sections below. Finally, a complete

list of any absent rings is listed by core.

Part 2 of the COFECHA output is a graphic representation of the length of the chronology

(Figure 6.13). It also summarizes the sequence number (which can provide an easy way to

navigate the COFECHA output), the beginning and end years of each series, and the total

number of years in each series.

Part 3 contains the master chronology in a three column format, including the year, the index

value, and the sample depth (Figure 6.14). Remember that COFECHA standardizes the series

with a cubic smoothing spline of your choice or the default 32-year spline. Each core is

standardized and then the master chronology is created by averaging together the index series for

each core. Because the master chronology in Part 3 records the index value for each year, it can

be used to identify extremely small rings that may be missing in other cores. I find the sample

depth to be the most useful column in this part of the output because it is the only place where it

is recorded for the master chronology. This information can be used to determine where your

master chronology has enough samples included in the chronology to provide an accurate stand-

level signal. A general rule is to have a minimum sample depth of 10 cores for a well-replicated

stand-level signal, although 20 is more robust and chronologies have been used with fewer cores.

A site’s sensitivity to climate will determine how many cores are necessary to average out the

individual tree-level noise and to reinforce the stand-level signal. Earlier in this chapter I

described the Expressed Population Signal (EPS) which is a statistical measure of adequate

replication. In order to maintain sufficient sample depth, many years often have to be cut out of

198
 

199
Figure 6. 13 Part 2 of COFECHA lists and graphically depicts the length of each core.
200
Figure 6. 14 COFECHA Part 3 shows the index values and sample depth for the master
chronology.
the final analysis because there are not enough overlapping series to demonstrate a stand-level

signal. It is important to note, however, that as long as there is agreement of dates between

series (at least two cores for a time period) then it may be possible to record events at the tree

level for those cores and have some confidence that the dating is accurate.

Part 4 is a graphical representation of the master chronology and is a good way for quickly

observing the dates of narrow rings and determining how small the rings are compared to the rest

of the chronology (Figure 6.15). An @ symbol indicates an average ring, upper case letters are

wider than average rings, and lower case letters are smaller than average rings. The further up

the alphabet, the larger or smaller the ring is. In the example given in Figure 6.15, a lower case

“s” for 1652 means that this ring is extremely small and is actually one of the smallest rings in

the chronology. An uppercase “C” at 1670 means that the ring is larger than, but not much

different than the mean.

Part 5 summarizes the correlation of each segment against the master chronology (Figure 6.16).

Remember that to date the core, COFECHA took all of the measured series and, by default,

broke them into 50 year segments with 25 years of overlap. Then it statistically correlated each

segment to the master chronology, minus the core being analyzed, at the date those rings were

assigned. This section of the output reports the correlation of each 50 year segment to the master

for each core with 25 years of overlap in the segments. COFECHA also marks poor correlations

with either an “A” or a “B” flag. An “A” flag next to a segment correlation means that segment

correlated below the critical level designated in this COFECHA run. A “B” flag means that

COFECHA found a better correlation for that segment in a 20 year window of -10 to + 10 years

201
 

202
Figure 6. 15 COFECHA Part 4 provides a graphical representation of the master chronology.
The @ symbol is an average ring, uppercase letters are larger than average rings, and lower case
letters are smaller than average rings. The further up the alphabet, the larger or smaller the ring.
203
Figure 6. 16 COFECHA Part 5 shows the correlation of each 50-year segment to the master. An
“A” flag means that the series dated best where it was, but the correlation was below the critical
level defined on page 1, and a “B” flag means that segment correlates better within a 20 year
window of where it is currently dated.
from the place where it is currently dated. This section is useful because it displays the

correlation of each segment to the master and is the only place in COFECHA where all of the

correlations for all of the segments are reported. Part 6 will report poorly correlated segments,

but not the results of the highly correlated segments. One has to refer back to part 5 for that

information.

Part 6 presents each core, one at a time, with a closer look at how well it correlates to the master

and reports any measurements that are outliers (Figure 6.17). The core name is given in the top

left corner and a series number is assigned to each core based on its order in the file. These

series numbers can be used to find the cores more easily in COFECHA or other

dendrochronological programs such as EDT (also called EDRM). Section A in Part 6 is printed

only when segments have a low correlation (as shown by an “A” flag in Part 5) or a better date

somewhere else in the 20 year window around the present date for the segment (as shown by a

“B” flag in Part 5). Section B is always presented showing the five years that added the most

weight to the correlation, labeled “higher”, and the five years that lowered the correlation the

most, labeled “lower”. This section also provides the correlation of each series to the master.

Section C presents any year-to-year differences (such as an acute increase or decrease in growth

from one year to the next) that were unexpected based on the master chronology. Any absent

rings in a core will be presented in section D, along with a comparison to what the master shows.

Section E presents any ring width measurements that are more than three standard deviations

from the mean. Because environmental effects on the trees are likely to cause rings that are

larger or smaller than the mean, I am concerned mainly when rings are five or more standard

deviations off of the mean; then the measurements should be rechecked for human error.

204
 

205
Figure 6. 17 COFECHA Part 6 provides core-level analysis of how well each core dates against the
master, shows the 20 year window of possible other dates for problem segments, shows the effect of
the best and worst segments on the overall correlation, and shows any outliers that are more than 3
standard deviations from the mean so that those measurements can be checked for accuracy.
The four cores presented in Figure 6.17 all date well, although the middle two (LCW01B and

LCW02A) have been flagged with a “B” flag and “A” flag respectively in Part 5. Core

LCW01B correlates better at a -1 shift for the 1875-1924 segment. If the outside (bark side) is

the known date of coring and should not be shifted, then the segment to be shifted falls in the

middle of the series and necessitates not only a missing ring near the modern part of the segment,

but also a false ring near the older part of the segment. Although the occurrence of both missing

and false rings in a 50 year segment is certainly possible, note that the correction would only lead

to a 0.029 increase in correlation (from 0.321 to 0.350), a small increase for a lot of change in

dating. This kind of flag obliges the researcher to go back to the wood. In this case after

checking the sample, I ruled this a spurious correlation and left the core the way it was dated.

Core LCW02A has an “A” flag because the segment from 1450 – 1499 correlated at 0.29 with

the master which is below the critical level. Again, after checking the growth on the sample and

seeing nothing anomalous, I left this core alone with no correction. The four cores shown on this

page of Part 6 correlated with the master at 0.622, 0.560, 0.558, and 0.524, which are good

correlations for this site. It is important to realize that there is no specific threshold that will

guarantee that a core is well dated or not. The correlation depends upon the species being

analyzed and the site characteristics. Once a researcher has worked in a region and with a tree

species for some time, one can learn what a good score is and use that for a benchmark.

In contrast to the case of LCW01B in figure 6.17, LCW10B in figure 6.18 has three segments in

a row that suggest a clear -1 shift: they are at one end of the core and the correlation of each

overlapping segment increases dramatically with the shift (for example from 0.16 to 0.70 for

1788-1837. This type of pattern clearly designates that there is a missing ring in the series and

206
 

207
Figure 6. 18 A second page from COFECHA Part 6 showing when a core has a missing ring.
Note that LCW10B, also called series 16 by COFECHA, has three inside segments that correlate
better if they are shifted back 1 year in time. If we assume that the outsides of the core is
anchored in time by the coring date, a negative shift usually means that a missing ring needs to be
inserted and a positive shift means that a false ring needs to be removed.
that ring will be in the area around 1825-1850 because the 1850-1899 segment dated without any

problems (check Part 5 to see this correlation). The other series on this page date well with the

master chronology.

Part 7 summarizes all of the descriptive statistics for each core including its correlation with the

master chronology and its mean sensitivity (Figure 6.19). At the bottom of the chart, we see

again the average of all of the series intercorrelations (0.520) and the average of the mean

sensitivities (0.254) that were reported in Part 1.

Conclusions from COFECHA

COFECHA is one of the most useful programs in dendrochronology and it can provide standard

statistics which enable researchers to compare between sites and species. It is often misused as

the sole dating method for samples and care should be taken to mainly use it as a quality control

check on previously dated samples. The COFECHA program is designed to assist in dating and

to develop individual series that are well dated. In a later section, I will describe ARSTAN

which is a much more powerful program that is used for chronology building. The tree-ring

series that are vetted in COFECHA will be input into ARSTAN for final chronology

development. On the way to chronology development COFECHA is used iteratively with

EDRM (meaning Edit Ring Measurements) to make corrections of problem segments that are

identified in COFECHA and confirmed on the wood.

208
209

Figure 6. 19 COFECHA Part 7 provides a table of the descriptive statistics for each core including
the sequence number, sample ID, start and end dates, number of years, number of segments, number
of segments with flags, and then statistics on each core for the ring width series before and after
filtering. This last part shows the autocorrelation and how it was removed from the series.
EDRM

EDRM (used to be called EDT) enables you to edit ring width measurements (Figure 6.20) and is

most often used after COFECHA has identified some sections of a core that need to be corrected

or eliminated. EDRM and COFECHA are often run many times to correct a series. EDRM

takes an input file name and accepts the standard dendrochronological file formats such as

compact, measurement (with 0.01 or 0.001 mm precision), indices, Accurite measurements,

meteorological data, spreadsheet data, or one or two column data. It asks for an output file name

so that a new file is always created instead of overwriting old data. This is a good safety

procedure so that original data files are not accidentally corrupted. The output file can be in

compact, measurement at 0.01mm precision, or measurement at 0.001mm precision. The

program asks if you want to use the first line of data as a header or title for the file. As with all

of the DOS programs, the option in <brackets> is the default response. The program then takes

one core at a time and allows the user to conduct various procedures on that core such as copy as

is, insert a value, eliminate a value, change the first year of the core, cut the core from the

beginning or end, omit a series, or change the core ID.

ARSTAN

ARSTAN is one of the main programs in dendrochronology that is used to build the final stand-

level chronologies. ARSTAN differs from COFECHA in that it has a broader range of

standardization techniques that can be used on individual series before a master chronology is

compiled. This should not be confused with the master chronology that is developed in

COFECHA. COFECHA also uses standardization (usually a 32-year cubic smoothing spline) to

210
Figure 6. 20 EDRM showing the options for editing a file.

211
create a master chronology for the dating of other cores. This master chronology though was

created specifically for dating purposes and is not the master chronology that should be used for

the final analysis. In ARSTAN, I will describe how different standardization techniques can be

used to maximize the signal of interest and remove noise from the final chronology. It was

developed to be able to mathematically standardize tree-ring series and to remove or control the

autocorrelation component in the time series (Cook 1985, Cook and Holmes 1986). The

program fits a curve to the measurements from each core, divides the ring width by the modeled

curve value, averages together the resultant index for each core to create a tree-level index, then

averages together the tree indices to develop a stand-level chronology (Figure 6.21; Fritts and

Swetnam 1989).

Historically, a negative exponential curve was considered a conservative standardization

technique because it removed a known age-related geometric curve from the ring width series. A

negative exponential curve describes the decreasing thickness of rings from pith to bark that can

develop in open grown pine trees putting the same volume of wood each year on an ever

increasing cylinder. More recently dendrochronologists have come to realize that this curve

works best where the trees are open grown and do not experience many disturbance events.

Cook (1985) demonstrated the need for more complex standardization techniques in closed-

canopy forests that have more stand dynamic signal than open grown forests. Cubic smoothing

splines take into consideration autocorrelation (the effect of previous growth or climate on the

current year’s growth) and Cook (1985) suggested the use of cubic smoothing splines as an

empirical fit to the growth of the trees (Figure 6.22). Today these spline fits are commonly used,

but too often they are applied with little rigor; it is essential that the researcher know what signal

212
 
A B C D
CPN01A CPN01A

CPN01
CPN01B CPN01B

CPN02A
Ring width (mm)

CPN02A
CPN02 CPN01A

CPN02B
CPN02B 1800 1900 2000

CPN03A CPN03A

CPN03
CPN03B CPN03B

1800 1900 2000 1800 1900 2000 1800 1900 2000

Figure 6. 21 Standardization and tree-level index series. ARSTAN matches a curve to the
individual series, as seen in column A, with a negative exponential curve used for cores CPN01A
and CPN01B, a straight line fit for cores CPN02A and CPN02B, and a decreasing trend for cores
CPN03A and CPN03B. Note that in this case the site ID is CPN the tree ID is tree number 01,
02, and 03 respectively, and the last digit (A or B) stands for the first and second core from the
same tree. ARSTAN then takes the ring width and divides it by the model fit (from column A)
and the resultant series is plotted in this column as a dimensionless index value with the average
of the series drawn as a straight line of value 1. ARSTAN then averages the index series from
the two cores from each tree together to create a tree-level index series in column C. Finally, the
tree-level series are averaged together to produce the stand-level master chronology in column D.
The intermediate step of the development of a tree-level series avoids the circumstance of
overrepresentation of one tree in the master chronology from which multiple cores may have
been taken (modified from Fritts and Swetnam 1989 p126).

213
Figure 6. 22 Examples of four tree ring chronologies that have been standardized using a 15-year
cubic smoothing spline. This was a very flexible spline that was used to remove as much climate
as possible to enhance a mast (synchronous fruiting) signal in oak trees from the southeastern
United States. This graph shows the original ring width chronology, with the model fit on top of
that chronology, followed by the resultant index chronology for four trees (Buell Plot Tree 50,
47, 10, and 2) (From Speer 2001).

214
is being removed and what signal is being kept in the resultant chronology (see Chapter 2 for

more description about standardization and spline length choice).

ARSTAN was first developed for the Macintosh operating system and continues to have the

most features. It has been made available in two different formats for the PC. A DOS version of

ARSTAN exists that is a black box where the researcher chooses the standardization procedure

which is then applied to all cores. This technique assumes that the researchers know what they

are doing in their choice of standardization and are manually plotting out the ring-width

chronologies, standardization curve fits, and the resultant index series to make sure they are

choosing the correct standardization technique for keeping the signal that they are pursuing. A

version of ARSTAN for Windows was recently developed that enables researchers to

interactively detrend their series, showing the curve fit and the resultant index series for each

core. Users can choose to fit different standardization curves to the data and see how well the

curve fits (Figure 6.23). This procedure is a good way to visualize the data and to see how

standardization affects the process.

Keystroke Tutorial for ARSTAN for Windows

To begin, the ring-width file (from a measuring program, checked with COFECHA, and edited

in EDRM) should be placed in the ARSTAN directory and the ARS37win_5f.exe file should be

executed. Enlarge the windows so that they fill the whole screen. Next, hit “enter” twice to get

past the introduction to the program. At this point the user is prompted for the name of the data

file. Following that, the user can identify a second file to include in this run or hit enter to use

only the first file. The user should then enter a descriptive title for the run that will allow for the

215
A

216
Figure 6. 23 Comparison of standardization with a negative exponential curve (A) versus a 100
year cubic smoothing spline (B) on a 600-year chronology. Notice the greater flexibility of the
100 year cubic smoothing spline for removing the slow growth when the tree is establishing
followed by the spurt of juvenile growth and then the age-related growth trend (from Clark and
Speer unpublished data).

217
identification of this run at a later date. The next option allows the user to run ARSTAN in batch

mode, enabling the user to run ARSTAN on many file sets. The default response is no.

The main menu in ARSTAN that controls the whole program appears next (Figure 6.24). There

are more than 20 options one can access at this point in the program. The options that I find

most useful are [4] first detrending, [7] interactive detrending, [15] site-tree-core mask, and [19]

summary plots. ARSTAN provides the most powerful standardization options out of any of the

dendrochronological programs. With option [4], the user can choose to fit a negative

exponential, linear trend, or various cubic smoothing splines. Option [5] allows for a second

detrending, but I am personally opposed to manipulating the data more than necessary, and a

second detrending is rarely warranted. Some researchers will use second detrending when a

deterministic model such as a negative exponential or regression line is used first to remove

noise from a known cause, i.e. age-related growth trend. Many of the standard detrending

methods, for example most cubic smoothing splines, will remove noise such as a negative

exponential curve so that two runs at detrending the series are not necessary. Also, two separate

detrending curves will move the data farther from the raw ring widths that were observed and

measured on the actual the wood. I suggest using the interactive detrending option [7] in

ARSTAN because this is the best way to visualize the data, as seen in figure 6.24. Option [15]

allows the site, tree, and core mask to be changed to fit your identification tags, but your tags

have to be consistent with the same number of characters for the site ID and tree number. The

mask fits the tree identification code so that the program can differentiate separate cores from the

same tree. In Figure 6.24 option [15] the site-tree-core mask is “sssttcc” where “sss” allows

three letters for the site ID, “tt” allows two numbers for the tree ID, and “cc” allows two letters

218
 

Figure 6. 24 Main menu for ARSTAN.

219
or numbers for the core ID. Option [19] provides summary plots so that you can visualize your

final chronologies. Option [20] is also useful for some disturbance quantification techniques.

Reading the Output of ARSTAN

When ARSTAN is run in interactive mode it plots the ring-width measurements, the curve fits,

and the resultant indices so that the user can see how well each curve fits the data (Figure 6.25).

These curves are not saved, so it is useful to do a screen capture (Ctrl + Prnt Scrn) of the plots

and then paste them into another document such as Word or PowerPoint.

The output from ARSTAN summarizes all of the descriptive statistics for the raw ring widths

and then goes through the same descriptive statistics for the standard, residual, and arstan

chronologies (explained below). These statistics include the start and end dates of each core, the

mean, standard deviation, skewness, kurtosis, mean sensitivity, and first order autocorrelation for

each core. The ARSTAN output also lists the detrending curve for each core so that any changes

that have been made in the interactive detrending part of the analysis are recorded for later

reference.

Four chronologies are produced by ARSTAN. The raw chronology is a simple average of the

raw ring widths, in other words, no standardization was done on these series. The standard

chronology is an average of the index values from the standardization process chosen by the

operator (see Chapter 2 for more details on standardization). This chronology still has all

autocorrelation included in the final chronology, which may be an issue when conducting

regression analyses later as one of the assumptions of regression analyses is that the series are

220
 

Figure 6. 25 Master chronologies for the Mokst Butte Lava Flow Ponderosa Pine Chronology.
A) raw ring width chronology B) standard chronology C) residual chronology D) arstan
chronology E) sample depth curve (from Clark and Speer unpublished data).

221
not autocorrelated. The residual chronology has had all autocorrelation stripped from the series

making it a more suitable chronology for regression analysis, but not necessarily the most

sensitive to the signal of interest. The arstan chronology has been calculated by removing the

autocorrelation, modeling it, and reintroducing a stand-level autocorrelation back into the

chronology. All three chronologies are output in the .crns file, meaning chronologies file. A

benefit of interactive mode is that these chronologies are also plotted on the screen along with a

sample depth curve for all of the chronologies (Figure 6.25). Chronology statistics such as the

running rbar and EPS value (described above) are also graphically presented to help the

researcher determine when the sample depth is so low that the that the stand-level signal is

degraded (Figure 6.2).

Regional Curve Standardization (RCS)

The Regional Curve Standardization (RCS) technique was developed as an alternative

standardization procedure that can maintain low-frequency variability in tree-ring chronologies

while removing the age-related growth trend that is unique to each site (Cook et al. 1995, Esper

et al. 2002, Esper et al. 2003). This is now a standardization option in ARSTAN. Low

frequency signal in climate reconstructions would be useful to determine long-term trends in past

climate. Short-term cubic smoothing splines remove this low frequency signal making the

reconstruction of the Medieval Warm Period and Little Ice Age impossible in thousand year-long

climate reconstructions (Cook et al. 1995). In the RCS method, the pith for each individual tree-

ring series is set to zero, regardless of the actual calendar year (Figure 6.26). It is important to

note that because this technique is based on the biological age of each ring, obtaining the pith is

especially important. Esper et al. (2003) demonstrated that this method was relatively robust for

222
differing pith offsets and sample depths, but was very sensitive to the calculation method used to

obtain the RCS. The Regional Curve represents the average growth for that stand, which can be

removed from each core by either calculating the difference from the mean growth curve (Figure

6.26) or as a ratio to the growth curve as is done in the classical method of standardization.

Esper et al. (2003) indicated that a minimum sample depth of five series for any section of the

curve is required and 40 series should be achieved at some point along the curve for the best

results. This standardization technique has been programmed into ARSTAN for Windows which

also presents the standardization curves so that its validity can be examined on a site-level basis.

YUX

YUX is a useful program that enables the user to convert a file with many chronologies to a

spreadsheet where each chronology is in a subsequent column. This spreadsheet file can then be

read into Excel as tab, comma, or space delimited, based on user specification in the program. It

is the most efficient way to convert output chronology files from ARSTAN (such as the .crns

file) or to convert raw ring width measurement files into a format that Excel can read.

Climate Analysis Packages

Two stand-alone programs have been developed to facilitate climate analysis called PRECON

(Fritts and Dean 1992) and DENDROCLIM2002 (Biondi and Waikul 2004). Similar analyses

can also be conducted in Excel or SAS but PRECON and DENDROCLIM2002 are written

specifically for dendrochronological applications, making it easier to enter tree ring and

meteorological data and incorporate more advanced principal component analyses (PCA) along

223
Figure 6. 26 Example of series lined up by a) pith date versus b) establishment time for the
Regional Curve Standardization Method (from Clark and Speer unpublished data).

224
with bootstrap techniques (Guiot 1991). Bootstrapping is a statistical technique that can be used

to determine the significance of any statistic of interest even when the data is autocorrelated, not

normally distributed, or when the data set is small. This technique creates pseudo-data sets by

randomly sampling the original data with replacement (which means that data point can be

selected again) then calculating statistical parameters for this new data set that can then be

compared to the actual data. The result is a set of confidence intervals for any regression or

correlation analysis that enables tests of significance (Guiot 1990).

PRECON

PRECON is a program written by Harold Fritts that is “an empirical model of climatic and prior

growth factors preconditioning annual ring growth in trees” (Figure 6.27; Fritts and Dean 1992).

PRECON 5.17B is the latest version, written in April 1999, and is a DOS program that allows

the user to read in a tree-ring chronology file and climatic data sets. The program runs

correlation matrices and principal components analyses (PCA) resulting in response functions for

each variable, which are then displayed in a graphical form. PRECON uses a bootstrapping

method to determine significance of the response function analysis (Guiot 1991).

DENDROCLIM2002

DENDROCLIM2002 is a C++ computer program with a Graphical User Interface (GUI) that

also conducts correlation and response function analysis, but uses a bootstrapping technique to

determine significance levels for both types of analysis, whereas PRECON uses bootstrapping

with only the response function analysis (Figure 6.28; Biondi and Waikul 2004). This program

225
 

Figure 6. 27 The opening page to PRECON. This is a program used to determine the response
function of how tree-ring chronologies respond to monthly climate data.

226
Figure 6. 28 Correlation results comparing tree rings to climate in DENDROCLIM2002 (from
Biondi and Waikul 2004).

227
also calculates the correlation and response function analyses in moving intervals through time to

determine if the climatic response is stable or if it changes through time.

OUTBREAK

The computer program OUTBREAK was written by Richard Holmes at the Laboratory of Tree

Ring Research at the University of Arizona in consultation with Tom Swetnam (Holmes and

Swetnam 1994a) for the purpose of quantifying and differentiating spruce budworm and tussock

moth outbreaks as recorded in tree rings (Figure 6.29; also see the following authors for

application of the program Swetnam et al. 1985, Swetnam and Lynch 1989, Swetnam and Lynch

1993, Swetnam et al. 1995, Speer et al. 2001, Ryerson et al. 2003). OUTBREAK allows for

data from host trees to be entered into the program along with a non-host chronology to control

for climate. The program runs on tree-level ring width indices that can be obtained from the

DOS version of the ARSTAN program. Tree-level index chronologies are developed by

averaging together the standardized “a” and “b” cores from the same tree and the results are

output as the .tre file. The tree-level chronologies give a better representation for growth in the

tree than a single core and assure that individual trees are not overrepresented in the final

chronology in the case that more cores were taken from one tree than another. The non-host

chronology is usually a master chronology from a site similar to the host sample site but of a

different tree species that is not affected by the insect or pathogen that the researcher is studying.

In 1996, this program was modified and calibrated for pandora moth outbreaks, which allowed

for host chronologies to be entered, but did not require the non-host control (Speer et al. 2001).

This modification was necessary in a ponderosa pine system where no long-lived non-host

species were available growing in the same climate conditions. In a case such as this, other

228
 

Figure 6. 29 The opening page of the program OUTBREAK.

229
efforts should be made to control for climate, such as determination of the climate response of

the host trees using modern climate.

There are two difficult assumptions made with the host/non-host comparison. First, it is assumed

that the non-host trees, usually of a different genus, have a similar climatic response to the host

trees. The climate response can be tested for each species and the reliability of this assumption

can be determined. The second assumption is that the non-host trees are not affected by the

outbreak. Non-host trees that are growing in the same stand as the host trees may change growth

due to a reduction in competition, nutrient cycling, or possibly by some damage to the trees

associated with the outbreak. If a non-host species of a different genus is lacking, it may be

tempting to use a host species for climate control that seems to be spatially separated from the

outbreak. Spatial arguments are tenuous, however, because the spatial distribution of outbreaks

is likely to have been different in the past. Therefore, a lack of modern outbreaks in a stand of

host trees does not validate that stand as a long-term climate control site.

OUTBREAK should be calibrated in an iterative process on the specific site of interest in each

new study. It was intended to quantify the effects of insect or pathogen outbreaks, and to

automate the process of identifying past outbreaks. OUTBREAK is pre-programmed with three

insect types (western spruce budworm, Douglas-fir tussock moth, or pandora moth) that have

default values for outbreak duration, severity, and onset rate. This program can be used to

quantify the growth reduction of any insect or pathogen, but it should be calibrated with known

outbreak occurrences. Characteristics of the wood should be the primary indicator that an

outbreak occurred. Once the signature of the outbreak has been identified in the wood, then the

230
program OUTBREAK can be run in an iterative process until it records the start and end dates of

historically known outbreaks. Once the program is accurately representing known outbreaks,

then it can be used to infer outbreaks in the past and to quantify the outbreak characteristics.

Four main parameters control the ability of OUTBREAK to recognize events in ring-width

measurement. These four parameters are the standard deviation of the maximum growth

reduction, the shortest length of an outbreak, the longest length of an outbreak, and amount of

the growth reduction at the beginning of the outbreak. The duration variables enable researchers

to tease apart the effect of multiple insects in the same host tree, such as spruce budworm and

tussock moth (Swetnam et al. 1995).

Spectral Analysis

Maximum entropy method (MEM; Burg 1978, Dettinger et al. 1995), singular spectrum analysis

(SSA; Vautard and Ghil 1989), and wavelet analysis (Torrence and Compo 1998) are all types of

spectral analysis that can be used to examine cyclicity in time series (Villalba et al. 1998, Speer

et al. 2001). This is a common technique in insect outbreak studies to document the return

interval of periodic outbreaks (Speer et al. 2001, Zhang and Alfaro 2002, Ryerson et al. 2003).

EVENT

The program EVENT runs a superposed epoch analysis (SEA) that overlays an event year (such

as the occurrence of a fire or insect outbreak) every time it occurs in the chronology to examine

previous and subsequent years of some variable such as climate or tree growth (Holmes and

231
Swetnam 1994b). The inputs for the program are either a time series of the comparison variable,

such as a climate variable like PDSI to see the effect of climate on the event, or a tree-ring

chronology to see the effect of the event on tree growth combined with a list of event dates. An

event window in time is identified by the user to look at a number of years prior and subsequent

to the event. Each event year is taken as year zero for that event, and then the lag years are taken

from the chronology and averaged for all of the events. Subsequently a bootstrap technique is

employed that uses a large number of random simulations (default of 1000) with randomly

selected event years to produce confidence intervals to determine if any lag year has a significant

response or correlation to the event. This analysis enables the researcher to determine if climate

is forcing fire or insect outbreaks and can also be used to examine the effect of known repeated

insect outbreak emergences (such as periodical cicadas (Magiciada sp.)) on tree growth (Figure

6.30).

Conclusion

Most of these programs and statistics are the basic tool set of dendrochronologists. There are

other specialized programs and statistics used by researchers in the various sub-disciplines but

they are more specialized and will either be described in subsequent chapters on the different

applications of dendrochronology or in the references cited in each chapter. The rest of this book

will expand on a different sub-discipline in each chapter and provide some of the specific

methods that are involved in each application along with citation of many of the main works in

that field of study.

232
 

0.20
0.15
Growth Departure

0.10
0.05
0.00
-0.05 -7 -6 -5 -4 -3 -2 -1 0 1 2 3 4 5 6
-0.10
-0.15
-0.20
Lag Years

Figure 6. 30 A superposed epoch analysis showing the growth departure in pin oak (Quercus
palustris) ring growth associated with periodical cicada emergences. Year zero is the overlay of
four emergence events back through time (every 17 years) and the analysis shows growth before
and after the emergence with the horizontal lines indicating 95% confidence intervals. This
analysis shows a decrease in growth the year after emergence, presumably from damage to these
trees from oviposition scarring. A significant increase in growth occurs five years after
emergence, which could be related to nutrient cycling from the decay of dead cicadas at the base
of the trees (Speer unpublished data).

233
Chapter 7: Dendroarchaeology

Introduction

Dendrochronology has gained recognition in archaeology as an accurate tool for chronological

control. Dendrochronologists have used tree rings to date the construction of archaeological

structures (Douglass 1929, Haury 1962, Dean 1978, Dean et al. 1985, Billamboz 1992, Cufar

2007), scars from Native American use of the inner bark of pine trees (Kaye and Swetnam 1999),

and to verify the dating of historical works of art (Lavier and Lambert 1996, Jansma et al. 2004,

Cufar 2007) such as the panels in paintings (Bauch and Eckstein 1970, Eckstein et al. 1986) or

the wood in violins (Grissino-Mayer et al. 2002). Tree rings can also be used to dendro-

provenance archaeological or historical wood (Eckstein and Wrobel 2007). This is a fast

growing sub-field of dendrochronology that uses wood anatomy and correlation to regional

master chronologies to determine the origin of and trade routes for wood that has been

incorporated into artifacts.

The first contribution of dendrochronology to archaeology was made by A.E. Douglass, who

determined the exact occupation dates of approximately 45 archaeological sites in the

southwestern United States (Figure 7.1; Douglass 1929, Haury 1962, Nash 1999). This work

started in 1914 when Clark Wissler (Figure 7.2), Curator of Anthropology with the American

Museum of Natural History suggested that Douglass use tree rings to date the Native American

structures in the American Southwest. Douglass began to examine samples that were submitted

from archaeological sites in New Mexico. In 1921, Neil Judd (Figure 7.3) of the United States

National Museum approached Douglass about continuing his dating efforts in the southwest and

234
 

Figure 7. 1 A.E. Douglass (1867-1962) coring a ponderosa pine. Douglass’ first major project
was to date some 45 archaeological ruins in the southwestern United States. Here he is coring a
ponderosa pine tree in the Forestdale Valley in Arizona in 1928 (Laboratory of Tree-Ring
Research, University of Arizona, from Webb 1983 and Nash 1999).

235
Figure 7. 2 Clark Wissler (1870-1947) and W. Sidney Stallings (1910-1989) with specimens
from Pueblo Bonito and Aztec Ruins in 1932. Clark Wissler attended a Carnegie lecture by
Douglass in 1914 on tree ring dating. Wissler realized the value of tree ring dating to obtaining
dates on archaeological wood in the southwestern US and was the first to send Douglass
archaeological samples to date. (Negative number 280306, by Clyde Fisher, Department of
Library Services, American Museum of Natural History, reprinted in Nash 1999)

236
Figure 7. 3 Neil Merton Judd (1887-1976). Judd was enthusiastic about dendrochronological
dating of dwellings and sent Douglass many samples. He also suggested that Douglass pursue
funding from the National Geographic Society for his work in developing a long chronology in
the southwest which started three major beam expeditions to complete the long master
chronology for this region. This photograph was taken at Alkali Ridge, Utah in 1908.
(Photograph from the Peabody Museum, Harvard University. Reprinted in Nash 1999).

237
suggested applying for funds from the National Geographic Society (NGS), which provided

funds for Douglass’ research from 1923-1930.

Douglass’ efforts to build a long chronology in the southwestern U.S. to date the archaeological

ruins and to build a climate chronology for himself is a classic story in dendrochronology and

also demonstrates many of the basic principles of dendrochronology. The foundation for

Douglass’ chronology came from living trees that he sampled throughout the Flagstaff and

Prescott area. Funding from the National Geographic Society for the first two “beam

expeditions” in 1923 and 1928 resulted in a 700-year modern chronology that was anchored in

time by living trees with known sampling dates and extended further back in time with

archaeological wood that had been submitted by Clark Wissler, Neil Judd, and Earl Morris

(Figure 7.4). Samples from the beam expeditions and from previous work enabled Douglass to

build a 585-year floating chronology that provided relative dates for a number of the

archaeological ruins in the southwestern U.S., but did not date against the modern chronology.

Douglass acquired funding from the National Geographic Society to conduct a third beam

expedition in 1929 to search for wood from archaeological sites that would bridge the gap

between the modern and floating chronologies. This expedition was led by Lyndon Hargrave

(Figure 7.5) and Emil Haury (Figure 7.6), with intermittent visits by Douglass himself. On June

22nd, 1929, Hargrave and Haury were leading an expedition at Whipple Ruin in Show Low,

Arizona. With the help of the Whipple family, they excavated a sample which was labeled HH-

39 (Figure 7.7). That same day, Douglass visited the ruin and spent the evening examining the

sample in the local hotel. After his analysis, he was able to announce that sample HH-39 bridged

the gap between his modern chronology and his floating chronology. In truth, there was no gap

238
 

Figure 7. 4 Earl Halstead Morris (1889-1957). Earl Halstead Morris with a charred beam at
Broken Flute Cave, Arizona in 1931 (from Nash 1999).

239
Figure 7. 5 Lyndon Lane Hargrave (1896-1978) examining a conifer cross section. Note the stone
axe cut end of the beam to the right of the cross section that Hargrave is examining. (Photograph
from the Museum of Northern Arizona and reprinted in Nash 1999).

240
Figure 7. 6 Emil W. Haury (1904-1992) examining a buried beam at Pinedale Ruin, Arizona
during the Third Beam Expedition, 1929. Photograph from the Laboratory of Tree-Ring
Research, Arizona and reprinted in Nash 1999.

241
Figure 7. 7 Sampling HH-39. This is Farmer Whipple removing sample HH-39 from an
archaeological site in Show Low Arizona. This is the famous sample that bridged the gap
between Douglass’ modern chronology and his floating chronology, allowing Douglass to
provide absolute dates to the archaeological ruins in the southwest (Courtesy of the Laboratory
of Tree-Ring Research, University of Arizona, reprinted in Nash 1999).

242
at all, but the overlap was so small that it had not been noticed. HH-39 bridged the gap with

enough rings covering the chronologies on either end that it made Douglass confident of the date

of the floating chronology (Haury 1962, Nash 1999). The work of Douglass, Haury, and the

beam expeditions resulted in the creation of a 1200-year long chronology that extended back to

A.D. 700 (Douglass 1929, Nash 1999) and revolutionized southwestern archaeology by

anchoring cultural traditions in time with great accuracy long before the advent of radiocarbon

dating and other dating techniques.

Archaeological Methods

Many of the methods used in Dendroarchaeology are similar to those employed in basic

dendrochronology, such as crossdating, sample preparation, and standardization. Other methods,

such as site selection, cannot be employed, because the site is determined by the location of the

archaeological dwelling, and the original locations of the trees are chosen by the residents of the

dwelling. Dendroarchaeology also has some unique field methods of its own.

Sample Collection

Samples are often taken from structural beams in houses or wood that is in place and has been in

position and drying for hundreds of years. To reduce the damage to the original structure and to

be able to get a sample from dry wood, a special archaeological borer is used (Figure 7.8). A

drill guide can be used to hold the drill bit steady as the researcher begins to core the beam. This

drill guide is a metal plate with a hole in the center of it, just larger than the diameter of the drill

bit. It is affixed to the beam with two short nails and is removed once the core is started. The

243
 

Figure 7. 8 An Archaeological borer. An archaeological drill can be used to cut a 12 mm core


from dry wood. This archaeological borer uses a hollow bit that cuts away the wood around the
core as the bit is drilled into the beam. A starter plate can be used to hold the bit in place while
the core is taken. A long piece of metal is inserted along the side of the core, and then twisted to
break the core off on the inside of the beam. This tool is being demonstrated by Dr. Darin
Rubino (photo provided by Darin Rubino).

244
archaeological borer is driven by an electric drill and uses a specially made extra long hole-saw

to cut the wood away from around a 10-12mm diameter core. The core is then removed from the

hole with a bent wire which is inserted down the side of the hole and twisted to break the core off

at the center of the beam. The most difficult part of this type of coring is the fact that the dust

and wood chips from drilling can clog the borer. To remove this debris from the drill hole, one

can frequently run the drill bit in and out of the hole, or core up into the beam so that the dust

falls out with gravity; although the best surface of the beam is seldom in a convenient coring

location. The dust and wood chips can also be removed by spraying a stream of air into the cut

from a can of compressed air. The coring hole is often plugged with a cork to obscure the fact

that core samples have been removed and to keep insects from making the core hole a home

(Figure 7.9). Plugging the hole in archaeological samples differs from leaving the bore hole open

in live samples because the live tree has mechanisms to defend itself, while the “dead”

archaeological sample does not. The sample ID can then be written on the cork so that any

dendroarchaeologist can refer back to a sample that was previously removed.

Archaeologists must collect the outer surface of a beam to be able to get the cutting date of a

tree. That is the most important date for a dendroarchaeologist. This outer surface can be

identified by bark, a smooth outer surface that may gain a patina with age, or by bark beetle

galleries on the outer surface of the stem. The bark beetle will feed in the cambium layer while

the tree is still alive and leave a small indentation in the xylem of the tree. Other wood boring

insects, however, leave galleries in the xylem which should not be mistaken for an indication of

the outer wood surface.

245
 

Figure 7. 9 Cores can be taken from window lintels. Although smaller than primary beams,
window lintels can also be used for dendrochronological dating of structures as long as the cores
have enough rings from bark to pith of the tree. The core holes in archaeological samples are
often plugged with cork to protect the beams from insect invasion into these spaces and also for
aesthetic reasons (note the cork plug on the right side of the front most lintel beam) (photo by
Jim Speer).

246
Cross sections can also be obtained from wooden beams and artifacts. This process is more

destructive, but it provides a larger amount of wood for analysis and when searching for micro or

locally absent rings. Cross sections also give the researcher a greater chance to find more rings

towards the outside of the tree, and thus get closer to a cutting date. One test for a cutting date

on a tree sample is the continuity of the outer ring around the circumference of the section, so a

full section can provide this data where a core cannot. Collecting a cross section from the end of

a beam in a door frame or window frame can also reduce the visibility of sampling. The amount

of wood available and the integrity of the artifact may constrain how much wood one is allowed

to sample.

In the southwestern U.S., the Pueblo cultures used large primary beams and smaller crossing

secondary beams to support multiple stories in their structures (Figures 7.10 and 7.11). These

structures have provided extensive samples for the development of local chronologies from

which cutting dates have been determined. Archaeological samples such as cross sections and

cores taken from these beams can be surfaced with sandpaper using the methods discussed in

Chapter 5.

Charcoal can also be used to date archaeological structures (Figure 7.12). Once the wood is

carbonized, it is relatively inert and can last on a site for hundreds or even thousands of years

without any biological decay. The cell structure is preserved in carbon, but it is extremely

fragile, and may be mechanically broken down over time. The surface of carbonized wood

cannot be prepared in the same way as other dendrochronological samples; instead of sanding,

the wood must be snapped to produce a freshly broken surface along the cross sectional view or

247
 

Figure 7. 10 Primary and secondary beams. Native Americans in the southwestern United States
constructed complex above ground dwellings and incorporated large beams to support multiple
stories. The major beams making up the support for the floor are called primary beams and the
smaller poles making up the floor are called secondary beams (photo by Jim Speer).

248
Figure 7. 11 Cross section of a primary beam. Crosssections from the main support beams can
be sampled to collect a complete series of years. Cutting dates can be obtained by cross dating
the ring series of these dead trees against a master chronology of the area to assign exact felling
dates, demonstrating when the trees were cut for incorporation into the architecture (photo by
Jim Speer).

249
Figure 7. 12 Charcoal samples can also be used for archaeological dating. Because the charcoal
is inert, it is well preserved in the soil so that the wood does not decay through time. Charcoal is
very fragile and care has to be taken when collecting these samples. Unlike green wood or old
beams, charcoal samples should not be sanded. These samples can be broken to expose a clean
surface of the cross sectional view (from Stokes and Smiley 1968).

250
can be cut with a very sharp blade. A freshly broken surface is perfect for dating because all of

the wood structure is visible in reflected light. Snapping carbonized wood is a finicky procedure

and takes some practice. The dendrochronologist is also effectively breaking an archaeological

artifact or ecofact (a natural object found at an archaeological site) in this type of sample

preparation, and therefore should proceed conservatively.

Dean (1978) has developed a key to inside and outside dates that are identified in archaeological

samples (Table 7.1). This key can really be used on any tree-ring sample, and helps researchers

determine the quality of the dates. For identification of the inside date of the sample a “p” is

used for a pith date, while “fp” is used to designate a date far from the pith, and “+/- p” is used to

indicate that the pith is present, but because of poor ring condition on the inside of the sample, an

exact pith date cannot be determined. Likewise for the outside of the sample, a “B” designates

that bark is present, “G” means beetle galleries are present, and “L” means that the sample has a

smooth surface and patina suggesting that the outside is the true cutting date of the sample. A

“c” means that the outer ring is present all the way around the circumference of the sample. This

usually only occurs when the outside date is the true cutting date. Erosion of the outer surface

will usually cut across ring boundaries, differentially removing the outer surface of the sample.

With the degradation of the outer surface of the sample, various symbols such as “r”, “v”, “vv”,

“+”, and “++” indicate a lessening confidence that the outside date represents a cutting date for

that sample. This nomenclature can be used for any application of dendrochronology where the

inside or outside date of the sample is important.

251
Table 7.1 Symbols used to mark archaeological samples. Dendroarchaeologists working on wood use a series of
codes to demonstrate the quality of the outside dates. The presence of beetle galleries, patina, the smoothness of the
outer surface, and presence of a complete ring around the circumference of the section can all indicate whether an
accurate death date can be determined for the tree (From Nash 1997 and 1999).
Table --- Symbols used to mark archaeological samples.
Symbols used with the inside
date
Year No pith ring is present.
p Pith ring is present.
fp The curvature of the inside ring indicates that it is far from
the pith.
±p Pith ring is present, but because of the difficult nature of
the ring series near the center of the specimen, an exact
date cannot be assigned to it. The date is obtained by
counting back from the earliest date ring.
Symbols used with the outside date
B Bark is present.
G Beetle galleries are present on the surface of the specimen.
L A characteristic surface patination and smoothness, which
develops on beams stripped of bark, is present.
c The outermost ring is continuous around the full
circumference of the specimen.
r Less than a full section is present, but the outermost ring is
continuous around the available circumference.
v A subjective assessment that, although there is no direct
evidence of the true outside of the specimen, the date is
within a very few years of being a cutting date.
vv There is no way of estimating how far the last ring is from
the true outside.
+ One or more rings may be missing from the end of the ring
series, whose presence or absence cannot be determined
because the specimen does not extend far enough to
provide an adequate check.
++ A ring count is necessary because, beyond a certain point,
the specimen could not be dated.
Note: The symbols B, G, L, c, and r indicate cutting dates in order
of decreasing confidence. The + and ++ symbols are
mutually exclusive but may be used in combination with all
other symbols.

252
Chronologies Used in Dendroarchaeology

Long-term tree-ring series have been constructed that have application for archaeological regions

worldwide (Cufar 2007). Excellent preservation of wooden beams in arid environments has

made dendrochronology prominent in archaeology in the American southwest. This preservation

has enabled the creation of pine (Pinus sp.), fir (Abies sp.), spruce (Picea sp.), Douglas-fir

(Psuedotsuga menziesii), and juniper (Juniperus sp.) chronologies that extend back 2,300 years

(Kuniholm 2001). The bristlecone pine chronology (Pinus longaeva) is over 8,700 years long

(Ferguson et al. 1985) and provides important dates for volcanic events (LaMarche and

Hirschboeck 1984) and climate reconstruction (LaMarche 1974) and has been used in the

calibration of the radiocarbon curve (Becker 1991, Friedrich et al. 2004). Long oak chronologies

extending back 11,000 years have been developed in Ireland and Germany, providing the master

chronologies needed for obtaining construction dates on structures throughout the region (Pilcher

et al. 1984, Becker 1993, Baillie 1995, Jansma 1996, Cufar 2007). The eastern Mediterranean

chronology has wood that goes back 9,000 years before the present, but has a number of gaps left

to be filled (Kuniholm 2003). These chronologies have been developed from modern specimens,

samples taken from consecutively deeper layers of oaks in bogs, and from archaeological

structures themselves. Many other long-term chronologies have been developed around the

world that could possibly be used as master chronologies for archaeological dating (Table 7.2).

Applications of Dendrochronology to Archaeology

Dean (1997) categorizes the use of dendrochronological evidence in archaeology under three

separate applications: chronological control, behavioral information, and environmental

information. Chronological control has been the standard use of dendrochronology in

253
Table 7. 2 Long-term chronologies from around the world. The length indicates how far back in
time from the present these chronologies extend. There are gaps in some of these chronologies
and some of the archaeological chronologies do not extend to the present.

Species Location Inside Length Reference


Date
Combined Pinus and Germany 10,461 BC 12,460 Friedrich et al. 2004
Quercus
Pinus sp. Germany 9,494 BC 11,370 Becker 1993
Quercus petraea Germany 8,021 BC 10,076 Becker 1993
Quercus robur
Juniperus sp. Eastern Mediterranean 7,020 BC 9,000 Kuniholm 2003
Pinus longaeva White Mountains, North 6,716 BC 8,700 Ferguson et al. 1985
America
Quercus petraea Ireland 5,218 BC 7,272 Pilcher et al. 1984
Quercus robur
Various species Swiss Chronology 4,086 BC 6,086 Egger et al. 1985
Quercus petraea France 3,659 BC 5,659 Girardclos et al. 1996
Quercus robur Lambert et al. 1996
Quercus sp. Netherlands – Floating 2,258 BC 1,100 Jansma 1996
Chronology
Fitzroya cuppressoides Chile 1,634 BC 3,622 Lara and Villalba 1993
Sabina przewalskii China 1,580 BC 3,585 Shao et al. 2007
Sequoiadendron gigantium California, USA 1,229 BC 3,220 Brown et al. 1992
Pinus aristata Arizona, USA 662 BC 2,262 Salzer 2000
Quercus petraea Poland 474 BC 2,474 Krapiec 1996
Quercus robur
Pinus sp. Southwestern United States 322 BC 2,327 Dean 1997
Pseudotsuga sp.
Taxodium distichum Southeastern USA AD 372 1,600 Stahle et al. 1988
Pinus sylvestris Fennoscandia AD 443 1,555 Briffa et al. 1990
Lagarostrobos franklinii Tasmania AD 900 1,210 Cook et al. 1992
Larix sibirica Polar Urals AD 961 1,008 Graybill and Shiyatov
1992

254
archaeology, although in Bannister’s (1963) article summarizing the state of dendroarchaeology,

the useful chronologies were confined to the Southwest and Great Plains in the U.S., Western

Europe, and Russia. Since that time, the application of dendrochronology has become more

popular around the globe, making master chronologies available in many more regions than

previously realized. Dendrochronological dates are useful in archaeology because they are

accurate to the year without any error. Because of this accuracy, tree rings have been used to

calibrate the radiocarbon curve and have, on occasion, upset previously determined

archaeological chronologies (Baillie 1995). Dean (1997) notes that tree-ring dating of artifacts

provides a suite of behavioral information “…including treatment of trees as natural resources,

use of wood as raw material, seasonal timing of tree felling, sources of wood, tools and

techniques of tree felling and wood modification, differential use of species, use of dead wood,

reuse of timbers salvaged from older structures, stockpiling, structure remodeling and repair…”.

Also, the expansion of new applications of dendrochronology has been producing environmental

records of climate and possible resource availability that can be used in archaeological

interpretation (Speer and Hansen-Speer 2007). Dendroecological records are becoming more

available around the world and are not dependent upon preservation conditions in archaeological

sites. The prevalence of dendroecological studies can extend the benefit of dendrochronology to

archaeologists that do not have preserved wood on their particular sites.

Construction dates

Construction dates are the most common application of dendrochronology to archaeology

(Bannister and Robinson 1975, Billamboz 1992). This is the information that Douglass (1929)

provided for the archaeological ruins throughout the southwestern U.S. Additional work has

255
been done since that time, providing initial construction dates as well as expansion and repair

dates, enabling the archaeologist to interpret human behavior and habitation periods.

Archaeological dates from the southeastern U.S. are becoming more common as cabins and other

historical structures from the settlement of North America are dated (Figure 7.13 and 7.14).

Stahle (1979) successfully dated 24 cabins from Arkansas that had cutting dates that ranged from

1825 to 1911. This work helped to extend the living chronologies for yellow pine (Pinus sp.),

eastern red cedar (Juniperus virginiana), white oak (Quercus sp.), and baldcypress (Taxodium

distichum) further back in time, providing a resource for future dating attempts. Bonzani et al.

(1991) were able to use wooden planks from a lock system on the Main Line Canal in Pittsburg,

Pennsylvania to extend a white pine (Pinus sp.) chronology back to A.D. 1658.

Dendrochronology provides the ability to verify or reject previously held beliefs for construction

dates of historical homes. Bortolot et al. (2001) dated a cabin that was thought to have been

constructed in 1814, but was actually constructed in 1876. It has often been the case that these

homes have been built later than previously thought.

The wood in historical structures throughout Europe is an important resource that has been

extensively used to obtain dates of construction and to develop long chronologies (Eckstein

1972, Becker and Delmore 1978, Becker 1979, Baillie 1982, Laxton and Litton 1988, Billamboz

1992). Extensive archaeological collections now enable broad scale analysis of towns in Europe

and allow researchers to compare construction dates to earliest historical documentation of the

towns. Westphal (2003) used 5,002 beam samples from 87 towns that were constructed between

A.D. 800 and A.D. 1300 between the Elbe and Lower Oder rivers in Germany. He founds that

256
Figure 7. 13 A log cabin from the southern Appalachian Mountains. Log cabins from the
eastern United States are a great resource for old chronologies and a dendroarchaeologist can
provide construction dates for these dwellings (photo by Jim. Speer).

257
Figure 7. 14 Cross section of a beam from a log cabin. Many pioneer log cabins have very old
wood incorporated in their structures. Some of these beams can prove problematic in
determining accurate outside dates because the outer rings were often removed as the timbers
were shaped for construction. Care must be taken to sample through an area that has complete
outer rings as observed in the cross section (photo by Jim. Speer).

258
on average the towns were constructed 40 years prior to any written comment of it being a place

and 50-60 years prior to it being called a town. In some cases, 250 years passed before there is

any written mention of the town. Eckstein’s (1972) summary article noted at that time

dendrochronology laboratories were established in most European countries or analysis could be

done in neighboring countries. He also noted that most of this work was conducted on historical

structures which provided a large amount of wood and extended their chronologies back in time.

Work in Finland has dated wooden causeways that both provide behavioral information on past

cultures and a large quantity of wood for extending our chronologies (Zetterberg 1990).

Dating Artifacts

Lavier and Lambert (1996) report on research conducted at the Laboratoire de Chrono-Ecologie

in France where they frequently date wood from paintings, furniture, sculptures, and covers of

books. They take this work further and examine where the wood came from, how the artwork

was made, how wood was chosen, and the time between felling trees for the artwork and when

the work was completed (Lavier and Lambert 1996). All of this work demonstrates some of the

unique contributions that dendrochronology and wood anatomy can make to archaeological

research, specifically dealing with the behavioral information to which Dean (1997) referred.

Wooden panels were used in the Netherlands and England as the medium for paintings of the

14th through 16th centuries (Fletcher 1976, 1977, Eckstein et al. 1986). These panels can be

dated to determine when the paintings were actually completed and to verify their authenticity.

Also, if the date of the painting is known from historical records, dendrochronological dates can

be used to determine behavioral aspects of how the wooden panels were processed. Exact dating

259
on wooden panels is hampered by the practice of cutting away the outer surface of the wood,

possibly in an attempt to remove damage by wood boring insects (Baillie 1982).

Dating musical instruments is also possible in dendrochronology if there are enough rings in the

instrument and the proper master chronology can be found for comparison (Figure 7.15). It is

often the case that wood from exotic locations can be used in the construction of artifacts. This

foreign wood makes finding the proper master chronology a challenge in dating art and artifacts.

Besides dating the Messiah violin of Stradivari (Grissino-Mayer et al. 2004) as related in chapter

1, Grissino-Mayer et al. (2005) dated the Karr-Koussevitzky double bass (Figure 7.16). With

this analysis they found that the instrument had 317 rings on its face plate (the most rings ever

recovered from a musical instrument) but the last rings grew in 1761 demonstrating that the

instrument was not made in 1611 by the Amati Brothers as was originally thought.

Climate Reconstructions

Archaeologists have made good use of dendroclimatic reconstructions of temperature and

precipitation to explain environmental resource limitations and subsequent migration patterns

(Dean et al. 1985, Grissino-Mayer 1995, Ahlstrom et al. 1995, Kaye and Swetnam 1999, Van

West and Dean 2000) and have used streamflow reconstructions to provide paleoenvironmental

information for an area (Nials et al. 1989). Stahle et al. (1998a) reconstructed the last 800 years

of climate variability from bald cypress in the southeastern U.S. This chronology provided the

background information of the climate during the establishment of the Roanoke and Jamestown

colonies along the east cost of the U.S. The Roanoke colony was established during the most

extreme drought recorded in the 800-year chronology and the Jamestown colony was established

260
 

Figure 7. 15 Rings on the face of a cello. Tree rings are clearly evident on the face plates of
musical instruments and can be crossdated against master chronologies from the region where
the wood for the instrument was harvested (photo from Topham and McCormick 1997).

261
Figure 7. 16 The Karr-Koussevitzky double bass marked up for measurement (from Grissino-
Mayer et al. 2005).

262
during one of the driest 7-year stretches during that same time period. Such climatic data can

help archaeologists interpret the archaeological record in the context of the past climate of the

area.

Ecological Reconstructions and Anthropogenic Ecology

Dendroecology is a recent branch in the field of dendrochronology (starting in the 1970s) that

uses tree rings to reconstruct environmental records other than climate (Fritts 1971, Fritts and

Swetnam 1989). It is a field that has not been used to its full capacity in archaeological research.

Dendroecology can be used to develop records of fire history (Swetnam and Baisan 1996), insect

outbreaks (Swetnam et al. 1985, Speer et al. 2001), and acorn production (Speer 2001).

Researchers are developing long records of these variables that can be useful to archaeologists

interested in anthropogenic ecology and resource availability (Speer and Hansen-Speer 2007).

Billamboz (1992) used the cutting dates for timbers in two lake dwellings in southwest Germany

and found some distinct periods of forest clearance in 1767-1730 B.C. and 1511-1480 B.C.

These clearance events were associated with settlement phases and were documented by the

gradual shift to smaller timbers and a change in the tree species that were used for structural

timbers over time. This use of archaeological timbers to understand silvicultural practices of

past cultures has been termed dendrotypology and demonstrates another set of information that

can be obtained from archaeological wood (Billamboz 1992, Billamboz 2003).

Fire in the southwestern United States. Native American use of fire is an issue that has been

debated for the last half century (Pyne 1982, Swetnam 1990, Agee 1993, Vale 2002, Wagner

2003). Native Americans may have used fire to aid in hunting, to improve grasslands, and in

263
warfare (Stewart 1936, Shinn 1978, Pyne 1982). Lightning ignition of fires is also very common

in the southwestern United States and produces a natural background of fire occurrence

(Swetnam and Baisan 1996). Swetnam and Baisan (1996) argue that ignition sources are not the

limiting factor, but that the appropriate fuel and climatic conditions control the occurrence of

fire.

Work by Wilkinson (1997) has shown some effect from Native American burning as

demonstrated by an increase in fire occurrence during times of Spanish pressure on Native

American encampments. In the Sacramento Mountains of New Mexico, she found that broad

scale disturbance from anthropogenic sources did not occur until the introduction of grazing in

the 1880s and fire suppression in the early 1900s. This local effect on the fire regime was

identified by comparing fire occurrence over a broad area and in a specific forest type to

individual sites histories. Such an approach makes the broad scale pattern the norm to which

irregular fire histories can be compared and described.

Fire in the eastern United States. In the southeastern United States, the fire issue is not so clear.

Many people believe that fire is a natural part of the oak woodlands (Abrams 1985, 1992, 2000).

Recent work, however, argues that much of past fire occurrence is from the direct effect of

Native American and Euro-American burning (Jenkins et al. 1997, Sutherland 1997, Guyette et

al. 2002). In the southeastern United States, few fire histories extend much before 1800. Most

of fire history chronologies in the eastern U.S. are from oak trees but the full suite of hardwood

trees have not been examined for fire history. More regional work, use of other hardwood tree

264
species, and a longer time perspective may help to answer questions of Native American burning

in the southeastern United States.

Culturally Modified Trees. Culturally modified trees (CMT) provide direct evidence of Native

American use of trees (Figure 7.17) (Swetnam 1984, Mobley and Eldridge 1992, Wilkinson

1997, Towner et al. 1999, Lewis 2002). These trees provide the year and season of Native

American occupation and can be related to social forcing factors of the time. CMTs are found

throughout the ponderosa pine zone from Mexico into Canada. Native Americans were thought

to peel the bark from these trees in the spring time and eat the inner cambium as a starvation

food. Peeled trees generally occur in clusters of about 20 individuals (personal observation) and

the scar left on the trees can easily be dated to the year of damage and sometimes the season.

Mobley and Eldridge (1992) conducted a systematic examination of CMTs reporting on 967

peeled trees in the Pacific Northwest region with the oldest scarred tree dating back to A.D. 1467

(error approximately +/- 10 years, based on a ring count). While this work demonstrates the use

of tree rings to determine the use of culturally modified trees, it would be much improved if

crossdating was used so that the exact year of Native American activity could be determined.

Slash pine (Pinus elliotii) and longleaf pine (Pinus palustris) have been modified in the

southeastern U.S. by Euro-Americans since the mid-1700s for the production of turpentine as

part of the naval stores industry (Grissino-Mayer et al. 2001). Workers would cut through the

bark and into the wood of these pines, a process called chipping and collect the sap that came

from these wounds. Grissino-Mayer et al. (2001) found a concentration chipping events in two

southern Georgia sites in 1925, 1947-48, and 1954-56. These studies show that any preserved

evidence of tree modification can be used to interpret the timing of past human behavior.

265
Figure 7. 17 Peel bark tree. Peel bark trees or culturally modified trees can provide a date that
Native Americans were active on a site. These trees have axe marks in the wood at about knee
height and again above head height. The bark was then peeled from the tree and most likely
used as a starvation food source. In Canada these are considered artifacts and are protected by
law whether they are living or dead trees in the forest (photo by Jim. Speer).

266
Culturally modified trees are now protected as archaeological artifacts in Canada and many

locations in the United States are also starting to protect these trees.

Insect Outbreaks. Speer et al. (2001) developed a record of pandora moth outbreaks that extends

back 622 years in south-central Oregon (see Figure 9.10). Pandora moth is a phytophagous

insect that defoliates ponderosa pine, Jeffrey pine, and lodgepole pine in the western United

States. The Klamath and Piute Indians used the pandora moth larvae and pupae as a traditional

food source when it was available, indicating they had knowledge of its life cycle (Blake and

Wagner 1987). This led early forest entomologists to speculate that pandora moth outbreaks had

often recurred in the past (Aldrich 1912, 1921, Patterson 1929). These types of reconstructions

can be used to demonstrate resource availability for native peoples.

Mast. Recent work in dendroecology has produced a new technique for developing mast

(massive fruit production in trees; specifically acorns in this example) reconstructions from tree

rings (Speer 2001) (Figure 7.18). Native American groups have been present in the southern

Appalachians for at least the past 12,000 years (Yarnell 1998) and have been using nuts as a food

source throughout much of their history in North America. One use of a mast reconstruction

would be to determine the dependability of mast as a human food source in prehistoric times and

as livestock feed in historic times.

Dendrogeomorphology in Archaeology. Geological applications of dendrochronology can also

be used to inform archaeological interpretation. One of the better examples of this is the

267
Figure 7. 18 White oak (Quercus alba) regional mast reconstruction from the southern
Appalachian Region. The reconstruction is based on 165 white oak trees from Tennessee, North
Carolina, and northern Georgia. Mast years are shown as z-scores with numbers larger than 1.2
and less than -1.2 considered extremely good or poor mast years respectively (from Speer 2001).

268
reconstruction of the eruption of Sunset Crater in A.D. 1064 in Northern Arizona by dating a

growth reduction due to projectile damage from the eruption (Smiley 1958). On a broader scale,

short-term global or regional temperature changes have been identified that were caused by

major volcanic eruptions (LaMarche and Hirschboek 1984, Baillie 1995). Tree rings can also be

used to determine bounding dates on land surfaces and some archaeological earthworks, helping

to determine the chronology of archaeological sites.

Future of Dendroarchaeology

Dendroarchaeology will continue to find new applications for the chronological control that

dendrochronology provides. The use of dendrochronology to determine construction dates has

long been used to great benefit in archaeology. I recommend that archaeologists look more to

climatic and ecological reconstructions from the various subfields of dendrochronology to

develop a richer dataset for the interpretation of the archaeological record.

269
Chapter 8: Dendroclimatology

Introduction

One of the first and most publicly debated applications in dendrochronology has been the ability

to reconstruct climate from tree rings. Because trees respond to their surroundings, they are

subject to climatic stresses such as variations in temperature, rainfall, soil moisture, cloudiness

days (number of days with clouds which reduces photosynthesis), and wind stress. In fact,

climate seems to be one of the main controlling factors of most tree-ring growth across all spatial

and temporal scales. The basic steps in a climate reconstruction are relatively simple and are

often normal procedures that are done even before ecological reconstructions. But the statistical

analyses of tree-ring chronologies for dendroclimatic reconstructions have become increasingly

sophisticated.

Dendroclimatologists are interested in past climate so that the variation and trend of modern

climate can be put into perspective. The natural range of variation of the climate system can be

reconstructed from examination of the past through tree rings (Morgan et al. 1994). From

various types of climate reconstructions (based on ice cores, marine and lake sediments, and

dendrochronology) we have learned about the glacial/interglacial cycle (100,000 years), the

shorter-term Holocene climate variation (past 10,000 years), and documented recent warming in

the modern era (Figure 8.1). Mann et al. (1998) reconstructed climate variation from multiple

proxies including tree rings for the past six centuries showing an abrupt increase in temperature

associated with the industrial revolution (Figure 8.1). This reconstruction has been questioned

from many quarters with the most constructive criticism stating that it does not take low

270
 

Figure 8. 1 Tree-ring climate reconstruction for the past 1,000 years. The Mann, Bradley, and
Hughes (1998) “Hockey Stick” graph, which was used by the Intergovernmental Panel on
Climate Change (IPCC), is built from long tree ring chronologies throughout the terrestrial land
surface. The solid black line is a running average to smooth the data. There are many more
recent papers that discuss improvements for this curve, because while tree rings are excellent at
capturing short frequency variability, they are not very good at capturing long-term variability.

271
frequency climate variability into consideration as shown by lack of evidence for the Medieval

Warm Period and the Little Ice Age (Moberg et al. 2005).

Climate phenomena, such as hurricanes, can be reconstructed from tree rings because of the

specific signal recorded in ring width and in the isotopic chemistry of the rings. Climatic

reconstruction, therefore, can be used to examine the proximal cause of ring width, such as

changes in temperature or rainfall, or it can be used to examine broader scale patterns and

phenomena that are recorded along with changes in temperature and rainfall. In the case of

hurricanes, an isotopic signature can be identified in the fluctuations of wood chemistry through

time (Mora et al. 2006). Another powerful tool is use of tree-ring networks to examine climate

variability on a broad spatial scale such that inferences can be drawn about long-term changes in

synoptic climatology (the flow in the climate system including pressure differences)

(Hirschboeck et al. 1996).

Tree growth is one example of a proxy, or a natural phenomenon that indirectly records an event

of interest, such as a hurricane or flood. Other examples of proxy records that record climate are

coral growth, ice deposition, sediment deposition, or cave dripstone. By studying the dynamics

of a region or watershed with multiple proxies, we can better understand the vegetation response

to changes in climate. For example, tree-ring reconstructions of climate can be examined

alongside pollen reconstructions of vegetation change to see how ecological systems respond and

interact with climate change (Friedrich et al. 2001). Given fine enough resolution in the proxy

records, a long-term record of climate and vegetation change can inform us about the

272
mechanisms involved in vegetation change and show us possible feedback loops through

microclimatic effects.

Methods for Dendroclimatology

Climate reconstruction starts with a site-level analysis of a tree species’ climate response.

Standard dendrochronological methods are used such as site selection, coring at breast height,

crossdating, and measuring the samples (see chapters 2 and 5 for a full description). Trees are

chosen from climate sensitive sites (e.g., Figure 8.2), such as steep rocky slopes or northern

treeline. A variety of tree ages can be used in climate reconstructions because trees may change

in their climatic response with age. The oldest trees are chosen to obtain the longest

chronologies, but older trees may have a weakened climate signal due to senescence. Trees

with obvious injuries or sub-dominant canopy position are avoided because of possible

complication of the climate signal with other micro-environmental factors. The climate signal in

growth of very young trees may similarly be distorted by such factors.

Sample depth is always an important issue in dendrochronology, but of paramount importance

in climate reconstruction. Sample depth is simply the number of samples that represent a

phenomenon back through time. The ring-width measurements are corrected for an age-related

growth trend (see standardization in chapter 2) and the resultant index values are averaged

together to create a chronology with a stand-level signal that is analyzed for its climate response.

The goal is to create a robust climate reconstruction that maintains a consistent climate signal

whether sample depth is increased or the ring width series are standardized in a different fashion.

Because we often use living trees for our climate reconstructions, we can sample at least 30

273
 

Figure 8. 2 Old preserved wood on a lava flow in Oregon. Trees struggling to grow in harsh
conditions, such as this lava flow in Oregon, tend to be the oldest individuals in the species and
also can be good recorders of past climate (especially precipitation records). Preservation on
these sites is also very good because of the lack of soil microbes, leading to the potential for very
long chronologies (photo by Tom Swetnam).

274
living trees (two cores per tree) for a robust sample depth in the modern era. Not all of those

trees, however, established at the same time, so they fall out of the record at different times as it

goes back into the past. When the chronology falls below 10 or 20 trees, growth variations of

individual trees may overwhelm the common growth signal that conceptually represents the

response to climate; a statistical side effect can be the increase in variance at the beginning of the

chronology.

An important part of the standardization procedure in the processing of tree-ring measurements

for dendroclimatology is the removal of non-climatic trends from the ring widths. A sensible

strategy in standardization is to retain as much variance as possible at the long wavelengths or

low frequencies. Such variance represents the gradual fluctuations in growth over periods of

decades and longer. A negative exponential or linear trend line can be used to remove the age-

related growth trend from the time series. These line fits should be examined on an individual

core basis to make sure that the curve accurately represents the ring-width series. Chronologies

can also be truncated after the irregular juvenile growth and the steep curve from an age-related

growth trend have leveled out. Such truncation will result in removing many years from the

beginning of each tree, but can sometimes simplify the standardization curve that is required to

process a site. Cubic smoothing splines are often used in many dendrochronological

applications because they produce a flexible curve that fit the data fairly closely. Splines,

however, should be used with caution and with full knowledge about the variance that is being

removed. LaMarche (1974) argued that the use of raw ring widths can sometimes provide a

more accurate climate reconstruction because real climatic trends in the data may be removed

during the standardization process. He was working in a unique circumstance with bristlecone

275
pine (Pinus longaeva and Pinus aristata) trees from above treeline. His chronologies were

longer than 1,000 years in length so that he could omit the juvenile growth from the trees and

still examine long-term climate changes.

The first step in a climate reconstruction (and most dendrochronological studies) is to examine

the climate response of the chronology which can be accomplished with a simple correlation

matrix or a response function analysis. Climate variables such as monthly temperature,

precipitation, and Palmer Drought Severity Index (PDSI) should be gathered for each research

area and entered into a spreadsheet. Individual climate station data can be obtained from the

Historical Climate Network (HCN) and climate division data can be obtained from the National

Climate Data Center (NCDC) for the entire United States back to 1895. Another recently

developed data set is the PRISM data set (see Appendix E for web address) which was developed

at Oregon State University and takes individual climate station data and models the signal over

the landscape based on a physiological model. This data set provides accurate climatological

information for locations that have not been previously monitored. Other historical data can also

be used to calibrate dendroclimatic reconstructions such as grape harvest data in France (Guiot et

al. 2005).

Climate division data is often preferred to individual site data for many reasons. First, it is often

hard to find an individual climate station that is near to your sampling site and similar in

microenvironmental condition. Second, most individual stations have some data gaps in their

records, while the climate division data draws on many climate stations and has been corrected

for changes in the location of individual stations and for these data gaps. A correlation matrix

276
can be created in almost any statistical package including Excel, to compare the master

chronologies to every month of climate data. This analysis will reveal which months of climate

data are correlated with ring width. The significant months can then be aggregated into seasons

that are appropriate for each tree species being examined. Such a posteriori determination of the

effective climate window for a particular site and species enables researchers to document the

climate response of the chronology rather than forcing a hypothetical climate window onto a

species for a given location (Fritts 1976: 34)

Simple linear regression can be used to develop a model for the reconstruction of one climate

variable (Duvick and Blasing 1981). With the 100 years or more of climate data from the local

climate division, a model can be built from half of the data (the calibration set) and tested against

the other half of the data (the verification procedure) (Fritts 1976). If the calibration and

verification procedures meet the statistical requirements of the study, then a significant climate

reconstruction can be completed for the length of the chronology by repeating the calibration on

the full length of overlap of climate and tree-ring data and substituting the long-term tree-ring

data into the regression equation. Sometimes the model is extended to multiple linear regression

by including more than one tree-ring series as predictor variables, or by including tree-ring series

that are lagged relative to the climate series (Meko et al. 1980). More sophisticated models can

also be used to examine the response of trees to climate. Graumlich (1993) used response

surfaces, Woodhouse (1999) used neural networks, and Meko and Baisan (2001) used binary

classification trees to find the most accurate way to identify the climate signal in tree-ring

chronologies.

277
Some research in climate reconstruction has used Principal Component Analysis (PCA) to

reduce the number of climate variables, produce orthogonal factors (completely independent

variables), and to create variables that represent similar climate measures (LaMarche and Fritts

1971, Mann et al. 1998). PCA plots a cloud of climate variables in as many dimensions as there

are variables, and then creates a best fit line through the data that represents most of the variance

of all of the climate factors. This first line is called the first eigenvector. Then a second

eigenvector is fit to the data so that it is orthogonal (or completely perpendicular and

independent) to the first eigenvector. This process continues until all of the variance in the

original data set is captured. PCA generally produces two to five eigenvectors which represent a

large part of the variance in the original data set (often some 70 variables). In climatic

reconstruction, PCA can be used to reduce the climate variables, the tree-ring variables, or both.

Regression analysis can then be used to model the relationship between the eigenvectors of

transformed climate data and transformed tree-ring variables. The process of combining PCA

and regression analysis usually explains more variance in the tree-ring chronology than simple

linear regression could with untransformed climate variables.

Spatial regression methods using PCA in dendroclimatology were reviewed by Cook et al.

(1994). A novel approach using PCA on the tree-ring chronologies in a moving spatial window

was developed by Cook et al. (1999) to combine a dense network of tree-ring sites and create a

regular grid of PDSI reconstructions throughout the United States. This use of PCA, which will

be discussed more in the next section, reduces chronologies to modes of tree-ring variation for

geographic regions centered on gridpoints. PCA is a powerful tool for reducing a complex set

278
of variables to the common signal and is likely to have more application in dendroclimatology in

the future.

Applications of Dendroclimatology

Tree response to climate from many sites in a region or across a continent can be used to map the

climate variables that affect tree growth in different regions (Fritts 1976: 35). With a network of

chronologies, spatial patterns of effective climate can be determined. Furthermore, with the

time depth provided by dendrochronology, the spatial patterns can be studied to determine the

size and distribution of climate events such as droughts through time. Brubaker (1980)

conducted one of the earlier dendrochronological network analyses to examine climate response

across the Pacific Northwest using PCA. She found that the first eigenvector responded to

Spring-Summer rainfall and the second eigenvector responded to summer temperature and

winter rainfall. This signal remained constant over the 400-year chronology that she developed.

LaMarche and Fritts (1971) started collecting a climate network of tree-ring chronologies to

reconstruct broad scale drought throughout the U.S. and Canada. A reconstruction of PDSI from

the network established a long-term context for the dust bowl drought of the 1930s in western

North America (Stockton and Meko 1975) and revealed a bi-decadal drought rhythm with a

weak statistical link to the Hale Sunspot Cycle (Mitchell et al. 1979). Expansion of the

geographical coverage of tree-ring collections, especially in the eastern USA, resulted in a

climatically screened network of chronologies whose statistical properties and drought signal

were analyzed by Meko et al. (1993), This initial network was further developed by Cook et al.

(1999), who examined tree response to PDSI in order to map drought reconstructions for the

continental U.S. on a 2° latitude by 3° longitude grid for 1700 to 1978. Their landmark paper

279
provided long-term drought reconstructions for the entire U.S. that were then used to examine

the intensity of the Dust Bowl drought and other climatic phenomena in comparison to this long

climate window. Stahle et al. (2000) used a similar network with a broader spatial coverage and

expanded temporal depth to examine drought across western North America, including Mexico,

and found a mega-drought in the 1600s that was a much more extreme event than the 1930s Dust

Bowl.

Broad-scale tree-ring networks, such as those used in the PDSI reconstructions, can be more

generally applied in synoptic climatology, the study of climate from the perspective of

atmospheric circulation. Circulation patterns can be inferred from reconstructed patterns of

precipitation, temperature and pressure (e.g., Fritts et al. 1971, Blasing and Fritts 1976, Fritts

1976, Fritts and Shao 1992, Hirschboeck et al. 1996, Barber et al. 2004, Girardin et al. 2006).

The compiling of Hemispheric climate reconstructions, begun in the 1970s, brought to light

standardization issues that arise when different species and age chronologies are compared

across different regions. Researchers from around the globe standardize their tree-ring

chronologies differently depending upon the expected signal in the chronology and the goal of

the research. When many chronologies are combined for broad-scale analysis, the raw ring-

width measurements have to be standardized with one technique across all sites, even though the

sites are likely to be affected by different climatic forces. Briffa et al. (2001) reconstructed mean

summer temperature from wood density of tree rings from 387 sites in the Northern Hemisphere

for the past 600 years. They used a new technique called Age Band Decomposition (ABD) in

which the tree-ring series are decomposed into predetermined age bands, averaged together,

scaled for equal mean and variance, then recombined into a master series of relative growth

280
changes. This technique may better preserve the low-frequency (multi-decade to century) signal

in long chronologies.

Dendrochronological records of global climate variability must contain a widely distributed

network of climate reconstructions from around the world. Outside of the North America and

Europe, much explorative dendrochronological research was conducted by LaMarche et al.

(1979a, 1979b, 1979c, 1979d, 1979e) in the Southern Hemisphere in Australia, New Zealand,

South Africa, Argentina, and Chile. Much more work was accomplished in South America as a

dendrochronology laboratory was established in Argentina and local researchers took on tree-

ring investigations (Villalba et al. 1985, Roig et al. 1988, Villalba and Boninsegna 1989, Lara

and Villalba 1993). Natural climate variability in northern Africa and the eastern Mediterranean

region is becoming of great interest as General Circulation Models are projecting severe drying

here associated with increased greenhouse gases (Seager et al. 2007). Several recent tree-ring

studies have been advancing dendroclimatology in the region and have developed a broad

network of tree-ring chronologies (e.g., Touchan et al. 2003, 2007). Recent work in India has

reconstructed the Indian Monsoon back to A.D. 1835 using teak (Tectona grandis) ( Shah et al.

2007). Work in China is examining the climate response on natural forests that are stressed by

the proximity of the Loess Plateau (Du et al. 2007) and a 680 year reconstruction has examined

the strength of the Asian Monsoon and the effect of the Little Ice Age on the Qinghai-Tibetan

Plateau (Huang and Zhang 2007). Work in Siberia has continued to develop long chronologies

from high latitudes that are useful in examining modern climate changes (Vaganov et al. 1996,

Hantemirov et al. 2004). Cook et al. (1992, 2000) reconstructed warm season temperature for

Tasmania from Huon pine (Lagarostrobos franklinii) back to 1600 B.C. This reconstruction

281
showed a weak signal representing the Medieval Warm Period in the 1100s and the Little Ice

Age in the 1600s, suggesting that these events were stronger in the Northern Hemisphere than in

the Southern. Their record also showed higher temperatures over the last 25-year period than

any other time during their 1090 year reconstruction (Cook et al. 1992). Although this finding is

not conclusive proof of global warming, it does support the theory. Climate reconstructions from

New Zealand also show that tree growth since 1950 is significantly higher than any prior period

since A.D. 1500 when the oldest chronologies start (D’Arrigo et al. 1998).

Beyond simple reconstructions of temperature and precipitation, researchers are exploring

correlations with other natural variables and strengthening our understanding of the connectivity

of the global circulation system. The location of the Cook et al. (2000) reconstruction on the

relatively small landmass of Tasmania enabled a sea-surface temperature reconstruction for the

southern Indian Ocean between 30° and 40° S. This teleconnection between marine and

terrestrial systems can help us understand the climate system with its fully complex interactions.

Work in West Africa in a variety of tree species has also been able to show a relationship

between tree growth and sea surface temperature (Schoengart et al. 2006). Villalba et al. (1998b)

examined the connection between tree growth in South America and sea-level pressure over the

Pacific to explain long-term precipitation changes. These connections with the climate system

are leading a better grasp of the complexity of broad scale atmospheric circulation patterns and

interconnectedness of the global system.

282
Climate Indices

The El Niño Southern Oscillation (ENSO) is a prominent example of teleconnections in our

climate system that has raised popular as well as scientific awareness of broad scale climatic

processes. El Niño was a phenomenon noted in the Peruvian fisheries where the waters became

warm around Christmas time and the fisheries failed. Later investigation discovered that this

change, called El Niño, was the same change as the Southern Oscillation Index (SOI) which is

quantified as the variability in the pressure difference between the town of Darwin on the

northern tip of Australia and the island of Tahiti. The scientific community first took note of

ENSO as a global phenomenon during the 1972-1973 El Niño event. The 1982-1983 El Niño

drove greater scientific interest in this phenomenon. Finally, the 1997-1998 El Niño was strong

enough and public media was active enough that El Niño became a household word. The

escalating interest in teleconnections (the interconnection of physical parameters over long

distances such as sea level pressure in the Pacific Ocean affecting weather around the world) has

lead scientists to look for other long-term oscillations in the climate system that may lead to

climate prediction and a deeper understanding of the linked marine-terrestrial system. Stahle and

Cleaveland (1993) used networks of chronologies from Mexico, Texas, and Oklahoma to

reconstruct the SOI for the period from 1699 to 1971. Forty one percent of winter SOI was

explained by tree-ring predictors and the reconstruction correlated significantly with an

independent winter SOI measure. Stahle et al. (1998b) continued to examine winter SOI effect

on terrestrial chronologies from North America. They were able to explain 53% of the variance

in the tree-ring chronologies for the period from A.D. 1706 to 1977. D’Arrigo and Jacoby

(1991) examined millennial length chronologies developed from archaeological wood samples

from the northwest corner of New Mexico. The desert southwest is a region that is strongly

283
affected by El Niño, receiving drier conditions on either side of an El Niño. Principal

components from five of their six chronologies explained 30% of the variance of the SOI data for

the period A.D. 1865-1970 (D’Arrigo and Jacoby 1991).

The Pacific Decadal Oscillation (PDO) (Mantua et al. 1997, MacDonald and Case 2005),

North Atlantic Oscillation (NAO) (Barnston and Livezey 1987), and the Atlantic

Multidecadal Oscillation (AMO) (Gray et al. 2004) are three of the more important marine

phenomena that affect climate in the Northern Hemisphere besides ENSO. Warm phases of

PDO occur when the eastern North Pacific is warm and the central North Pacific is cool; this

temperature gradient switches during cool phases of PDO. Biondi et al. (2001) reconstructed

PDO back to A.D. 1661 based on a network of chronologies of Jeffrey pine (Pinus jeffreyi) and

big-cone Douglas-fir (Pseudotsuga macrocarpa) in a transect from northern Baja California to

southern California. MacDonald and Case (2005) were able to develop a millennial length

record of PDO from limber pine in California that demonstrated that PDO had a strong 50 to 70

year period, but that it was not consistent through time. Over the past 1,000 years, this signal

was only evident for about half of the time. On the eastern coast of North America, NAO is

measured by the height of the 500mb isobar with a positive phase representing below-normal

heights in the high latitudes of the North Atlantic and above-normal heights over the central

North Atlantic. The phase of the NAO affects both the North Atlantic jet stream and the

meridionality of the Rossby Waves (meanders in high-altitude winds associated with the polar-

front jet stream) (Hurrell 1995). The NAO has a 1.7-7.5 year periodicity that creates alternately

cold or warm conditions in Europe associated with this pressure variation. D’Arrigo et al.

(1993) demonstrated a tree-ring response to NAO from Scots pine (Pinus sylvestris)

284
chronologies located in Scandinavia and successfully reconstructed sea-surface temperature for

the North Atlantic back to A.D. 1713. Gray et al. (2004) found that the AMO, which is a 60-100

year variation in Atlantic sea-surface temperature, was a consistent pattern throughout their

record extending back to A.D. 1567.

Climatic Gradient Studies

Dendrochronological studies along gradients can be useful to examine climatic and

environmental changes. Because ecotones are by definition transition zones, they are

particularly sensitive to climate change and will rapidly bear evidence of shifting vegetation

responses. High latitude studies are likely to show the first evidence of global warming because

these locations are likely to warm more than the middle latitudes. Elevational gradients show

the same vegetation changes as latitude gradients, but over a much shorter spatial scale. Both

ecotones and high latitudes are being studied to see if predicted changes to temperature are

occurring and to determine the vegetational response to these changes.

Latitudinal Gradient

Jacoby et al. (1996) examined climate response of trees for a North-South transect through

Alaska covering from 62°N to 72°N latitude. In the 300-year record of their northern

chronologies, they found recent decades exhibited a warming trend. Their southern chronologies

and those located along the coast with a strong maritime influence did not show this warming

trend. This evidence fits with the predictions of general circulation models and our

understanding of the importance of gradient studies. Jacoby and D’Arrigo (1999) reviewed four

285
climate reconstructions from northern latitudes and high elevation (in Mongolia, Siberia, Alaska,

and a general Northern Hemisphere reconstruction) and found that all of these reconstructions

showed unusual and persistent warming since the 1800s. They also noted other evidence such as

glacial retreat and that trees once limited by low temperatures were becoming limited by

moisture stress instead. D’Arrigo and Jacoby (1993) found that trees growing at their northern

limit showed an increase in growth since the mid-1800s which is consistent with expectations of

global warming. The authors successfully modeled the observed increase in tree growth with

projected changes in climatic parameters such as temperature and precipitation with little

residual signal. Because other climatic parameters explained these changes, it was concluded

that there was no direct affect from CO2 fertilization.

Treeline Studies

Trees growing at treeline are often limited by cold temperature and can be useful for long term

temperature reconstructions (e.g. Esper et al. 2003). Treeline studies are an alternate use of

dendrochronology to reconstruct climate through its effect on establishment of trees at the cold-

limited high elevation extent of the species (Nicolussi et al. 2005) and can be an indication of

broad scale climate changes. Esper and Schweingruber (2004) report on a recent broad-scale

treeline advance throughout the northern Arctic region, suggesting that this present trend is

associated with the reported Northern Hemisphere warming. LaMarche and Mooney (1967) and

LaMarche (1973) examined remnant bristlecone pine (Pinus longaeva; Figure 8.3) wood from

above treeline and found that warmer conditions lasted from the beginning of their record at

5,300 B.C. to 2,200 B.C. extending treeline. Cooler and wetter conditions lasted from 1500 B.C.

286
 

Figure 8. 3 Bristlecone pine trees in Methuselah Grove. For climate reconstructions, we go to


high latitude or high elevation (as in this picture) to get tree ring chronologies that are sensitive
to temperature. This is the Methuselah Grove of the White Mountain Bristlecone Pine Trees in
California. Most of the trees in this picture are between 3,000 and 4,000 years old. The old age
of these trees, combined with crossdated samples of the sub-fossil wood on the ground, allow the
reconstruction of long chronologies of temperature fluctuations extending back approximately
8,000 years before present (photo by Jim Speer).

287
to 500 B.C., and then a cool dry period dominated from A.D. 1100 to A.D. 1500 resulting in the

lowering of treeline.

Treeline sites can also exhibit stressful conditions in which trees in tropical environments

produce annual rings, even though excess moisture prevents annual ring formation at lower

elevations (Speer et al. 2004). In the Dominican Republic at 19.5°N Latitude, heavy rain in the

lowlands makes West Indian pine (Pinus occidentalis) trees produce 3-4 rings per year (FAO

1973, van der Burgt 1997). At the highest elevation on the island, and along the margin of an

area of loose rocks without much soil development, the trees are stressed enough by the lack of

moisture that the January to March dry season forces them to systematically shut down,

producing reliable annual rings (Figure 8.4).

Dendrohydrology: Water Table Height and Flood Events

Water table changes, land subsidence, flood height and energy, and streamflow can all be

reconstructed using tree-ring data. Dendrohydrology is the subfield of dendrochronology that

uses tree rings to reconstruct these phenomena (Schweingruber 1996). Dendrohydrological

records can be reconstructed through suppression or release events in trees associated with water

table changes and land subsidence, scars and growth changes associated with flood events,

establishment of trees on newly deposited surfaces (Figure 8.5), and changes in growth as a

response to climatic phenomena that drive river discharge.

Stream behavior can be documented by a variety of effects on tree growth that include flood

scarring, tree leaning from undercutting, and establishment of trees on new sediment surfaces

288
 

Figure 8. 4 Old Pinus occidentalis growing on a high elevation site in the Dominican Republic.
Annual dating of tree rings in the tropics is possible on unique sites where the tree growth shuts
down for some part of the year. This site is called Conuco del Diablo (Cornfield of the Devil)
and is located on the flank of one of the highest peaks in the Dominican Republic. Trees stop
growing in the dry season from January to March in part because the rocky earth prevents much
soil development or water retention. These trees have been used to build a chronology for the
region (see Speer et al. 2004).

289
 

Figure 8. 5 The age of a delta or any sedimentary deposit can be determined from trees growing
on that sediment. In this photo, four different age surfaces are discernable based on the structure
of the vegetation, with the youngest being the bare sediment in the delta, the second oldest is the
low vegetation in the left of the image, the third oldest is just above that and inland from the
road, the oldest vegetation is on the hill slope at the top of the image (photo by Jim Speer).

290
(Gottesfeld and Gottesfeld 1990) (Figure 8.6). For example, a flood history of the Potomac

River in Washington D.C. was determined using tree rings to date flood scars (Sigafoos 1964).

The height of the scars can also be used to document the height of flood waters in the past.

Begin (2000) recorded ice scarring on the trees surrounding lakes in Quebec, Canada to record

lake flood events. Yanosky and Jarrett (2001) found distinct variations in the wood anatomy of

oak trees; they identified white rings that were formed from open fibers when a tree’s root were

submerged in water and earlywood vessels when a tree was submerged and stripped of leaves at

the end of the growing season. These distinct anatomical changes are an excellent indicator of

past flood damage.

Information about long-term patterns of streamflow, flooding, and water level in reservoirs is

relevant to anyone who makes decisions about water allocation (Meko and Woodhouse in press).

Streamflow reconstructions, for example, can help municipal water managers plan for the natural

variability in water resources (Woodhouse 2001). Correlations of ring width to streamflow data

from the Colorado Front Range was used to reconstruct streamflow along the South Platte River

and Middle Boulder Creek back to A.D. 1703 (Woodhouse 2001). Stockton and Jacoby (1976)

reconstructed stream flow for 12 stream gauge stations in the Upper Colorado River Basin and

found that flow was at record high levels in the early 1900s based on their 450-year

reconstruction. This meant that the water allocation for the Colorado River based on the early

1900s levels could not be met during a normal year of stream flow. This was actually known at

the time of the allocation decision based on research done by Douglass and Schulman. The

commission reduced the amounts that were allocated because of this higher growth shown in the

tree-ring chronologies, but they did not adjust it enough (see Schulman 1938 for published

291
 

Figure 8. 6 Flood events can damage trees in many ways, providing dendrochronologists with
different approaches to reconstruct flood activity. Establishment dates can also provide timing
for these events (from Gottesfeld and Gottesfeld 1990).

292
reconstructions). Cook and Jacoby (1977) used standard climate reconstruction techniques to

document drought frequency in the past as it relates to water supplies in the Hudson River Valley

of New York. Other work has demonstrated the direct connection between frequency of drought

events and the reliability of water reserves in various reservoirs (Stockton and Jacoby 1976, Jain

et al. 2002, Woodhouse et al. 2006).

Another application of dendrohydrology is to provide information about the timing and extent of

ecological changes as water levels rise or lower. Phipps et al. (1979) used the growth of loblolly

pine near the margin of the Great Dismal Swamp to document anthropogenic ditching and

subsequent drainage of the swamp. Schweingruber (1996) observed a similar phenomenon with

spruce trees growing along the margin of a bog in the Swiss Jura Mountains. The marked release

in growth made dating the drainage events readily observable. Changes in hydrology can also be

related to landslide events, broadly called mass movement events.

Segment-Length Curse

The ability to obtain a low-frequency climate signal from a chronology is dependent in part upon

the length of the individual ring-width series that contribute to the chronology. The limitation is

imposed by the detrending of individual ring-width series in tree-ring standardization, and

becomes a problem when trying to reconstruct climate over a long period of time from a

chronology that has been formed by splicing together relatively short tree-ring segments.. This

phenomenon, called the segment length curse, was originally proposed by Cook et al. (1995),

who demonstrated the issue with modeled chronologies composed of sine waves with 1000, 500,

293
and 250 year wavelengths, and supplied illustrations with real chronologies from the bristlecone

pines in the White Mountains of California.

The segment-length curse must be considered when examining long chronologies that have been

constructed from shorter series. Cook et al. (1995) used a very conservative spline of a

horizontal line fit through the mean of the series and, with the bristlecone pine, a negative

exponential curve fit through the series. They suggested that the Regional Curve

Standardization (RCS; Briffa et al. 1992) technique may better preserve the low-frequency

information of long chronologies composed of many short series such as the European oak

(Quercus sp.; Pilcher et al. 1984) chronology or Scots pine (Pinus sylvestris) chronology from

Fennoscandia (Briffa et al. 1990). St. George and Nielson (2002) applied this technique to oak

trees in southern Manitoba to reconstruct hydroclimatic events while maintaining the long-term

variability in their chronology.

Archaeological Uses of Climate Reconstructions

Climate reconstructions have long been used to provide an environmental backdrop to the

settlement patterns of native populations (Dean et al. 1985, Dean 1997). Grissino-Mayer (1996)

developed a climate reconstruction for El Malpais, New Mexico, that extends back to 100 B.C.

This 2000-year long precipitation record delineates drought and moisture episodes for much of

the southwestern U.S. He compared this climate reconstruction to major changes in Native

American settlement patterns and found that settlement patterns change during periods of high

variability in climate. Stahle et al. (1998a) examined bald cypress (Taxodium distichum) growth

in southern Virginia and found that the Roanoke and Jamestown colonies were established

294
during two of the most severe droughts recorded in their 800-year record. Both of these colonies

struggled and Roanoke failed soon after they were established which demonstrated the

communities’ reliance upon the resources provided by a temperate climate. (See Chapter 7 for

additional discussion.)

Use of Climate Reconstructions for Future Prediction

Policymakers and laypeople are interested in discerning the future climate of the Earth. Studies

of past climate can give us an idea about climatic variability and the causal mechanisms that

should hold true in the future (Vaganov et al. 1999, Briffa 2000). For example, Cook et al.

(2004) examined climate variability for the past 1200 years in the western U.S. and found that

instances of higher temperature (such as the Medieval Warm Period, here reconstructed as A.D.

900 – A.D. 1300) corresponded to drought, suggesting that future climatic warming may result in

an increase in aridity in this area. Other tree-studies have examined the effects on past stream

flow and water supplies, changes in tree response to climate, and changes in related climate

parameters (see Chapter 10). Studying the past provides researchers with an understanding of

the natural range of variability so that we can prepare to adapt to climatic changes in the future.

295
Chapter 9: Dendroecology

Introduction

Dendroecology uses dated tree rings to study ecological events such as fire and insect outbreaks.

Dendroecology was developed as a field of study by Theodor Hartig and Robert Hartig in the

late 1800s in Germany, with Bruno Huber continuing the tradition from 1940-1960

(Schweingruber 1996). In the United States, dendroecology did not develop until the 1970s with

early work proposed by Hal Fritts (Fritts 1971). Since the 1970s, dendroecology has greatly

expanded (see Fritts and Swetnam 1989) to include the study of fire history (Dieterich and

Swetnam 1984), insect outbreaks (Swetnam et al. 1985), masting (synchronous fruiting in trees;

Speer 2001), stand-age structure (Lorimer and Frelich 1989), pathogen outbreaks (Welsh 2007),

and endogenous disturbance history (Abrams and Nowacki 1992). I exclude from

dendroecology the subfields of dendroclimatology and geological applications of

dendrochronology which are included in the definition given for dendroecology in the

Multilingual Glossary of Dendrochronology (Kaennel and Schweingruber 1995). All three of

these subfields of dendrochronology have a sufficient amount of research, refined methods, and

techniques to be addressed independently. For this work, I define dendroecology as analysis of

ecological issues such as fire, insect outbreaks, and stand-age structure with tree rings.

Therefore, dendrohydrology, dendroclimatology, and dendrogeomorpology are described in

Chapters 8 and 10. Furthermore, some research tools used in dendroecology such as stable

isotopes, dendrochemistry, and x-ray densitometry are treated in greater depth in Chapter 11.

296
Methods for Dendroecology

General methods of dendroecology usually involve the standard field and laboratory analyses

described in Chapter 5, with particular emphasis on establishment dates for succession studies,

scarring from fires, or suppression and release events to document insect outbreaks or episodes

of logging. Some methods are specific to dendroecology; for example, in order to determine

exact establishment dates, cores are often taken at ground level and special care is directed

towards obtaining pith.

Stand-Age Structure

A stand-age structure analysis is a useful technique for many of the studies that follow. Stand-

age structures require that all living and dead trees be sampled in a plot to quantify the current

forest composition and past conditions (Lorimer and Frelich 1989, Abrams et al. 1995, Bergeron

2000, Daniels 2003). Exact age dating is important to provide a complete picture of the

establishment and mortality of all tree species on the plot (Gutsell and Johnson 2002). This

method can be used as the basis for a study on succession dynamics (Abrams et al. 1995,

Bergeron 2000), endogenous disturbances such as gap dynamics (Kneeshaw and Bergeron

1998), or for fire history in a stand replacing fire regime (Bergeron 2000).

A stand-age structure analysis can be conducted in a circular plot, square plot, or along a

transect. One preferred technique is to use a band transect. In this method, a tape measure is laid

out for the length of the transect (50 m for example) and all living and dead trees within a

designated distance (e.g. 1 meter) of either side of the tape are sampled. Although sampling

along the transect is very intensive, the larger size classes are often not well represented because

297
larger trees tend to grow at lower densities than smaller trees. Subsequent bands can be added to

the outside of the base transect, therefore, in which trees of large diameter at breast height (DBH)

are sampled. For example, in addition to the complete sample within one meter of either side of

the tape, all trees greater than 20 cm DBH within two meters of either side of the band can be

sampled, and all trees greater than 40 cm DBH within three meters of either side of the tape can

be sampled. This sampling technique results in a 100 percent sample of all age trees in a 2 X 50

meter transect and all larger trees within a four meter or six meter swath, increasing the sample

depth of the larger trees. Only the data from the inner transect is used when estimating the

number of trees in each age class or when quantifying the percentage of each species recorded.

Careful field notes should be kept during sampling that record the location of each tree that is

cored, the species, DBH, and sample ID of each tree. With this data, one can examine the

interaction among trees and plot out the distribution of different tree species along the transect.

Transects are particularly useful for crossing boundaries or ecotones, such as examining

vegetation patterns on either side of a stream channel or across an edge from woodland to forest

interior. Transects can also be run along contours to maintain a similar sampling pool of trees

that are growing at the same elevation and aspect.

Selective and opportunistic sampling can be used to achieve a stand-age structure study that

would otherwise be too difficult to sample or be untenable due to a lack of permission to sample

extensively. Samples were taken from stumps in a clear-cut forest in coastal British Columbia in

order to analyze the regeneration and growth characteristics of western red cedar (Thuja plicata)

(Daniels 2003). Western red cedar develops heartwood decay and can grow to diameters of 160

cm, making accurate sampling of pith dates difficult with an increment borer. Daniels (2003)

298
was able to determine that tree size was a poor indicator of tree age, recruitment was continuous

and most likely affected by light gaps more than broad-scale disturbance, and mortality events

were gradual and continuous through time. She concluded that current regeneration rates were

sufficient to maintain the species on these sites (Daniels 2003). Daniels’ research demonstrates

the usefulness of opportunistic sampling when cross sections or intensive sampling are required.

Ring Width Analysis

Analysis of ring widths can be used to deduce disturbance events such as suppression from insect

outbreaks, decline in growth associated with atmospheric pollution, trade-off between

incremental growth and reproductive effort (masting), and release events associated with growth

into canopy gaps. These applications use the same basic techniques as a dendroclimatic study,

except that climate is often noise in these chronologies and should be controlled or removed so

that the effect of the disturbance agent can be isolated. If a disturbance is expected to be

recorded throughout the entire stand, such as with an insect outbreak study, older trees will often

be targeted for sampling because they provide a greater temporal depth. Sometimes, complete

stand inventories are conducted and ring widths are examined on trees throughout the stand such

as with a stand age structure (Filion et al. 1998). Individual disturbances can also be targeted,

for example trees growing in or around a light gap, in order to document the timing and response

to such a disturbance (Thompson et al. 2007).

Tree Scars

Various types of ecological phenomena may leave scars on trees that provide a record of events

for the dendrochronologist. A commonly studied disturbance is fire which causes scars in the

299
trunk when part of the cambium is killed by excessive heat during a surface fire (see Swetnam

and Baisan 1996). Animal herbivory and damage to roots by mifrating caribou on trees also

leaves behind a record of their effect and can even be used to estimate population levels (Spencer

1964, McLaren and Peterson 1994, Payette et al. 2004).

Basal Area Increment

Basal Area Increment (BAI) can be used as a type of standardization (see Chapter 2 for more on

standardization) in which annual growth increments are calculated by subtracting the area of a

cross section in year t-1 from year t. The result is an estimation of the two dimentional growth

increment added to the cross sectional area of the tree. This calculation is useful because it

removes any age-related growth trend resulting from adding the same volume of wood on an

ever increasing cylinder, while maintaining suppression and release events that may be due to

forest disturbances (LeBlanc 1990a, Phipps 2005).

Applications of Dendroecology

Gap Phase Dynamics

Intensive sampling coupled with a detailed analysis of tree growth histories and stand-age

structure can result in a chronology of gap-phase dynamics in which trees respond to openings in

the canopy due to dominant tree mortality. Tree-level ring-width series record suppression and

release events, enabling dendrochronologists to document major disturbances that affect the

growth of mature and understory trees (Lorimer and Frelich 1989). Gap-phase dynamics are the

most common disturbances in many closed-canopy forest sites including the tropics and

temperate forests such as the eastern deciduous forest in the United States (Picket and White

300
1985) and old-growth forests of the Pacific Northwest (Lertzman et al. 1996). Gaps provide

limited resources, sunlight and growing space, to the lower levels in a dense forest, and may

provide the chance for suppressed understory trees to recruit into the canopy. Kneeshaw and

Bergeron (1998) examined gap dynamics in the southern Boreal forest near Quebec. They found

that aspen (Populus tremuloides) established soon after fire and as the forest aged it became

dominated by individual tree gap dynamics. Over time, balsam fir (Abies balsamea) replaced

aspen and larger gaps occured as a result of spruce budworm defoliation. As gap size increased,

shade intolerant species were preferred and eastern white cedar (Thuja occidentalis) dominated.

Daniels (2003) found a similarly complex establishment pattern in coastal British Columbia

where western redcedar (Thuja plicata), western hemlock (Tsuga heterophylla), and Pacific

silver fir (Abies amabilis) competed for dominance. Western hemlock and Pacific silver fir

depended on gaps to recruit into the canopy while western redcedar did not require gaps and

could regenerate under the closed canopy.

Forest Productivity and Succession

Ring-width analysis can be used to examine forest health and productivity over time (Kienast

1982, Eckstein et al. 1984, Greve et al. 1986, Graumlich et al. 1989, Biondi 1999). For example,

Biondi (1999) examined the growth of ponderosa pine (Pinus ponderosa) in a forest stand near

Flagstaff, Arizona. The stand did not experience wild fires during the 20th century and Biondi

(1999) documented a decline in growth since 1920 associated with increased forest density.

These changes were compared to the past 400 years of tree growth to determine if this was a

unique pattern in the history of the stand. In another project, Graumlich et al. (1989) examined

changes in net primary productivity (NPP) of forests by measuring ring widths, demonstrating

301
that forests in the Cascade Mountain Range in western Washington State were increasing in

growth. They attributed this recent increase to warming summer temperatures and increased

absorbed solar radiation rather than any CO2 fertilization effect. Cook et al. (1987) took a

different approach to examine forest decline in red spruce in the Southern Appalachians. They

examined the relationship between climate and tree growth and found that the trees stopped

responding to climate after 1967 and started to decline. They suggested that this response, not

related to climatic forcing, could be due to anthropogenic pollution. Fir waves are another

phenomenon in the eastern United States in which large swaths of fir trees die synchronously,

but the agent for this mortality was unknown until the 1980s. Marchand (1984) examined tree

growth and age structure in these fir stands and found that wind abrasion on the windward side

of the stand caused these trees to die back while other fir trees established on the leeward side of

the stand. The fir wave phenomena did not kill the stand but created a banded pattern of

different stand ages. All of these specific studies are examples of how dendrochronologists can

reconstruct the past to determine if modern growth of trees is similar to growth in the past,

providing temporal perspective to current forest health issues.

Recent work in dendroecology has produced a new technique (dendromastecology) for

developing mast (massive fruit production in trees; specifically acorns in this case)

reconstructions from tree rings (Speer 2001). Mast reconstructions require the researcher to

consider multiple variables in a tree’s signal, peeling apart one layer of signal after the other,

until a large percentage of the overall pattern can be explained by climate and biological factors

that control tree growth. The aggregate tree growth model discussed in Chapter 2 is a conceptual

model of this approach, where the age-related growth trend can be standardized out of the series,

302
and then climate can be removed through regression analysis with significant correlates in the

climate variables. Finally, the signal that is left may be a biological signal such as masting in the

tree (Speer 2001). Mast reconstruction is a new technique that still has to be tested in multiple

species and locations around the world. Speer (2001) demonstrated that mast reconstructions

were achievable on oak species in the eastern United States; good climate data and a long

calibration data set of masting in the tree species of interest, however, are necessary for this type

of work. Speer (2001) also found that there was not a strict trade-off between incremental

growth and reproductive effort as was suggested by ecological theory. Instead, the oaks in his

study had the ability to store carbon as carbohydrate and starches (most likely in their roots) so

that the carbon drain for acorns in one year was actually carried by energy production over

multiple years. This suggests that mast reconstructions will not be evident on all sites, because

Speer (2001) found only 25% success in mast reconstructions in his site/species chronologies.

Mast fruiting may be a strong control of seed predator populations and also affect the ability of

masting species to compete in a diverse forest.

Forest succession, the development of the forest community through time on a site following a

stand-repalcing disturbance such as glacial retreat, volcanic eruptions, or farming can be assessed

by examination of establishment dates of different tree species on one or multiple sites (Abrams

and Nowacki 1992, Fastie 1995). Fastie (1995) explored primary succession dynamics in

Glacier Bay Alaska after the glaciers had retreated. He used historical records of glacier retreat

for the past 250 years to document the age of his plots located at increasing distances from the

present foot of the glacier. Fastie (1995) found that in this area, primary forest succession was

accelerated because of proximity to seed sources along the trimline of the glacier, which is the

303
highest elevation that the glacier ice reaches. Mature trees left above the trimline produced seed

that repopulated the newly exposed ground. Abrams and Nowacki (1992) demonstrated variable

mechanisms of succession by documenting a case of accelerated succession in the deciduous

forests of Pennsylvania that was due to an exclusion of fire followed by thinning of the overstory

trees in multiple logging operations. This management practice removed fire tolerant oak and

pine trees from the site and encouraged late-sucessional red maple (Acer rubra), sugar maple

(Acer saccharum), and black cherry (Prunus serotina) to gain dominance in a forest that used to

be maintained by frequent fire. Dendrochronologists can add information to larger ecological

theories through exact dating of establishment and growth of trees in complex forested

ecosystems.

Old Forests

Dendrochronologists tend to find old trees in surprising locations. The concept of “longevity

under adversity,” first published by Schulman (1954), has lead dendrochronologists to find

extremely old trees on lava flows, cliff faces, and high mountain peaks. Since then, many

dendrochronologists have continued to find old trees in unexpected places. Stahle (1996)

documents old trees throughout the eastern United States in locations that were previously

assumed to have been completely logged around the 1900s. Most eastern states still have living

trees that established prior to the founding of the United States (Stahle 1996). Orwig et al.

(2001) document four sites within 80 km (50 miles) of Boston that have trees in excess of 250

years old. Kelly et al. (1992) record eastern white cedars growing on the Niagara Escarpment in

southern Ontario that are over 1,000 years old. In the western U.S., Grissino-Mayer et al. (1997)

developed a 2,000 year climate reconstruction from trees growing on the El Malpais lava flow in

304
northwestern New Mexico and from dead trees that were well preserved on the rocks of the lava

flow. They found some of the oldest documented Douglas-fir (Pseudotsuga menziesii) and

ponderosa pine trees growing in this extremely harsh environment with roots delving into the

volcanic rock under a thin film of soil. Bristlecone pine trees (Pinus longaeva), the oldest living

non-clonal organisms (Currey 1965, Ferguson 1968), epitomize the principle of “longevity under

adversity.” Found in the White Mountains of eastern California in cold, arid conditions at

elevations of more than 10,000 feet, these trees grow so slowly that more than 100 rings may be

contained in less than an inch of wood. As in many areas where long-lived trees are found,

bristlecone pine forests have very little understory, usually only a few tundra herbs. The oldest

trees studied by dendrochonologists are often not found in classic old-growth forests with

complex understory vegetation, but are instead discovered in sites with sparse tree density on

harsh sites (Orwig et al. 2001).

Dendrochronologists have been documenting the age of the oldest trees in each species that

produces tree rings, providing an understanding of the maximum age attainable by each species

(Brown 1996). By recognizing the maximum age attainable by each species, managers can

formulate strategies that take into account the temporal scale of the trees they manage.

Documentation of the maximum age of trees and the sites on which they grow enable researchers

to better understand the concept of old-growth forests and even provide insight into how

organisms age and survive to extreme old age (Appendix B; see the World Wide Web for an

updated OLDLIST and Eastern OLDLIST).

305
Dendropyrochronology

Reconstruction of fire histories is one of the major applications of dendrochronology for use in

management of forests and the reestablishment of fire as a disturbance agent. Any prescribed

fire policy in the United States must be supported with scientific evidence for that tract of land.

These federal and state laws have motivated fire reconstructions on many parcels of land which,

in turn have generated a comprehensive view of the role that fire plays in these fire-prone

landscapes. The goal of the dendrochronologist is to determine the natural range of variability

for fire on a particular site (Landres et al. 1999). The natural range of variability describes the

past occurrence of fire, how frequently it affects a site, and the area that it has covered in the

past. From this information, forest managers can determine how fire has behaved on their land

in the past and how fire regimes have changed in the 20th century (Heyerdahl and Card 2000).

Three main fire types occur around the world. A surface fire is one that burns over the ground

surface, consuming duff and fine fuels. These fires usually move through an area fairly quickly

and burn at a low to moderate severity. Many forest types, such as ponderosa pine (Figure 9.1),

red pine, and giant sequoia (Figure 9.2) depend on these frequent low-severity surface fires to

remove competition and to burn off the duff layer, allowing seedlings access to mineral soil.

Oak woodlands also seem to be dependent upon frequent fire to maintain this forest type.

Stand-replacing fires occur less frequently when fuels have built up to a critical level and often

cause high tree mortality. These fires will often burn through the canopy of the trees and,

therefore, are also called crown fires (Figure 9.3). Some pine forests, such as lodgepole pine,

are adapted to this type of fire. Stand-replacing fires burn through a forest and kill the mature

trees. Many of the trees that are adapted to a stand-replacing fire regime have serotinous cones

306
 

Figure 9. 1 A ponderosa pine stand in Oregon that has received multiple thinning and prescribed
fire treatments. Note the triangular catface (fire scar wound) at the base of the closest tree on the
left. The trees record multiple fires that burn through the site at low intensity (photo by Jim
Speer).

307
Figure 9. 2 A catface can be a huge scar when it occurs in giant sequoia. Giant sequoia is a fire
adapted tree species that needs fire to regenerate (photo by Jim Speer).

308
Figure 9. 3 Stand replacing fire in Pinus sylvestris. This is a crown fire that will burn through
most of the stand, killing the mature trees. These trees have serotinous cones that open and
disburse seeds when the cones are heated, leaving behind a seed bank on a rich mineral soil
which starts the regeneration process after the fire (photo by Tom Swetnam).

309
which only open to spread their seed when they are heated, as a coating of resin or woody layer

is burned off. Stand-replacing fires occur frequently in the boreal forest where dry summers in a

continental climate combined with little topographic relief, warm winds, and convective storms

result in fire that can burn a large area of the landscape. The third type of fire is a ground fire

that actually burns under ground in the organic-rich soils of histosols. These fires are common in

Alaska where they can burn for more than 30 years as they smolder through the thick organic

layers of plant material on the ground.

Surface Fire. Each type of fire requires a different sampling method to accurately record the

occurrence of fire in the past. Surface fire regimes will burn frequently on a site but leave

mature trees alive that record the fire. Pine ecosystems seem to be the most adapted to surface

fires and are most frequently sampled for long-term fire history. Fire histories in this type of

forest are most productively accomplished by cutting fire scarred samples from stumps, down

logs, and living trees from these sites. Surface fires are often pushed by the wind and move

uphill as they consume fuels. Fire can burn more intensively on the uphill sides of the trees

because the fire can eddy there in a vortex from the rising air currents. Also, pine needles and

cones collect on the uphill side of the tree and provide more fuels for the fire to burn hotter at

this location. The first time a tree is scarred, the fire does not usually damage the xylem of the

tree. The cambium is killed because the fire heats it through the bark, later causing the bark to

slough off. The scarred part of the tree will have an area of thinner bark and, if a pine tree, pitch

will collect in it. Once the tree is initially scarred, it is more likely to along the exposed portion

of the cambium when subsequent fires burn. Repeated fires create a triangular scar, called a

catface, at the base of the tree (Figure 9.4).

310
Figure 9. 4 A catface scar on a living ponderosa pine tree with a partial section removed from
the left base of the tree. Subsequent fires kill off the living cambium and leave behind a scar
which the tree tries to heal over. By tracing the vertical fissures in the wood that follow the ring
boundaries, one can count the number of these scars on the face of the wound to get an estimate
of the number of fire scars preserved on the sample. Then a decision is made about which trees
to sample based on the number of scars recorded and on the preservation of the sample (whether
there is much rot) (photo by Jim Speer).

311
Repeated fire scars in a catface can be sampled by taking a partial section from living or dead

standing (snags) trees (Figure 9.5; Arno and Sneck 1973, Cochrane and Daniels 2008). This

sampling technique involves two cuts with a chainsaw along the cross sectional surface through a

cat face followed by two plunge cuts along the edges of the sample to break it loose. The

resultant sections have all of the fire history information from bark to pith (Figure 9.6) while

leaving most of the base of the tree for stability and transport of substances through the xylem.

Heyerdahl and McKay (2001) reexamined 138 trees six years after sampling partial sections for a

fire history reconstruction in order to investigate the impact of fire scar sampling on the health of

the trees. They estimated that only 8% of the cross sectional area was removed in sampling and

that these trees did not have greater mortality than a control group of 386 similar sized trees that

were not sampled for fire history. They conclude that partial sampling from the catface of pine

trees is a non-lethal sampling technique that provides needed information for land management

(Heyerdahl and McKay 2001).

Hardwood trees can also scar from surface fires. This has most frequently been recorded in oak

trees growing in open woodland settings (Abrams 1985, Smith and Sutherland 1999, 2001). In

this case, the trees tend to grow on flat ground and fire scars are not recorded in a catface, but are

recorded on multiple sides of the stem wherever a fissure in the bark allows the cambium to heat

up to a temperature which can kill the cambium. Therefore, old oak trees can record multiple

fire scars, but a full cross section is needed to document past fires. Some hardwood trees

develop catfaces due to successive fires, but these are not common as most hardwood trees

develop rotten wood near the wounds, which can obscure the tree rings and the scar (Speer

unpublished data).

312
 

Figure 9. 5 We take partial sections from living trees to get a complete history of fire through
the modern era, but leave the tree standing and healthy. In this process, the fieldworker uses a
chainsaw to take two horizontal cuts and two plunge cuts to remove the fire scarred section while
leaving most of the tree behind for support and conduction of fluids (from Arno and Sneck
1973).

313
 

Figure 9. 6 A partial section from ponderosa pine with a close up showing the fire scar dates.
This partial section only removes a small portion of the living cambium and the rest of the tree is
left for support and growth. The individual scars are obvious from the bark to the inside of the
tree where it started to record fires. By looking closely at the fire scarred samples, the season of
the fire can be determined based on the position of the scar on the earlywood or latewood within
the ring (photo by Jim Speer).

314
Stand-replacing fire regime. Past fire occurrence of a stand-replacing fire type can be

documented by stand-age structure recording the time since last fire (Heinselman 1973). This

technique uses a stand-age structure in areas that have been identified as different fire events

based on areal photos or a stratified random sampling protocol across the landscape to determine

where the age-breaks occur (Johnson and Gutsell 1994). Within the bounds of each area, a plot

can be established and a stand-age structure analysis conducted. This stand-age structure is

likely to demonstrate that all of the trees in each patch are a single cohort with similar

establishment dates controlled by the time since the last fire. This type of study requires the

researcher to take a core near the ground surface and hit pith to get the most exact age estimate

for each tree (see stand-age structure methodology above). Extensive landscape studies that

examine fires scars along with establishment dates and survivorship curves can be used to

interpret the entire fire history of a landscape that can include surface and stand-replacing fires in

lodgepole pine and sub-alpine fire forests (Sibold et al. 2006, 2007).

Ground fires. Ground fires can possibly be documented from root damage or scarring to living

trees, although I am unaware of any dendrochronological studies that have examined these types

of fire regimes. Most ground fires burn in regions with a rich organic peat layer that formed

from an old bog. These areas do not always support many trees and any trees that were able to

grow there may be killed by the passing of the fire. As long as the mortality of the tree can be

attributed to the fire, the death dates of those trees can be used to determine the time of the event.

315
Seasonal Resolution of Fire Scars. Just as the year in which a fire occurred can be ascertained

through crossdating, the season of burn can be determined from the position of fire scars in the

earlywood or latewood of some trees (Swetnam and Baisan 1996, Figure 9.7). For example, if a

few cells of earlywood were formed before the cambium was killed and seasonal growth was

stopped in that part of the tree, researchers can infer that the fire occurred early in the tree’s

growing season, often in the spring. A latewood scar indicates a fire burned the tree late in its

growing season, and a scar between two fully formed rings signifies a dormant season fire.

Knowledge of the season of past fires enables land managers to reintroduce fire to the landscape

in a natural way. If fires are forced on the landscape in a different season than occurred

naturally, different plant species will be favored by affecting sprouting, fruiting, and flowering.

Native American use of fire can also potentially be determined by looking for a change in the

fire season from the natural fire regime.

Fire in the southwestern United States. Studies of fire effects on trees have been conducted since

the early 1900s (Clements 1910, Anonymous 1923, Show and Kotok 1924, Presnall 1933). In the

1980s, crossdated fire histories became much more common in the southwestern U.S., leading to

a better understanding of the spatial and temporal patterns of fire in many pine ecosystems

(Madany et al. 1982, Dieterich and Swetnam 1984, Swetnam et al. 1999). Forest managers can

use prescribed burning to return the forests to a more natural condition and improve forest health

(Swanson et al. 1994, Morgan et al. 1994, Fule et al. 1997, Landres et al. 1999, Swetnam et al.

1999). Fire history records easily extend before the late 1800s, which was a time of heavy

grazing by sheep and cattle and was followed by fire suppression by the U.S. Forest Service

(Savage and Swetnam 1990). Swetnam and Baisan (1996) demonstrated how

316
 

Figure 9. 7 A fire scar is a three dimensional wound where the living cambium meets the dead
cambium. Right at that point we can determine the year and the season in which the fire
occurred based on the position of the scar in the ring. If a few early wood cells formed before
the cambium was killed, then it is termed an early-earlywood scar. If the scar appears in the
middle of the earlywood, then it is named a middle-earlywood scar. The same is true for the
late-earlywood fire scar. When the scar appears only in the latewood, it is called a latewood
scar. If the scar appears right on the ring boundary with neither early wood formed before it or
latewood formed around it, it is called a dormant season fire. The season of the dormant fire
scars can generally be assigned to spring or fall by the dominance of other scars in the earlywood
or latewood from that site (from Swetnam and Baisan 1996).

317
dendrochronological records can be used to examine fire across multiple spatial scales (Figure

9.8). Fires history data is collected on the individual tree basis. Researchers can also examine

whether fires are recorded on multiple trees, and thus reconstruct a fire’s spread through a site.

Then multiple sites can be examined in a watershed to see how the fire has spread across the

watershed. Finally sites throughout a region can be examined to identify common fire years that

occur in many separate watersheds because of the appropriate climatic forcing. The result is a

reconstruction of fire on multiple spatial scales and different driving factors can influence fire

events at each scale. For example, Swetnam and Betancourt (1990) have shown that fire

occurrence in the Southwest can often be explained by climate patterns, especially at the broad

scale. Swetnam et al. (1999) examined fire histories from 55 sites throughout the southwestern

U.S. that extended back to the 1600s (Figure 9.9) and found that regional fire occurrence was

driven by long-term climate fluctuations such as El Niño.

Fire in Scandinavia. New innovations in the methodology to examine the spatial dimension of

fire are being conducted in Sweden (Niklasson and Granström 2000). Dendropyrochronology

has had great success in documenting the temporal component of past fires, but has lacked a

rigorous systematic sampling across space that could enable the reconstruction of the spatial

aspect of surface fires. Niklasson and Granström (2000) collected 1133 samples from 203 points

that were approximately spaced two kilometers apart across an area covering 19 X 32 km. This

network of sampling points enabled researchers to examine the area burned by past fires in a

Pinus sylvestris chronology extending into the 1100s, which successfully documented the spread

of fires across the landscape so that spatial and temporal patterns could be compared.

Researchers also found an impact from human-caused ignitions in Scandinavia by early settlers

318
 

Figure 9. 8 Fire history data can be collected on multiple spatial scales to understand the driving
factors of this natural disturbance. Fire scars are collected from an individual tree, but are
usually analyzed at the stand level to examine spreading fires. Multiple stands can be sampled in
a watershed to look at fire spread across the landscape. Many such watersheds can be sampled
on a regional basis to understand how climate can drive fire occurrence at this broadest scale
(from Swetnam and Baisan 1996).

319
Figure 9. 9 A fire history chart for a network of 55 site-level chronologies extending back to
A.D. 1600 throughout the southwestern United States (Swetnam et al. 1999). Each horizontal
line in the top part of the graphic represents an individual site while each tic mark on that line
represents a separate fire event that affected many trees on the site. The bottom part of the graph
is a composite of fire charts throughout the western United States and Northern Mexico.

320
around the 1600s, probably for the purpose of improving cattle grazing conditions and for slash-

and-burn agriculture (Lehtonen and Huttunen 1997, Groven and Niklasson 2005). Fires tend to

cease around the late 1700s because of increased value of timber and a likely resultant attention

to conservation of this resource (Groven and Niklasson 2005).

Fire in Canada. Considerable fire history research has been done in Canada from the mixed

hardwood forests of south-central Ontario (Dey and Guyette 2000), to the boreal forest

(Bergeron 1991) and even at the latitudinal treeline at 58° North latitude (Payette et al. 1989).

Bergeron (1991) examined fire occurrence on mainland sites versus island sites to determine the

driving factors that control fire frequency. He documented more frequent/less intense fires on

the red pine (Pinus resinosa) and common juniper (Juniperus communis) dominated rocky

islands because of fuel limitation, in contrast to less frequent/stand replacing fires on the

mainland that was dominated by balsam fir (Abies balsamea), black spruce (Picea mariana), and

paper birch (Betula papyrifera) on a site that was more moist and had greater fuel accumulation.

A decrease in fire frequency over the past 120 years was found in this study and attributed to a

decrease in long-term droughts associated with warming since the Little Ice Age (Bergeron 1991,

Bergeron and Archambault 1993). Studies from 55° to 59° North latitude have examined the

variability of fire rotation periods across biomes from boreal forest through the forest tundra and

into the shrub tundra (Payette et al. 1989). It was estimated that the northern boreal forest fire

rotation period was 100 years while shrub tundra fire rotation period was greater than 7,800

years.

321
Conclusions from Dendropyrochronology. Fire history has been one of the main tools of

dendroecologists, supplying much information about disturbance ecology over the past 30 years.

While working with this application, researchers have perfected techniques for seasonal

resolution of dating scars preserved in the trees and have made advances in spatial as well as

temporal analysis. In the next section, the study of insect outbreaks adds to our knowledge of

disturbance ecology.

Dendroentomology

The study of insect outbreaks has become a major subfield of study in dendrochronology because

forest managers are interested in the historical effects of insects on their managed lands.

Dendroentomology documents past occurrence of insect outbreaks and gives an understanding of

insect population dynamics including duration of outbreaks, interval between outbreaks, and the

spread of insect outbreaks (Swetnam et al. 1985). As with all dendrochronological applications,

dendroentomology provides a long-term perspective for ecological dynamics in a forest system.

The earliest study of insect outbreaks using tree rings was conducted by a German botanist,

Ratzeburg (1866), who dated outbreaks of a defoliating caterpillar with annual resolution

(Ratzeburger 1866, as cited in Wimmer 2001 and Studhalter 1955). In an introductory textbook

on forestry, Hough (1882) shows a graphic of the reduced growth of tree rings related to the

defoliation of insects in the eastern U.S. The presence of this graphic and statement in a forestry

textbook in the late 1800s demonstrates the general knowledge of tree growth, the ability to date

ecological phenomena with tree rings, and the effect of insects on trees and their growth. The

field of insect outbreak reconstructions started in earnest in the 1950s and 1960s when a series of

322
publications made this discipline more accessible to researchers (Blais 1954, 1957, 1958a,

1958b, 1961, 1962, 1965, Hildahl and Reeks 1960). Blais (1958a) documented a decrease in

ring width of balsam fir and white spruce due to the effects of eastern spruce budworm

(Choristoneura fumiferana). In 1960, Hildahl and Reeks published a study on the effect of forest

tent caterpillar (Malacosoma disstria) on trembling aspen in Manitoba and Saskatchewan. Blais

(1962) did much to establish the techniques for studying insect outbreak dynamics in his study of

eastern spruce budworm in Canada. Long-term reconstructions covering the past 200-300 years

have demonstrated that spruce budworm has increased in frequency, extent, and severity caused

by human changes in the forest ecosystems (Blais 1983). Swetnam et al. (1985) published a

manual on how to approach insect outbreak studies that became a standard in the field. This

publication helped to codify an approach to insect outbreak reconstruction that was quickly

followed from the 1990s to the present with a flurry of dendrochronological insect outbreak

publications. Canada is one of the more active areas in insect outbreak reconstruction with work

by Krause and Morin (1999), Zhang and Alfaro (2002), and Campbell et al. (2007).

Insect outbreak studies come in many different forms depending upon how the insects affect the

trees; insects may be defoliators, cambium feeders, or root parasites. The defoliators focus on a

type of tree and consume leaves or needles from those tree species. Examples of these types of

insects include western spruce budworm (Choristoneura occidentalis) (Swetnam and Lynch

1993), Douglas-fir tussock moth (Orgyia pseudotsugata) (Swetnam et al. 1995, Mason et al.

1997), and pandora moth (Coloradia pandora) (Speer et al. 2001) (see Table 9.1 for a more

comprehensive list). Insects that feed on the cambium, usually killing the tree, include bark

beetle larvae (Dendroctonous and Ips species) (Eisenhart and Veblen 2000). Finally, an example

323
Table 9. 1 Insects that have been studied using dendrochronology.

Common Name Scientific Name Type Publication


Spruce Beetle Dendroctonos rufipennis Cambium feeder Eisenhart and Veblen
2000
Mountain pine beetle Dendroctonos ponderosae Cambium feeder Shore et al. 2006,
Campbell et al. 2007
Forest tent caterpillar Malacosoma disstria Defoliator Hildahl and Reeks 1960
Gypsy Moth Lymantria dispar Defoliator Asshof et al. 1999
Larch Sawfly Pristiphora erichsonii Defoliator Case and MacDonald
2003
Pandora Moth Coloradia pandora Defoliator Speer et al. 2001
Western Spruce Choristoneura Defoliator Blais 1962
Budworm Swetnam and Lynch
occidentalis
1993
Tussock Moth Orgyia sp. Defoliator Mason et al. 1997
Two-Year Spruce Choristoneura biennis Defoliator Zhang and Alfaro 2002
Budworm
Periodical cicadas Magicicada sp. Root parasite Speer unpublished data

324
of a root parasite is the periodical cicada (Magicicada sp.), a group of insects that are restricted

to the eastern United States and spend 99 percent of their life cycle feeding on the xylem fluid in

the roots of trees. Recent research has demonstrated that they do not greatly affect the trees but

may cause a reduction in growth when they oviposit in the branches of the trees. They may also

increase the growth of the trees by providing a nutrient pulse when their carcasses decompose

after a massive emergence in the eastern United States (Speer unpublished data). Researchers

can also document the spread of invasive species such as the hemlock wooly adelgid (Adelges

tsugae), gypsy moth (Lymantria dispar), and the emerald ash borer (Agrilus planipennis) and use

the techniques of dendroentomology to reconstruct the effects of fungus on tree populations such

as the chestnut blight (Cryphonectria parasitica) and the Dothistroma needle blight

(Dothistroma septosporum) (Welsh 2007). Insects that cause mortality of the host trees, such as

wooly adelgid, emerald ash borer, and many bark beetles, are more difficult to reconstruct

because a mortality event could be caused by many different factors and that ends the record for

those particular trees. Repeated outbreaks of these insects may be recorded by the response of

other trees in the stand that are released by the mortality of the host species.

From the different groups, dendroentochronologists reconstruct a greater variety of defoliating

insects than any other type. Calibration of the tree-ring record with historical documentation of

past insect outbreaks has been very helpful in determining the insect's effects on the trees

(Brubaker and Greene 1979, Swetnam et al. 1985, Wickman et al. 1994). Comparisons of ring

patterns during periods of known outbreaks in a study area can identify a tree-ring signature

specific to that insect species. By examining the defoliation effects of tussock moth and spruce

budworm in the same tree, researchers have, in some cases, been able to differentiate between

325
the signature of the two species and subsequently document past outbreaks of each species

(Brubaker and Greene 1979, Wickman et al. 1994, Mason and Torgerson 1987). Douglas-fir

tussock moth produced a four to five-year signature of sharply reduced growth, while the

western spruce budworm entailed more gradual but longer outbreak periods (often 10 years),

leaving a signature of less abrupt but more persistent growth reduction (Wickman 1963,

Brubaker and Greene 1979). The differentiation of outbreak patterns aptly demonstrated the

effectiveness of dendrochronology in identifying specific ring signatures for phytophagous

(leaf-eating) insects throughout the length of the tree-ring chronologies. However, when western

spruce budworm and Douglas-fir tussock moth outbreaks occur simultaneously or closely spaced

in time, they cannot always be differentiated in trees or stands that were defoliated in the past by

both species (Swetnam et al. 1995).

Reduction in tree growth reflects the period when defoliation significantly impacts tree health

and does not usually begin precisely with the onset of the insect population’s increased growth

(Swetnam and Lynch 1993). Stored food reserves can delay defoliation-induced growth loss by

one or more growing seasons (O’Niell 1963, Kulman 1971, Brubaker and Greene 1979). Since a

tree requires time to replace lost foliage following severe defoliation, its growth may be inhibited

for several years after the insect populations have crashed (Duff and Nolan 1953, Mott et al.

1957, Wickman 1963, Brubaker and Greene 1979, Alfaro et al. 1985, Lynch and Swetnam

1992).

Researchers have developed techniques for differentiating climate-related ring-width

suppressions in the host trees from those produced by insect outbreaks (Wickman 1963, Koerber

326
and Wickman 1970, Brubaker and Greene 1979, Swetnam et al. 1985). Climate subtraction

techniques have been developed and widely-tested in studies of the western spruce budworm

(Brubaker and Greene 1979, Swetnam et al. 1985, Swetnam and Lynch 1993, Wickman et al.

1994, Swetnam et al. 1995, Weber and Schweingruber 1995). In the Swetnam et al. (1985)

approach, a non-host "control" tree species is collected from an adjacent site as the host species

and its tree-ring chronology compared to the host series. The common climate signal can be then

subtracted from the host chronology, thereby isolating the species-specific factors for further

study. However, some error or noise may be introduced into the analysis due to differing

responses to climate between the host and non-host tree species (Swetnam et al. 1985).

Speer et al. (2001) developed a record of pandora moth outbreaks that extends back 622 years in

south-central Oregon (Figure 9.10). Pandora moth is a phytophagous insect that defoliates

ponderosa pine, Jeffrey pine, and lodgepole pine in the western United States (Figure 9.11). The

Klamath and Piute Indians used the pandora moth larvae and pupae as a traditional food source

when it was available, indicating they had knowledge of its life cycle (Blake and Wagner 1987).

This led early forest entomologists to speculate that pandora moth outbreaks had often occurred

in the past (Aldrich 1912, 1921, Patterson 1929). Pandora moth and ponderosa pine trees are

well adapted to each other so that only two percent tree mortality occurs with the outbreaks

(Patterson 1929, Massey 1940, Bennett et al. 1987). Speer et al. (2001) were able to identify a

distinct ring width pattern or signature that is associated with an outbreak of this insect in

ponderosa pine forests (Figure 9.12). This signature was calibrated from sites with historically

documented outbreaks and was applied to reconstructing outbreaks in the past. Further analysis

327
 

Figure 9. 10 A 622-year pandora moth reconstruction from south-central Oregon. The top line
represents the number of trees recording outbreaks through time. The sample depth decreases
further back in time with 20 trees still recording outbreaks at A.D. 1500. The dark area shows
the number of trees recording outbreaks through time throughout the entire south-central region
of Oregon (from Speer et al. 2001).

328
Figure 9. 11 A ponderosa pine forest denuded of needles by pandora moth (from Speer et al.
2001).

329
Figure 9. 12 A tree ring signature has been identified for pandora moth in which the first year is
half the size of normal, the next two years are the smallest in the series, and subsequent years
gradually return to normal growth. Thin latewood throughout the outbreak is another
characteristic of the signal. Here, the upper photo shows a tree affected by a documented
pandora moth outbreak in the 1960s which was used as a calibration for inferred pandora moth
outbreaks with the signature starting in A.D. 1661, as seen in the bottom photo (from Speer
1997).

330
demonstrated that pandora moth was recorded on multiple trees within a site and between sites

(Figure 9.13).

Mountain pine beetle (Dendroctonus ponderosae) has become a major influence in the western

United States and Canada, with an outbreak from 1999 to the present affecting more than seven

million hectares in Canada alone (Taylor et al. 2006). Bark beetle outbreaks can be

reconstructed by documenting the mortality of host trees and occasionally from scars that are

preserved on trees that live through the outbreak. Tree death is recorded indirectly as release in

trees that survive the outbreak (Taylor et al. 2006). These types of reconstructions are more

difficult because a distinct signature for cambium feeding insects does not exist like it does for

most defoliating insects. The economic impact of the recent mountain pine beetle in Colorado

and neighboring states and the large outbreak in British Columbia has brought more attention on

this insect to try to understand if these events are natural or triggered by other factors such as

forest management practices and/or climate change.

Stem Analysis. Potential wood volume increases of a forest stand can be reduced during insect

outbreaks either through mortality of the host trees or suppression of radial growth and this

reduction has implications for forest management policies. For example, while pandora moth

outbreaks typically cause almost no loss due to mortality, the amount of volume reduction due to

the effects of defoliation can be quite substantial (Massey 1940, Wickman 1963, Koerber and

Wickman 1970, Speer and Holmes 2004). Growth loss during outbreaks may be offset by a

growth increase after the insect population has crashed, a phenomenon observed with spruce

budworm and Douglas-fir tussock moth outbreaks (Wickman 1980, Alfaro et al. 1985, Swetnam

331
 

Figure 9. 13 Insect outbreaks often affect many trees both on an individual site and on multiple
sites. In this graphic, each horizontal line is an individual tree with its ring width index plotted
through time, and two sites (pf and ef) are represented. Dashes at the bottom of the graphic
indicate known (K) outbreaks and inferred (I) outbreaks. Each outbreak event is recorded not
only by all trees from the same site, but by trees on different sites, demonstrating that pandora
moth outbreaks occurred on a broad scale and affected multiple sites (from Speer 1997).

332
and Lynch 1993). This effect may be attributable to factors such as reduced competition among

trees for resources and nutrient cycling associated with frass (excrement) accumulation.

Stem analysis has been used extensively to investigate the effects of insect defoliation on growth

allocation (Figure 9.14). The standard technique, with refinements by Duff and Nolan (1953 and

1957), involves taking multiple cross sections along the stem of the tree. The ring widths are

then measured from each cross section and used to estimate three dimensional growth throughout

the entire tree (Figure 9.15). Duff and Nolan (1957) and LeBlanc (1990a) suggested sampling a

section midway between each internode to allow for quantification of height and radial growth in

every year. Yet for trees a few centuries in age, it is very difficult or impossible to identify

internodes on the external surfaces of the main stem. Thus, with increasing age it becomes

impractical to determine all of the annual height increments; however cross sections can be taken

at regular intervals along the trunk instead (Figure 9.15). LeBlanc et al. (1987) note that stem

analysis affords increased accuracy in determining the overall tree response to disturbance, but

mentioned the added effort might preclude its widespread application. They also mentioned the

additional difficulty when studying older/larger trees because of the obscurity of the internodes

as trees age. The effort of conducting a stem analysis is worthwhile, however, if the researcher

wants to visualize the changes in wood volume due to defoliation and loss of photosynthetic

potential (Figure 9.16).

Conclusion of Dendroentomology. The tools developed from dendroentomology are now being

used for other forest health agents such as fungal pathogens (Welsh 2007) and to address

complex disturbance systems involving multiple agents (e.g. Thompson 2005). These concepts

333
 

Figure 9. 14 Here, the author is taking samples from a 600 year old ponderosa pine tree for a
stem analysis to examine the wood volume lost due to pandora moth defoliation. This is also an
example of opportunistic sampling, because this tree was killed by a winter storm in 1993,
enabling easy sampling in 1996 without having to cut down a living tree (photo taken by Tom
Swetnam).

334
Figure 9. 15 Diagram showing how samples taken every 3 meters up a tree can be used to
calculate wood volume for the whole tree (from Speer and Holmes 2004).

335
Figure 9. 16 Examples of four trees showing changes in wood volume with height of the tree.
The dates and arrows along the x-axis show where pandora moth outbreaks resulted in a decrease
in wood volume (from Speer and Holmes 2004).

336
take advantage of the aggregate tree growth model, limiting factors, site selection, and

replication, just to name a few of the main principles of dendrochronology that are applied and

honed through insect outbreak studies. These techniques continue to grow as researchers expand

to new insect systems such as periodical cicadas and mountain pine beetle. The management

concerns of foresters force the research agendas of many scientists as we react to public concern

and governmental interest.

Wildlife Populations and Herbivory

Dendrochronology can be used to determine the dates of herbivory on trees, to estimate wildlife

populations through resource availability linkages, and even to study fishery populations based

on covariance with sea surface temperature measures (Spencer 1964, Schweingruber 1996, Speer

2001, Drake et al. 2002). Past fluctuations in animal populations can be documented through

scars left on trees such as those from porcupine (Erethizon epixanthum) feeding on small

branches of pinyon pine in Mesa Verde, Colorado (Spencer 1964) or expanding into northern

treeline in Quebec (Payette 1987). A study on Isle Royale National Park in Michigan

demonstrated how the removal of wolves (Canis lupus) resulted in an increase in moose (Alces

alces) populations which then overgrazed balsam fir trees (McLaren and Peterson 1994). Hessl

and Graumlich (2002) examined the effects of an elk herd on aspen regeneration in Wyoming,

finding high elk populations reduce the recruitment of aspen trees. Speer (2001) was able to

reconstruct masting and found a significant correlation between the regional white oak mast

reconstruction in the southern Appalachians and black bear population estimates. Wildlife

population reconstructions and examination of ecological interactions help zoologists study

wildlife population fluctuations and increase understanding of complex food web interactions.

337
The long-term perspective of dendrochronology provides the time depth needed to examine the

behavior of animals over multiple generations.

Dendrochronology can also provide valuable information for the management of fisheries.

Drake et al. (2002) documented a significant correlation (r2= 0.23, p < 0.05) between Sitka

spruce tree growth and Northeast Pacific salmon stocks, suggesting that both tree and fish

responded to, or were tracking, the same environmental variable such as sea-surface temperature.

They were able to take this relationship further by reconstructing salmon populations which were

verified against other known salmon stock data. Clark et al. (1975) documented fluctuations in

albacore tuna (Thunnus alalunga) populations by examining tree growth response to broad-scale

atmospheric flow patterns that affected sea surface temperature in the north Pacific, influencing

both tree growth and tuna populations. This technique of examining synoptic climate linkages

between terrestrial records from trees and sea surface temperature or circulation patterns has

enabled dendrochronologists to shed light on broad scale circulation and temperature phenomena

that also affect fish populations. This type of analysis effectively extends the usable range of

dendrochronological reconstructions and, in these examples, aids fishery managers.

Distributional Limits of Species

Biogeographers have long been interested in the factors that control the range limits of different

tree species. Range limits may also mark the ecotone boundaries between different biomes. An

ecotone is a zone of change from one vegetation type to another that often results in high plant

and animal diversity. Ecotones are interesting to study because they are the first place that will

demonstrate a response to climate change and because of their inherently high level of

338
biodiversity. Some of the main variables controlling tree species distribution seem to be

temperature, precipitation, and disturbance. In a study at the northern range limit of red pine,

Bergeron and Brisson (1990) found that the species is restricted to island sites that have more

frequent fire (a disturbance process), which may limit the competitive ability of boreal forest

species, enabling red pine to maintain itself at 48° North latitude. Conkey et al. (1995)

documented growth of jack pine in a marginal location at the eastern edge of its range limit in

Acadia National Park, Maine. Jack pine is an early successional species that is fire adapted, but

was able to survive in a marginal habitat because of the thin soil that prevented later successional

species from out-competing it (Conkey et al. 1995). Copenheaver et al. (2004) studied a

prominent ecotone at Buffalo Mountain, Virginia between dense forest and mountain top balds

(grassy openings on a mountain top that are not above tree line). They used transects and stand

age structure analyses, founding that some of these balds are stable while others are being

invaded by the surrounding forests. In a study across an elevational gradient in northern

Arizona, Fritts et al. (1965) documented that trees growing on the semiarid lower forest border

were more sensitive to climatic variability. This lower treeline is controlled by a lack of soil

moisture and produces sensitive tree-ring chronologies. Fine-scale analyses of the factors that

control a species’ range help biogeographers better understand the distribution of species. This

understanding of the controlling mechanisms also helps scientists predict future range limits

under a changing environment.

Treeline and Subarctic Studies

High elevation treeline (the elevational limit to which trees can grow because of temperature or

moisture limitations) is another interesting ecotone where mountain-grown trees may take on a

shrub-like form, called krumholz. Krumholz growth can occur in response to harsh winter

339
weather where conditions under the snow pack are more conducive to survival over the winter.

Historically, treeline has been studied as a proxy for temperature fluctuations, with the

understanding that elevational treeline is limited by cold temperatures (LaMarche and Mooney

1967, LaMarche 1973, LaMarche and Stockton 1974). Daniels and Veblen (2003) found that

precipitation controlled the local elevation of treeline in northern Patagonia and that disturbance

locally lowered treeline. They also documented that treeline in Chile and Argentina was limited

by lower precipitation as well as low temperatures and predicted that treeline advance may be

restricted in a warming climate because of a lack of precipitation (Daniels and Veblen 2003,

Daniels and Veblen 2004). Lloyd and Graumlich (1997) also found that treeline is dependent

upon moisture as well as temperature, suggesting that treeline response to future warming may

depend heavily on water supply. Similar studies have been conducted in high latitudes at the

transition to artic tundra. Au and Tardif (2007) examined tree rings in the shrub dryas (Dryas

integrifolia) in subarctic Manitoba, Canada. They found that these shrubs produced datable

annual rings and could demonstrate that the shrubs were sensitive to previous October

precipitiation and current year’s May temperature. As dendrochronological methods advance,

we can now begin to see the complex interactions of the effect of climate as well as disturbances

in controlling factors of altitudinal and latitudinal treeline.

Interactions of Multiple Disturbances

Dendrochronologists have separately studied disturbances such as fire, insect outbreaks, blown

down trees, avalanches, and herbivory. In the last fifteen years, researchers have started to

examine the interaction of multiple disturbances on the same site, giving managers a more

complete idea of the how processes affect one another on a given landscape (Hadley 1994,

340
Veblen et al. 1994, Kulakowski and Veblen 2002, Kulakowski et al. 2003, Thompson 2005).

Veblen et al. (1994) found that snow avalanches create fire breaks resulting in smaller fires,

while fires and avalanches kill mature trees which delays the onset of bark beetle outbreaks

because the remaining trees were not large enough for selection by bark beetles. Kulakowski et

al. (2003) found that spruce beetle outbreaks in Colorado also affected the fire regime resulting

in fewer occurrences of surface fires. They suggested a mechanism of the bark beetle outbreaks

resulting in higher moisture on the forest floor because Reid (1989) observed a proliferation of

mesic understory herbs. Similar observations have been made after mountain pine beetle

outbreaks in Bristish Columbia where an increase in soil moisture was documented due to a

decrease in transpiration (Kathy Lewis personal communication). Kulakowski et al.’s (2003)

finding of fewer surface fires following bark beetle infestation was contrary to previous research

that suggested bark beetle outbreaks resulted in an increase in forest fires due to an increase in

available fuels (Stuart et al. 1989, McCullough et al. 1998). Build up of fuels usually occurs

after stand-replacing fires, so a difference in the scale of the effect might result in these opposite

conclusions. Kulakowski et al. (2003) state, however, that their observation also holds for stand-

replacing fires. The time since defoliation can also be an issue in fire occurance, where

extensive needle fall occurs a few years after defoliation which could increase the spread of fire.

Those fine fuels decompose quickly and the stand is left with much standing fuel, but may not

have the fine fuel necessary to carry a fire, resulting in a decrease in fire occurance.

Controversial issues such as these remain to be examined with further dendroecological research.

Researchers have found that natural disturbances regenerate forests and make them less

susceptible to subsequent disturbances until the forest matures (Kulakowski and Veblen 2002).

341
After the 1997 blowdown event in the Routt National Forest, Colorado, which took down over

10,000 hectares of subalpine forest, Kulakowski and Veblen (2002) found that forests which had

experienced stand-replacing fire within the last 120 years were less susceptible to wind damage

because of their vigor. Anderson et al. (1987) found that fire suppression in a ponderosa pine

forest in western Montana enabled Douglas-fir to proliferate over the last 100 years providing

more host trees which resulted in an increase in duration and intensity of western spruce

budworm outbreaks. As a way to control the outbreaks, they suggest that fire be reintroduced to

maintain the ponderosa pine forest and reduce Douglas-fir density, bringing this forest back into

its historical condition. Only after managers understand the complex interactions among natural

disturbances, will they be able to manage their forests tracts in an ecologically sustainable

fashion.

Other Applications in Dendroecology

Schweingruber (1996) has developed a program of ecological examination in Switzerland that

focuses on the growth and adaptation of plants to environmental stressors. He is also expanding

his work to tropical environments and tundra environments by studying wood anatomy and rings

in perennial herbs and shrubs. Tree rings have been used to determine the factors that drive wet

heartwood occurrence in forest trees, which is heartwood that has unusually high water content.

Krause and Gagnon (2006) found trees growing in an area of a high water table were more likely

to have wet heartwood as well as suppressed growth due to the stress of growing in a frequently

saturated soil. Root age can be determined from tree-ring analysis, which provides an

understanding of how roots develop in different tree species (Krause and Morin 2005). Studies

of black spruce and balsam fir in Quebec demonstrated that adventitious roots grew more than

342
60% of their length in the first year of development while lateral roots produced 93% of their

elongation in their first 10 years (Krause and Morin 2005). This work is useful for our

understanding of how roots grow and suggests that the major structure of a root develops fairly

quickly. Other research in root wood anatomy will be presented in the chapter on

dendrogeomorphology (Chapter 10) because this can be a useful tool in examining soil erosion.

Maximum latewood density of annual rings has been found to correlate with Normalized-

Difference Vegetation Indices (NDVI) which are an estimate of vegetation productivity or net

primary productivity (NPP) from satellite imagery (Malmstrom et al. 1997, D’Arrigo et al.

2000). This correlation is useful for broad scale estimates of forest productivity and change

associated with global climate change, which is an area likely to take on greater importance in

the presence of global warming. This work demonstrates that interesting frontiers still exist to be

studied with dendrochronology. More of these new directions of research will be discussed in

Chapter 12.

Conclusion

Dendroecology has been and is becoming more useful for exploring a wide range of research

topics that can provide important information to wildlife, fisheries, and forest resource managers.

Combining the study of tree rings and ecology can help us understand the dynamics of natural

processes such as disturbance and the interactions between multiple natural phenomena. Trees

can provide long-term records on many different phenomena at different spatial scales, enabling

dendroecologists to contribute to important discussions on scaling laws that could aid

management in a changing environment.

343
Chapter 10: Dendrogeomorphology

Introduction

Geomorphology is the study of landforms and the earth surface processes that form and modify

them (Gärtner 2007a). Dendrogeomorphology uses tree rings to date geological processes that

affect tree growth such as landslides, river deposits, or glacial activity.. I consider

dendrogeomorphology to include the subfields of of dendroglaciology (the study of the

movement or mass balance of glaciers), dendrovolcanology (the study of past volcanic

eruptions), dendrohydrology (the study of stream dynamics), and dendrosiesmology (the study

of past earthquake events and fault movements through the use of tree rings) (Table 10.1).

The use of dendrochronology to reconstruct geological phenomena was pioneered in North

America over 100 years ago. Sherzer (1905) used the growth of spruce trees to estimate the ages

of glacial moraines in the Canadian Rockies and Selkirk Mountains in Canada as one of the first

dendrogeomorphological applications. In another early application, dendrochronology helped

resolve a boundary line dispute between Texas and Oklahoma (Sellards et al. 1923). The Red

River marked the boundary between the two states, but because its stream channel meandered

over time, the relative location of the state line changed. By examining the age of trees on

different land surfaces, the researchers were able to determine the location of the historical

boundary. Dendrogeomorphology really became established as a subfield in the United States in

the 1970s and its applications have expanded since this time (Alestalo 1971, Shroder 1978, 1980,

Shroder and Butler 1987, Butler 1987, Schweingruber 1996, Wiles et al. 1996).

344
Table 10. 1 Geomorphic events and how they can be reconstructed using tree rings. Table
modified from Shroder 1978, Sheppard and Jacoby 1989, and Wiles et al. 1996.
Process Event Possible responses Citation

Volcanic loss of photosynthesis suppression Smiley 1958, Yamaguchi 1983


lava flow and temperature
stress suppression

atmospheric cooling suppression LaMarche and Hirschboeck 1984


crown defoliation from
tephra fallout suppression Smiley 1958, Yamaguchi 1983
eluviation of leachates
harmful/favorable to
growth suppression/release
reduced aeration of soils
due to burial suppression

direct encounter with flow scarring

denudation of surface tree establishment Yamaguchi and Hoblitt 1995

gas release suppression

Flodplain debris impact during Gottsfeld and Gottsfeld 1990,


dynamics floods scarring McCord 1996

accretion of point bars tree establishment Gottsfeld and Gottsfeld 1990

erosion of banks tilting/mortality Gottsfeld and Gottsfeld 1990

inundation by sediment mortality and burial Sigafoos 1964

flow dynamics ring width variability Woodhouse 2001

Lake ice
dynamics direct ice push scarring Begin 2000

physical impact from Lawrence 1950,


Glacial glacial ice scarring Luckman 1988
inundation by glacial
sediment suppression/sprouting Wiles et al. 1999
temperature stress from Lawrence 1950,
proximity of ice suppression Wiles et al. 1996

345
missing rings/mortality/ Smith and Laroque 1996,
advance/retreat establishment Wiles et al. 1999
Sigafoos and Hendricks 1961,
denudation of surface tree establishment 1972, Wiles et al. 1999

mass balance change suppression/release Laroque and Smith 2005

isostatic adjustment tree encroachment Begin et al. 1993

Mass Corominas and Moya 1999


movement inclination reaction wood Fantucci and Sorriso-Valvo 1999

shear suppression Shroder 1978


Corominas and Moya 1999,
corrasion scarring Shroder 1978
suppression/root mortality,
change in cell thickness of
exposure of roots roots Danzer 1996, Gärtner et al. 2001
suppression/sprouting/ Hupp 1984, Begin and Filion
inundation mortality 1988, Corominas and Moya 1999
Hupp 1984,
denudation of surface tree establishment Hupp et al. 1987

change in hydrology release/suppression Fantucci and Sorriso-Valvo 1999

Earthquake shear suppression/missing rings Jacoby et al. 1988

inclination reaction wood Sheppard and Jacoby 1989


Atwater and Yamaguchi 1991,
change in water table suppression/release Sheppard and Jacoby 1989

extreme ground shaking suppression/missing rings Sheppard and Jacoby 1989

broken tree tops suppression/mortality Jacoby 1997


root system or major limb
damage suppression Jacoby 1997, Bekker 2004

346
Shroder (1980:165) outlined a process-event-response system for the analysis of

dendrogemorphological phenomena. He identified seven basic events:

1) inclination,

2) shear of rootwood or stemwood,

3) corrasion (which is abrasion or some removal of wood through contact),

4) burial of stemwood,

5) exposure of rootwood,

6) inundation, and

7) denudation (or the removal of vegetation).

In response to these events he categorized seven responses (Shroder 1980: 165):

1) reaction wood growth,

2) growth suppression,

3) growth release,

4) ring termination and new callous growth,

5) sprouting,

6) succession, or

7) miscellaneous structural or morphological changes in external or internal wood

character

This systemic approach is still useful today for dating possible events and responses of trees to

geomorphic phenomena.

347
Most tree-ring sampling for geomorphological research needs to involve targeted sampling in

which the direct effects of landslides, earthquakes, glaciers, or soil creep can be identified.

Because of the variety of phenomena that can cause reaction wood or scarring in a tree, samples

must be selected from areas that have been affected by the process of interest (Shroder 1980,

Butler 1987). Targeted or directed sampling uses the basic dendrochronological principle of site

selection discussed in Chapter 2. Random sampling on the landscape is likely to miss these

geomorphological events or require a huge amount of sampling to detect such localized events.

Sources of Information

Reaction wood

Trees will react structurally to being tilted by producing reaction wood. Gymnosperms

(conifers) will thicken cell walls and produce more cells on the downhill side of the tree while

angiosperms (flowering trees such as the hardwoods) will thicken the tracheids on the uphill side

of the tree (called tension wood), both in an attempt to straighten the tree (Figure 4.18). This

reaction wood can be used to determine the timing of events that tilted the tree trunk, such as

landslides or earthquakes (Gärtner 2007a). Changes in the circularity of stem growth can record

geomorphic changes through time. To sample for these events, a full cross section of the stem of

the tree is preferred, although one can core on the downhill and uphill side of the tree through the

reaction wood in an attempt to date such a tilting event. Note that this coring location is contrary

to previous sampling protocols discussed in Chapter 5 as it is targeting reaction wood, whereas

such studies as climate reconstruction try to avoid this irregular growth. By examining the

reaction wood of trees in an area of the Eastern Pyrenees in Spain that experienced frequent

landslides, researchers were able to reconstruct landslide activity there (Corominas and Moya

348
1999). The reaction wood documented slope instability over the past 70 years, revealing that

landslide activity had increased in modern times compared to the early portion of the chronology

from 1926-1959.

Death dates

Death dates can be obtained by crossdating dead wood samples against a living tree-ring

chronology to determine the advance of a glacier, when landslides occurred, or any other natural

event that results in the death of a tree. Preservation of the sample and its outermost rings then

determines how far back in time one can crossdate the event that caused tree mortality. Factors

that contribute to the quality of wood preservation include climatic conditions (hot and humid or

dry and cold for example), where the tree is located (whether the tree is a standing snag, sitting

on soil, suspended in the air, buried in sediment, or buried in a lake in anoxic conditions), and

innate resistance of the wood itself to weathering and decay (sequoia and cedar wood for

example).

Death dates can also be used to determine the sedimentation rates on a slope (Figure 10.1).

When a tree dies and falls across a slope, it will remain there for some period of time catching

sediment that comes down slope in overland flow of water. The sedimentation rate can be

determined from the amount of sediment present and the time since death of the log (Hart 2002).

Various factors complicate this process. If the tree died and stayed standing for 10 years, and

then fell to become a sediment trap, the accumulation rate would be underestimated because of

the 10 years that the tree remained standing.

349
 

Figure 10. 1 Coarse woody debris is composed of logs that fall in the forest. They may fall
across a slope and act to catch sediment during overland flow of water. These trees are very
important for sediment retention. If death dates can be established in the trees and the amount of
sediment can be measured that accumulates behind the log, then sedimentation rates can be
determined (photo from LaMarche 1968).

350
Establishment dates

Ecesis is the process of vegetation becoming established on previously bare ground that was

denuded by flooding or glacial activity (McCarthy and Luckman 1993). Establishment dates of

trees can be used to provide a bounding date on when that material was deposited or wiped clean,

but there is often a lag (from ecesis) between the time when the event occurred and when trees

first establish on the site. This lag can depend upon seed source, suitability of the substrate, or

climate. Estimates for this time period can be made on local sites of known disturbance to

calibrate a local record. When determining the bounding date for a surface and estimating time

of ecesis, one assumes that the oldest tree on that surface has been sampled (Wiles et al. 1996).

The techniques used for this process are the same as those for a stand-age structure employed by

dendroecologists (see Chapter 9). The goal is to document a cohort of tree establishment on a

surface that was cleared or deposited by some geomorphic agent such as a volcanic eruption,

landslide, or debris flow. One of the main problems with determining the age of a land surface

with tree rings is the lag time between when the surface formed and how long it took trees to

establish on the site (successional dynamics). This lag time can be driven by climate, the biology

of the trees, distance to a seed source, and presence of seed dispersers (Fastie 1995). All of these

factors combine, leaving some doubt as to the exact age of the surface, but establishment dates

do provide a bounding date of the earliest possible time that the surface could have been formed.

Wound Events

Trees can be damaged during geomorphic events and these wounds can be dated to reconstruct

rock falls or other damaging occurrences. For example, debris flow events were reconstructed

near Valais region of the Swiss Alps using tree scars and eccentric growth from trees being

351
dislodged in past debris flows (Stoffel et al. 2005). They were able to extend the debris flow

records from 80 years of historical data to 397 years from tree-ring records and found that the

peak of debris flow occurrence was in the 1800s. Scars can be caused by many sources, so

location of the scar and clustering of scar events is important for documenting past geomorphic

phenomena. Other possible geomorphic causes of tree scars include landslides, avalanches,

flood waters, and ice flows.

Coarse Woody Debris (CWD)

Course woody debris (CWD) is an important component of the dead wood in any forest because

it acts as a sediment trap, nutrient source, and increases habitat (Figure 10.1, Daniels et al. 1997,

Hart 2003, Campbell and Laroque 2005, Campbell and Laroque 2007). Foresters and

geomorphologists have defined stages of wood decay from recent (decay class I) to old (decay

class VI). In a dendrochronological study on the southwestern coast of British Columbia,

researchers successfully calibrated the decay classes of cedar by determining the time since death

of the tree (Daniels et al. 1997) enabling foresters to more accurately assess how long logs had

remained on the ground as a sediment trap. It was also noted that no classification system was

appropriate for snags because of their slower decay rate while they are standing above the

ground surface. Gore et al. (1985) developed a model to estimate the maximum likelihood

estimate of the average number of years that a bole would stand before it fell based on empirical

evidence; however their model does not take tree species or site conditions into consideration.

An understanding of CWD and its dynamics through time help in stream restoration and provides

information on habitat changes in riparian areas.

352
Roots

Root analysis can be used as part of a whole-tree analysis as demonstrated in the study conducted

by Krause and Eckstein (1993) who found that root increment was significantly correlated with

temperature. Gärtner (2007b) discusses the usefulness and difficulties of using roots to

determine rate of soil erosion, deposition, and other damages to trees through geomorphic agents.

He notes that little work to date has been successful in the use of roots, but current work with

anatomical features of roots have promise in providing information in the future. Gärtner et al.

(2001) and Gärtner (2007b) have demonstrated different anatomical characteristics for roots at

different depths in the soil (Figure 10.2) and have shown that this change in root wood anatomy

can be used to determine the year and even the season of subaerial exposure of roots through

erosion.

Subfields of Dendrogeomorphology

Dendrovolcanology

Volcanic eruptions can be documented with tree rings through 1) the direct effect on tree stems

from the shock wave, ash fall, or debris from the eruption, 2) mortality of trees on a site, or 3) a

global cooling event from the injection of gases and aerosols into the atmosphere. A number of

fascinating studies have linked trees to notable volcanic eruptions (Smiley 1958, LaMarche and

Hirschboeck 1984, Yamaguchi and Hoblitt 1995, Briffa et al. 1998, Jacoby et al. 1999).

Yamaguchi (1983) was able to use suppression events in Douglas-fir to document major

eruptions from Mount St. Helens. In a subsequent collaborative effort, Yamaguchi and Hoblitt

(1995) used establishment of trees to determine bounding dates on a series of lava flows since

A.D. 870 and determined that Mount St. Helens has had dormant periods that have lasted from

353
 

Figure 10. 2 Wood structure for roots of Larix decidua at four different depths in the soil (5, 9,
12, and 20 cm; from Gärtner 2003). Note the lack of latewood thickening for roots from deeper
locations in the soil.

354
123 to 600 years during their record. In Arizona, an eruption date of A.D. 1064 was determined

for Sunset Crater (Smiley 1958) by the presence of narrow rings in archaeological wood

collected from nearby Wupatki ruin, an Ancestral Pueblo site occupied from approximately A.D.

900-1275. Large eruptions near the equator can put enough sulfur dioxide and aerosols into the

stratosphere to reduce global temperatures for 1-3 years. If temperatures get cool enough during

the growing season of the trees, then a frost ring can be produced. LaMarche and Hirschboeck

(1984) identified frost rings in the bristlecone pine chronology at 3,000 m elevation in the White

Mountains of California that were related to major volcanic eruptions such as Krakatau in A.D.

1680, Vesuvius in A.D. 1785, and Tambora in A.D. 1815 among others. Jacoby et al. (1999)

documented an extremely low density ring in 1783 in Alaska and northern Canada. They were

able to compare this date to historical accounts of Inuit populations in the area that document a

decrease in population. They also found that the Inuit oral traditions speak of a great disaster

around that time period. This extremely cold summer apparently caused a major die-off in the

Inuit in 1783. Further discussion of volcanic effects on trees and how they can be used to date

events can be found in Schweingruber (1996).

Volcanic events may affect a broader spatial scale than normal climatic fluctuations, as

suggested by the work of LaMarche and Hirschboeck (1984). For example, in a study of ring-

width density chronologies from both North America and Europe, Jones et al. (1995)

documented consistent years of low density in 1601, 1641, 1669, 1699 and 1912. Four of these

years coincide with known volcanic eruptions. Decreased density of tree rings means that

individual cell walls are less lignified, which in the case of volcanic eruptions, could be due to an

increase in aerosols in the atmosphere resulting in cooler temperatures reduces lignification in

355
late summer. Jones et al.’s (1995) study is also a good example of the use of spatial scale to

tease apart different signals in tree-ring chronologies. They only wanted to record extreme

broad-scale events so examined years with low tree-ring density on both the European and North

American continents.

Dendroglaciology

Early work in dendroglaciology set the precedent for the variety of information that can be

obtained from tree-ring studies as they apply to glacial research (Tarr and Martin 1914,

Lawrence 1950, Sigafoos and Hendricks 1961, Sigafoos and Hendricks 1972). Glacial advance

can be documented by dating mortality events of sheared trees that are deposited in outwash till

and glacial retreat can be documented by the establishment of trees on newly exposed glacial till

(Wiles et al. 1999). Between these two techniques, mortality events from glacial advance is

more precise because the death of the trees can happen in a year or less, while it may take

decades for trees to establish on newly exposed rock and till after glacial retreat (McCarthy and

Luckman 1993). Through connections with climatic forcing factors, glacial mass balance

(periods of growth and ablation) can also be reconstructed from tree rings (Mathews 1977,

Laroque and Smith 2005). Trees that grow at the trimline (Figure 10.3) where the ice is directly

next to the trees, can record fine scale glacial fluctuation in time (Lawrence 1950, Wiles et al.

1996). Hard work over the past few decades has produced evidence for many alpine and

continental glacial changes over the past 2,000 years that have been combined into broad scale

interpretation of glacial dynamics (Wiles et al. 2008). Such dendroglacial reconstructions as

these extend many centuries into the past and clarify the mechanisms that drive glacial activity.

For further reading, Smith and Lewis (2007) document the history of the field of

356
Figure 10. 3 A glacier may kill trees as it advances and incorporate those trees in the till.
Massive glacial advances can then be dated from the mortality events of the trees. The glacier
leaves the trees alive above the trimline (the highest position of the glacier up slope on the
canyon walls). Trees at the trimline may survive, but their growth is stunted by the proximity of
the glacier. Their suppressed growth can then be used to determine the ice accumulation of the
glacier while fine scale fluctuations of the glacier can be tracked through time (photo by Jim
Speer).

357
dendroglaciology and outline the various techniques that can be used to glean information on

glacial activity.

Isostatic adjustment (the rise of land) after the retreat of glaciers has been documented using

the downslope establishment of trees towards present day sea level to document the rate of this

uplift (Begin et al. 1993). During the last glacial maximum at 21,000 calendar years ago there

were approximately six kilometers of ice above the Canadian Shield. Once the weight of ice

from a major glacier is removed from the terrain, the land begins to adjust to that lack of weight

and to rise, sitting higher on the mantle. New land surfaces are exposed upon which trees can

establish. In a research project located on the margins of Hudson Bay in Québec, Begin et al.

(1993) dated tree establishment along transects to current sea level parallel to the slope,

documenting the progressive advancement of this lower treeline. A similary pattern was

identified from land adjustment after Little Ice Age glacial retreat in Glacier Bay Alaska where

Motyka (2003) documented 3.2 m of uplift since the late 18th century.

Mass Movement

Tree rings can also be used to examine any mass movement such as rockslides (Figure 10.4),

landslides (Corominas and Moya 1999), rock glaciers (Giardino et al. 1984), debris flow (Hupp

1984, Hupp et al. 1987), or volcanic mudflow (also known as a lahar; Yamaguchi and Hoblitt

1995). Just as with flooding, trees can be scarred by mass movements of earth, they can be

killed, or fresh earth surfaces can be deposited or exposed on which trees can establish (Hupp

1984, Hupp et al. 1987, Corominas and Moya 1999, Fantucci and Sorriso-Valvo 1999). Soil

358
 

Figure 10. 4 Frequent rockfall down a landslide shoot may accumulate a large amount of
sediment. As the trees growing on the slope are killed and fall down slope to be incorporated in
the sediment pile (and in this case the lake) at the base of the slope, the wood is well preserved
and can be sampled to determine the landslide frequency for the area. Landslide debris piles are
also a great source of old deadwood that may have been accumulating for some time and could
result in a very long chronology (photo by Jim Speer).

359
creep can cause curvature of stems, although it can be hard to differentiate from curvature due to

wintertime snow pressure (Shroder 1980). Soil erosion can expose roots causing root mortality

(LaMarche 1968, Danzer 1996) or producing cells with different cell wall thicknesses depending

upon their depth in the soil (Figure 10.2 and 10.5; Gärtner and Schweingruber 2001, Gärtner

2007b).

Dendroseismology: Plate Boundaries, Faults, and Earthquakes

Any earthquake could cause damage to a tree by breaking fine and even large roots for the

locally affected trees. Jacoby (1997) provides a complete review of paleoseismology from tree-

ring analysis, noting that trees can be damaged directly from shaking, elevation changes, and

liquefaction, or indirectly through earthquakes that induce landslides and tsunamis. A massive

earthquake that triggered a tsunami and landslide killed thousands of trees sometime between

A.D. 894 and A.D. 897, and was documented by studying submerged logs from Lake

Washington near Seattle, Washington (Jacoby et al. 1992). This work combined

dendrochronology with 14C dating to determine a window of dates for a floating chronology (a

chronology not anchored in time) composed of trees that had been killed in the same season of

the same year by a tsunami triggered by this earthquake. Atwater and Yamaguchi (1991) also

found evidence for a major coastal event that submerged trees in the Seattle area in A.D. 1700.

Suppression of ring width over several years may be another indicator of seismological events in

addition to tree mortality and damage. For example, an 1887 earthquake in Kazakhstan resulted

in ring-width reduction for four to 15 years in most sampled trees (Yadov and Kulieshius 1992).

360
 

Figure 10. 5 These trees around Yellowstone Lake have been subject to gradual soil erosion;
adventitious roots grew as the soil was slowly removed from the site, enabling many of the trees
to survive. The flare of the roots demonstrates at what level the soil used to be (this lines up with
the modern soil level in the background). The soil is being washed away and the beach cut is
moving inland (photo by Jim Speer).

361
Considerable research has been conducted to study specific types of plate movement associated

with plate tectonics and fault types such as transform plate boundaries (Page 1970, LaMarche

and Wallace 1972, Wallace and LaMarche 1979, Meisling and Sieh 1980, Jacoby et al. 1988,

Sheppard and Jacoby 1989, Lin and Lin 1998, Vittoz et al. 2001, Wells et al. 1998), convergent

plate boundaries (Jacoby and Ulan 1983, Sheppard and Jacoby 1989, Atwater and Yamaguchi

1991, Veblen et al. 1992, Yadav and Kulieshius 1992, Kitzberger et al. 1995, Jacoby et al.

1997), strike-slip faults (Stahle et al. 1992, Van Arsdale et al. 1998), reverse faults (Ruzhich et

al. 1982, Stahle et al. 1992, Van Arsdale et al. 1998), and normal faults (Sheppard and White

1995, Bekker 2004). Because earthquakes are the results of plate tectonics, the study of

geological faults with dendrochronology uses the same techniques as earthquake reconstructions.

It is important that the researcher take into account damage to trees and the spatial distribution of

sampling when determining whether or not suppression of tree rings is directly related to the

seismic events under study (Jacoby 1997, Bekker 2004).

Limitations in Dendrogeomorphology

As with all tree-ring work, accurate crossdating of samples is paramount for producing a solid

dendrogeomorphological study. Reduced tree growth due to proximity of ice or damage from an

earthquake can cause trees to have micro or missing rings for a number of years. Only

crossdating can detect these locally absent rings and provide accurate dates for geomorphic

events. Other restrictions apply to the application of dendrogeomorphology which are harder to

correct (McCarthy et al. 1991, McCarthy and Luckman 1993). A lag can occur between the

deposition of a surface and the establishment of trees on that surface. This lag can be affected by

the climate, proximity to a seed source, and the substrate itself making it difficult to accurately

362
determine the timing of some event. Other potential sources of error include sampling above the

root collar which can underestimate the age of a surface and missing the pith of the tree

(McCarthy et al. 1991). Also when dealing with suppression events, it is hard to definitely

assign the cause of such an event, so climate and other confounding factors must be thoroughly

examined. Being aware of these possible sources of error and working to minimize them can

result in accurate dating of geomorphic events in the past.

Conclusions

Dendrogeomorphology has greatly expanded in the last 20 years with a wealth of applications

and studies documenting past geologic events. The researcher has to think creatively about how

a past event may have been recorded by the trees in the area. The variety of sampling techniques

and sources of data used in this subfield of dendrochronology are probably more diverse than in

any of the other applications. As with all of dendrochronology, these studies rely on the

accuracy provided by crossdating. Without it, researchers would not be able to definitely

determine the timing of events or to assign a specific event to a growth response. In the next

chapter, I will describe the use of chemical and isotopic analysis of tree rings which is one of the

newest forms of data being drawn from tree rings.

363
Chapter 11: Dendrochemistry

Introduction

In this chapter, I will detail the applications of dendrochemistry which is the measurement of

element concentrations within tree rings to make spatial and historical estimates of element

availabilities and to understand the physiological processes that control the uptake, transport, and

sequestration of elements in secondary xylem. This field also broadly includes stable isotope

dendrochronology (the measurement of stable isotopes such as 12C, 13C, 16O, and 18O for climate

and ecological applications) and calibration of the radiocarbon dating curve. All of these

applications involve the chemical analysis of tree rings and share some similar issues in sampling

and analysis.

Dendrochemistry is the use of tree rings as indicators of past chemical fluctuations in the

environment (Cutter and Guyette 1993). Trees take up nutrients and elements through their roots

along with absorbed soil moisture, directly from the atmosphere through their leaves, and

through the bark of the tree (Donnelly et al. 1990, Leavitt 1992). Most chemical analysis deals

with the examination of the uptake of heavy metals because they are one of the main

contaminants of soil and water. Trees will take up heavy metals that often travel as part of

soluble organic compounds (ligands) and usually become fixed as part of cell walls. Wood rays

can transport the ligands from the outer rings to the inside of the sapwood. Essential elements

(most commonly phosphorous) can also be transported from interior to outer rings to meet the

364
metabolic needs of the cambium. This process is called radial translocation and is one of the

major complicating factors of dendrochemistry which will be discussed in greater detail below.

Methods of Elemental Analysis

When sampling in the field for a dendrochemistry project, the researcher must take precautions

to avoid contaminating the sample with materials like WD-40 and metals from jewelry. Cross-

contamination between trees, cores, and even individual tree rings should also be prevented.

New borers (before the Teflon coating is worn off) can be used to reduce contamination from the

metal of the borer and the borer should be cleaned with acetone to make sure that no industrial

lubricant (such as WD-40) is contaminating the samples. In the field, I wear latex gloves to

avoid contaminating the core and rinse the increment borer with isopropyl alcohol or acetone

between each core. Two cores can be taken from the same side of the tree separated by a vertical

inch so that one core can be sanded for dating and the other can be sectioned for dendrochemical

analysis based on the dating on the first core. The cores are packed in normal paper straws in the

field for drying in an oven (at about 70°C) for 24 hours in the laboratory. If one wants to study

volatile elements, such as mercury, air drying the samples for a longer period of time rather than

oven drying is preferable because these elements may be lost through vaporization.

Normal dendrochronological preparation in which cores are glued to a core mount and their

surfaces are sanded with a belt sander cannot be used in dendrochemical research for a number

of reasons, which is why the two core technique described above or the vice clamp mounting

technique described below are good approaches. First, while the core samples need to be

surfaced and dated, they also need to be removed from the core mount for the subsequent

365
chemical analysis. Not only does gluing make removal difficult, but glue may carry

contaminants. Second, sanding with a belt sander creates a lot of dust from the often metallic

components of the sandpaper that fills the surface of the cells. A good technique is to use a

temporary vise clamp made of wood (Figure 11.1) to hold the core and surface the core with a

stainless steel scalpel so that the ring structure can be observed for dating of the sample. Normal

razor blades should be avoided because they are likely to leave metal flakes which can

contaminate the wood. In some cases where the composition steel itself may provide

contaminants, alternate cutting methods may be required such as the use of a laser (Sheppard and

Witten 2005). Then the sample can be removed from the clamp and cut into individual samples

for chemical analysis. The core samples are often cut into annual segments or blocks of rings (3-

20 years) for analysis using wet chemistry, depending upon the minimum amount of wood

needed for accurate analysis of the element(s) of interest. The precision of new chemical

analysis techniques and instrumentation enable sampling at annual and even subannual

resolution if needed. Many instruments have been used for elemental analysis of tree rings.

Inductively Coupled Plasma Mass Spectrometer (ICPMS; Guyette et al. 1991), Neutron

Activation Analysis (NAA; Guyette et al. 1989), Proton Induced X-ray Emission (PIXE; Hall

1987), and Proton Induced Gamma Ray Emission (PIGE; Hall 1987) are some of the more

commonly used methods of elemental analysis in tree rings.

In the past, analysis would be conducted on clusters of tree rings because enough wood could not

be extracted from a single year of growth. Instrument detection levels are consistently

improving, thereby enabling researchers to use smaller samples for analysis. Also, some tools,

366
 

Figure 11. 1 A dendrochemistry sample in a core clamp. The clamp is constructed from two
pieces of wood that are grooved on the top center to hold the core. They are drilled through the
center of the two boards in three places and a bolt with a wing nut is inserted through each hole
to clamp down the core. In this picture, note that the researcher wears gloves to avoid
contamination of the sample and uses a razor blade instead of sandpaper to surface the sample
prior to dating (photo by Jim Speer).

367
such as a 12mm borer, have been used to obtain a larger wood sample in each year. Although,

because of radial transport within the sapwood, annual resolution is not always expected.

Radial Translocation. Radial Translocation is the movement of elements across ring

boundaries; this transfer is problematic for dendrochronologists because the resultant signal of an

element does not necessarily relate to the time that it was taken up. Whether a given element is

translocated seems to be a function of both species and environment (Cutter and Guyette 1993).

Tom Yanosky has suggested that translocation may act as a “pressure valve” that helps maintain

physiologically favorable element concentrations within living parenchyma, especially near the

cambium (Yanosky personal communication). When concentrations exceed some threshold, an

element may be translocated away from the cambium, whereas at more typical concentrations the

same element may not be particularly mobile (Vroblesky et al. 2005). Trees growing over parts

of the aquifer contaminated by large concentrations of potassium (from chemical munitions)

showed large heartwood concentrations of potassium and sapwood concentrations that increased

from outer to inner rings, whereas trees growing over uncontaminated reaches showed smaller

concentrations of potassium in heartwood and an increasing gradient from inner to outer

sapwood rings. Translocation is an issue for a number of species and for many elements, but the

salinet issue is whether there remains a usable environmental signal (Yanosky personal

communication).

Hall (1987) examined translocation of 90 elements in pitch pine from New Jersey and found that

calcium (Ca), sodium (Na), and potassium (K) were all translocated, while titanium (Ti),

manganese (Mn), iron (Fe), copper (Cu), zinc (Zn), phosphorous (P), nitrogen (N), fluorine (F),

368
magnesium (Mg), aluminum (Al), strontium (Sr), and rubidium (Rb) were not. Lead (Pb) is a

toxic but common contaminant in which many researchers, public health inspectors, and

politicians have reason to be interested. Some studies indicate that lead is an element that trees

translocate across ring boundaries, complicating any conclusions about the timing of lead

contamination in the past (Ault et al. 1970, Bindler et al. 2004), while others have shown that

lead can be accurately recorded in the heartwood of trees with dry heartwood (Guyette et al.

1991). Bindler et al. (2004) found in a research project that studied 206Pb/207Pb in Scots pine in

Sweden, that the tree-ring records did not match the timing of other natural records of lead such

as peat sequences or soil lead contamination. In one of the first studies of lead uptake by trees in

the United States, Ault et al. (1970) looked at lead concentrations in tree rings around the New

Jersey Turnpike and found that the amount of lead contained in the rings greatly increased

towards the outside of the tree. They demonstrated almost a doubling of lead concentration in

the rings over a 30-year period, which was greater than the atmospheric increase of lead over the

same time. They speculate that the trees might be excluding lead from the inner xylem and

therefore are not an accurate record of the environment. However, some tree species with dry

heartwood have been shown to be excellent long-term records of changes in lead concentrations

in the soil (Guyette et al. 1991). This research was able to reconstruct lead for the past 300 years

from the heartwood of red cedar (Juniperus virginiana) in southeastern Missouri and showed an

increase in lead after mining operations started up close to the study site in the late 1800s.

Guyette et al. (1991) also identified that sites need to have acidic soils for efficient uptake of lead

and samples have to be taken from heartwood that is relatively dry. Without these conditions,

lead may not be brought into the tree from the soils and it may be translocated in the sapwood.

369
A tree will often accumulate metabolic wastes and unneeded elements in the heartwood, making

it resistant to decay and avoided by insects. Baldcypress in North Carolina was found to

compartmentalize excess chloride in its heartwood as a response to salt water intrusion due to

dredging and sea-level rise (Yanosky et al. 1995). The researchers could estimate the timing of

saltwater intrusion because the trees seemed to transport the chloride from the oldest part of the

sapwood to the innermost sapwood which then became irrevocably sequestered as the heartwood

boundary progressed towards the outside of the tree. From elevated levels of chloride in the

heartwood and an estimate of the location of the heartwood/sapwood boundary in the past, they

were able to estimate that this contamination may have started around A.D. 1850 (Yanosky et al.

1995).

Other Confounding Factors. The solubility of elements is often controlled by other

environmental conditions such as pH (Guyette et al. 1992). Guyette et al. (1992) used mangense

(Mn) concentrations in tree rings as a measure of soil pH in red cedars on four sites in the eastern

Missouri Ozark Mountains. They were able to document changes in soil pH back to A.D. 1700

and to use soil chemistry and historical records to demonstrate the validity of their record. They

suggest that this technique may be useful on other sites to document changes in soil pH due too

acid deposition, climate change, or ecological disturbances. Elemental concentrations in trees

may be dependent on more factors than the simple levels of that element in the environment, so

that soil pH needs to be considered in dendrochemical reconstructions. Although, this also

means that soil pH itself can be measured from tree rings under the right circumstances.

370
Event Reconstructions. Ring width is often adversely affected by the availability of toxic or

highly concentrated metals. A suppression in ring width has been observed when certain

elements in the surrounding area reach toxic levels. For example, a decrease in the size of tree

rings of shortleaf pine (Pinus echinata) at Cades Cove, Tennessee in the early 1900s

corresponded with peak levels of Fe and Ti in those same rings (Baes and McLaughlin 1984).

Further research showed that the probable source of the metal was the nearby Copperhill smelter

that was active at the turn of the century. In another example, Yanoksy et al. (2001) found an

increase in chloride in the rings of oak trees that had access to water contaminated with

chlorinated hydrocarbons and they also observed a decrease in ring width in these same affected

trees, demonstrating the timing of contamination. This use of multiple lines of information from

the same cores, helps to corroborate the timing and cause of injury in dendrochronological

records.

Elements that are useful in dendrochemistry. Guyette and McGinnes (1987) found elevated

levels of Al, Fe, Zn, Cu, Sr, boron (B), and Mn in red cedar (Juniperus virginiana) trees in

Missouri that matched smelting activity in nearby mining areas. Al, Fe, and Zn showed the

greatest change in concentration in conjunction with smelting and are suggested as possible

indicators to track the timing and influence of smelting on sites with unknown exposure. As

mentioned earlier, Hall (1987) found that Ti, Mn, Fe, Cu, Zn, P, N, F, Mg, Al, Sr, and Rb did not

translocate in pitch pine from New Jersey, suggesting that they may be reliable measures of the

timing of their introduction to the environment. Vroblesky and Yanosky (1990) measured Fe

and chloride (Cl) on the Aberdeen Proving Grounds in Maryland and found that both elements

371
demonstrated an increase in concentration in the tree rings of tulip popolar when historical

activity would have caused an increase in deposition or mobilization of these elements.

Conclusions on Dendrochemistry

Trees can be used as environmental indicators that provide the timing and spatial extent of

contamination events (Vroblesky et al. 2005). An excellent knowledge of soil chemistry is

required in dendrochemistry because most metals enter the tree through root uptake.

Dendrochemical work is complicated by translocation of elements and how well the trees will

take up certain elements, but it shows potential for future work in dendrochronology. The tree-

species being selected for analysis is an important consideration in dendrochemstiry. The ability

of a tree to become a recorded of environmental chemistry is based on habitat-based factors,

xylem-based factors, and element-based factors; all of which need to be carefully considered

when choosing where and how to conduct a dendrochemical analysis (Cutter and Guyette 1993).

Radiometric Isotopes

One of the earliest applications of dendrochronology involving isotopes was the calibration of

the 14C dating curve. Radiocarbon, like all radiometric isotopes, decay at a regular rate known as

its half-life. This decay rate enables researchers to determine how much time has passed since

the isotope was incorporated in the organism. With 14C, plants take in carbon from the

atmosphere as long as they are alive, remaining in equilibrium with atmospheric levels of carbon.

However, the natural production of 14C in the atmosphere varies slightly through time as affected

by solar activity and the Earth’s magnetic field, which contribute to variable differences in

372
calendar and radiocarbon ages in the past. Thus, 14C dates needed to be calibrated to account for

these differences and to ensure the most reliable ages on samples. Wood samples from giant

sequoia (Sequoiadendron giganteum) and bristlecone pine (Pinus longaeva) were taken for the

length of each chronology and submitted to various radiocarbon labs (Ferguson 1968). The labs

were able to verify each other’s dates and the measurements from the two species also confirmed

the temporal drift in the radiocarbon record, so that at 10,000 years before present the

radiocarbon curve is off by 2,000 years. This means that a 10,000 year BP date with radiocarbon

is really a 12,000 calendar year old sample. This distinction is important when comparing 14C

production in the atmosphere to absolutely dated sunspot records for the purpose of determining

what drives 14C production.

The current, widely accepted radiocarbon calibration curve is based on the European oak tree-

ring chronology (from Ireland and Germany) going back about 10,000 calendar years, which has

been extended back nearly an additional 2000 years with a European preboreal pine chronology

(Becker 1993; Friedrich et al. 1999).

Stable Isotopes

Stable isotope analysis of tree rings is becoming one of the fastest growing applications of

dendrochronology (Long 1982, Epstein and Krishnamurthy 1990, Leavitt 1992, McCarroll and

Loader 2004) (Table 11.1). This technique analyzes isotope ratios (usually 2H/1H, 13C/12C, and
18
O/16O), with the carbon coming from CO2 in the atmosphere and the hydrogen and oxygen

signatures deriving from soil moisture (McCarroll and Loader 2004). Other stable isotope ratios

such as 15N/14N (Bukata and Kyser 2005), 34S/32S (Yang et al. 1996), and 87Sr/86Sr (English et al.

373
Table 11.1 A summary of the tree-ring isotope studies for paleoenvironmental research (modified from McCarroll and Loader 2004).

Reference Species Site Age-range Wood component Isotopes Data treatment Environmental or other signal
Anderson et al. (1998) 4 A. alba C Switzerland 1913–1995 Pooled wholewood a-cell. d13C, d18O First difference Temp, prec, and RH
Anderson et al. (2002) 4Fir A. alba C Switzerland 1913–1995 Pooled wholewood a-cell. d18O None Prec and RH
Quercus sp. and Pinus Lateglacial–
Becker et al. (1991) sylvestris S central Europe Holocene 10-year blocks. Cellulose d13C, dD None Qualitative to Lateglacial–Holocene

Bert et al. (1997) 10 A. alba France 1860–1980 5-year blocks. Holocellulose d13C Discrimination Possible age related trend
Buhay and Edwards (1995) Elm, pine, maple Ontario Canada 1610–1990 10-year blocks. Cellulose d18O, dD None Modelled d18O of prec. and air RH,
Burk and Stuiver (1981) Various N America Spatial 3 years+blocks. Cellulose d18O None RH and temp.
Sequoiadendron 1027 BC–
Craig (1954) giganteum N America AD 1649 Wholewood d13C None Link to 13C in wood and the atmo
Recent and
Dubois (1984) Pinus sylvestris United Kingdom ancient Bulk cellulose dD None Prec and RH
Dupouey et al. (1993) F. sylv. France 1950–1990 Cellulose annual d13C Ci calculated Extractable soil moisture (July) and CO2
Duquesnay et al. (1998) F. sylv. NE France 1850–1990 Pooled 10-year cellulose d13C D, Ci and WUE Age effects and long-term trends
Edwards et al. (2000) 19 Fir A. alba S Germany 1004–1980 LW cellulose d13C, dD Detrended and shifted RH and temp.
Epstein and Krishnamurthy California and
(1990) 1 P. aristata (+22 sp) global 990–1990 3–5-year blocks. Cellulose d13C, dD 25-year moving average Qualitative link to temperature
Various (incl. P. Scotland and N 1841–1970,
Epstein and Yapp (1976) aristata) America 970-1974 Wholewood 10-year blocks dD 40-year running mean Winter temperature.
Q. robur, Larix
Farmer and Baxter (1974) decidua United Kingdom 1892–1972 Wholewood d13C 10-year running mean Atmospheric C
6 Widdringtonia
February and Stock (1999) cedarb S Africa 1900–1976 Whole ring cellulose d13C Corrected not detrended Air d13C, not prec.
1967–1996,
Feng et al. (1999) 2 Picea NE China 10,040BP 5-year blocks. Cell. dD None Monsoon influence
Feng and Epstein (1995a) Pine, juniper, oak N America 1840–1990 5-year blocks. Cell. d13C Polynom, 15-yr run ave. High freq=precip. Low freq = Atmo C
Feng and Epstein (1995b) 7 various N America 1840–1990 5-year blocks. Cell. dD 25-year running average +5.3%/°C to +17%/°C
Freyer (1979a) 26 various N Hemisphere 1850–1975 2–5-year blocks. Cellulose d13C None Trends in atmospheric C
Freyer (1979b) 10 various Germany 1890–1975 2-year blocks d13C None Influence of pollution on d13C
12 Q. robur and Pinus Germany and
Freyer and Belacy (1983) sylvestris Sweden 1480–1979 1-yr and 10-year, cellulose d13C First difference “Industrial effect", temp, prec.
Gray and Se (1984) 3 Picea glauca Canada 1883–1975 5-year blocks. Cell. dD None Temperature and source water;
Gray and Thompson (1976) 1 Picea glauca Canada 1880–1969 5-year blocks. Cellulose d18O None 1.370.1%/°C
Gray and Thompson (1977) Picea glauca Canada 1882–1969 5-yr blocks WW, cell, lignin d18O None Signal strength with temperature
F. sylv., Pin. sylv., Q.
Hemming et al. (1998) rob United Kingdom 1900–1994 Various d13C, d18O, dD Corrected and first diff RH>temp.>prec.>sunshine

374
1850–1970,
10th-20th
13
Jedrysek et al. (1998) 2 Quercus+fragments Poland cent. 1-and 5-year LW celluslose d13C, dD 5-yr Running average C May–July prec.
Kitagawa and Matsumoto 12 Cryptomeria 1862–1991, 5-and 10-year blocks, a-
(1995) japonica S Japan 1846 years cellulose d13C None Temp MWP and LIA
Ratio internal to
Krishnamurthy (1996) 1 Juniperus phoenica Sinai Peninsula 1550–1950 5-year wood blocks d13C ambient CO2 Air d13C and climate, possible moisture
Krishnamurthy and Epstein
(1985) 1 Juniperus procera Kenya 1834–1979 5-year blocks. Cellulose dD None Lake levels and water stress
Lawrence and White (1984) 2 Pinus strobus N America 1960–1980 Annual (C-bound H) dD None May–August precipitation amount
Leavitt (1993) 56 Pinus edulis N America 1780–1990 5-year blocks d13C None Moisture stress
5 Fitzroya
Leavitt and Lara (1994) cupressoides Chile 1700–1900 5-year block, holocellulose d13C Corrected and ci= ca “Anthropogenic effect’’ in S Hemisphere
Leavitt and Long (1985) 10 Juniperus sp. N America 1930–1979 5-year blocks. Cellulose d13C None Temp and Precip
1712–1954, 3–4-year blocks. 13C corrected for Suess,
Libby and Pandolfi (1974) Quercus petraea Germany 1530-1800 Wholewood d13C, d18O, dD 9yr running ave Temperature
1350–1950,
Q. pet., A. alba, 1660-1950,
Libby et al. (1976) Cryptomeria japonica Germany, Japan 137-1950 5-year blocks. d18O, dD Smoothed by eye Temperature
Lipp and Trimborn (1991) Picea abies, A. alba Southern Germany 1004–1980 LW cellulose d13C, dD Unclear d13C 0.48%/°C dD 2.2%/°C
Lipp et al. (1991) A. alba Germany 1004–1980 LW cell. nit. d13C, dD Detrended and shifted d13C August temperature, dD no Signal
Discrimination
Liu et al. (1996) 4 P. tabulaeformis N China 1885–1990 Annual pooled multi radii d13C calculated June temperature and May-June Prec
Loader and Switsur (1996) 3 Pinus sylvestris United Kingdom 1760–1991 1–10-year cellulose d13C First difference Correlation with summer temp
Corrected and
McCarroll and Pawellek (2001) 36 Pinus sylvestris N Finland, 4 sites 1961–1995 LW cellulose d13C detrended Summer sun or precipitation
360 Q. robur & Q. 4890 BC–
McCormack et al. (1994) patraea United Kingdom AD 1980 10–20-year, holocellulose d13C None Difference between land and bog oaks
4 radii, 5-year blocks.
Okada et al. (1995) 3 Chamae-cyparis Japan 1680–1989 Cellulose d13C None No direct external forcing identified
Athrotaxis
Pearman et al. (1976) selaginodies Australia 1895–1970 Wholewood 5-year blocks d13C Running mean February max. temp.
Pendall (2000) P. edulis SW USA 1989–1996 E and L wood a-cell. dD None RH dominates. LW more sensitive than EW
Ramesh et al. (1985) Abies pindrow India 1903–1932 Cell. and cell. nit. d13C, dD, d18O None Identified common forcing between radii
Ramesh et al. (1986) Abies pindrow India 1903–1932 Cell. and cell. nit. d13C, dD, d18O None RH, temp, prec.
Robertson et al. (1997a, b) 10 oak Q. robur SW Finland 1895–1995 LW a-cell. d13C Standardised by filtering Prec.>RH>temp.
Robertson et al. (2001) 4 oak Q. robur E England 1895–1994 LW a-cell. d18O No detrending d18O of winter prec. and summer RH
Saurer and Siegenthaler (1989) 4 F. sylv. C Switzerland 1935–1986 3-year block. Cellulose d13C None d13C temp. and prec.
Saurer et al. (1995) 12 F. sylv. C Switzerland 1934–1989 3-year block. Cellulose d13C None d13C soil moisture status, total prec.
Saurer et al. (1998a, b) F. sylv. C Switzerland 1935–1990 3-year block. Cellulose d13C, d18O Standardised, d13C temp. and prec. d18O of source
1861–1890,
Saurer et al. (2002) Larix, Picea, Pinus Eurasia 1961-1990 30-year blocks wholewood d18O None Precipitation

375
Schiegl (1974) Picea Germany 1785–1970 5-year blocks wholewood dD None Correlation with summer temperature.
Schleser et al. (1999) 5 Picea abies Germany 1957–1992 EW and LW, Cell. and WW d13C None July temp, mean annual temp.
Sheu et al. (1996) Abies kawakamii Taiwan 1873–1992 Annual (cellulose) d13C None May–October temp.
Sonninen and Jungner (1995) 1 Pinus sylvestris Finland 1841–1990 Annual (cellulose) d13C None July temp. 0.1%°C
Stuiver and Braziunas (1987) 19 conifers N America 1100–1850 10-year cellulose d13C Standardised Latitudinal trend. RH>temp. (0.32%/°C)
Switsur et al. (1994) Quercus robur United Kingdom 1890–1990 Annual, EW, and LW cell d13C, d18O, dD None d13C Jul T and RH d18O Jul T, Jul/Aug RH
Switsur et al. (1996) 1 Quercus robur E England, UK 1869–1993 LW a-cell. d13C, d18O, dD None Temp. (D, 13C18O), RH (13C, 18O), Prec. (1
Tang et al. (1999) Pinus longaeva California, USA 1795–1993 Every 5th ring, cellulose d13C Detrended, WUE WUE increases with CO2
Tang et al. (2000) Pseudotsuga menziesii NW USA 1934–1996 Annual cellulose nitrate dD None Source-water signal
Tans and Mook (1980) 3 oak, 1 beech Netherlands 1855–1977 Annual wood, cellulose. d13C Corrected for d13C Mean summer temp. and ann temp.
Treydte et al. (2001) Spruce Picea abies Swiss Alps 1946–1995 Pooled LW a-cell. d13C Corrected for d13C Late summer temp., prec. and RH
Treydte et al. (2006) Junpiperus sp. Northern Pakistan 950 - 2006 Annual wood, cellulose d18O Corrected to VSMOW Precipitation
Waterhouse et al. (2000) 5 Pinus sylvestris N Russia 1898–1990 Annual LW a-cell. d13C 3-year running mean Correlation with flow of river Ob
Yapp and Epstein (1982) Various (25) N America 1961–1975 Cellulose nitrate dD Corrected for outliers Correlation with mean annual temp.
Zimmermann et al. (1997) Juniperus cf. tibetica Tibetan Plateau 1200–1994 5-year blocks d13C Corrected for d13C Inferred soil moisture status

Abbreviations: Q=Quercus (oak); Q. pet=Qiercis petraea (oak); A. alba=Abies alba (fir); F. sylv.=Fagus sylvatica (Beech); LW=latewood; EW=earlywood; a-cell.=a-cellulose; cell.
nit.=cellulose nitrate; NiTP=nickel tube pyrolysis; D=discrimination; Ci=internal CO2 concentration; Ca=ambient CO2 concentration; WUE=water use efficiency; T or temp.=temperature;
prec.=precipitation; RH=relative humidity; WW = Wholewood

376
2001) have been examined but not as extensively as hydrogen, carbon, and oxygen. The

biological pathways that control how trees fractionate stable isotopes and then incorporate them

into their tree rings are fairly well understood, enabling researchers to reconstruct atmospheric

temperature, humidity, and water source from which precipitation came (McCarrol and Loader

2004). Most trees take up surface water, which comes directly from precipitation. If a tree is

taking up a significant amount of ground water from deeper underground (like mesquite trees

that can have a 200 foot deep tap root in the American Southwest) the isotopic composition

could be much different. Uncertainties arise because of the variety of factors that control the

concentration of these isotopes in the atmosphere and how the plant incorporates them. Oxygen

isotopes are controlled by the amount and source of precipitation, temperature, atmospheric

humidity, and transpiration of the plant while carbon isotopes are controlled by the CO2 source,

water stress, temperature, humidity, transpiration, and abundance of the isotopes (Figure 11.2).

For example, δ18O analysis was used to develop a 1000 year-long precipitation reconstruction

from juniper trees in the Karakorum Mountains of northern Pakistan (Treydte et al. 2006). δ18O

was used instead of ring width because the precipitation signal was enhanced due to the multiple

effects of atmospheric humidity on tree physiology. The dendrochronologists demonstrated that

the 20th century was the wettest period throughout the record and that this wet interval diverged

from normal cycles but matched predictions associated with anthropogenic warming for this

region. This research demonstrates the good replication of stable isotopic records over long time

periods with relatively few samples, and it highlights the relevance of the science of

dendrochronology in helping to understand modern issues.

377
Oxygen

• Precipitation
• Temp.
• humidity Carbon
• transpiration
• source water • CO2 source
• water stress
• Temp.
• humidity
•  transpiration
• limitation
Figure 11. 2 Controls on the isotopic signature in plants (from Anderson et al. 2003).

378
One of the more recent applications of stable isotopes in tree rings, are reconstructions of past

hurricane events by the variation of δ18O in wood cellulose (Mora et al. 2006). During normal

evaporation of ocean water, more H216O evaporates from the ocean surface than H218O such that

rainwater is relatively depleted in 18O. During hurricanes, a massive amount of water is

evaporated from the ocean surface resulting in even lower 18O/16O ratios in precipitation, which

is conveyed to tree rings formed during these events.

Limitations

One of the main limitations of isotopic analysis is that multiple environmental factors can cause

a change in fractionation of the various stable isotopes (McCarroll and Pawellek 2001). To

overcome this limitation, assumptions have to be made about the site from which a sample is

collected and how the tree is interacting with the environment as it ages. These assumptions

become more tenuous the further back in time our reconstructions run. The best solution for

circumventing this problem is the use of multiple proxies, for example using three stable isotopes

(like 2H, 13C, and 18O) from the same tree samples to constrain the possible temperature variation

in a paleoclimatic study (McCarroll and Loader 2004) or using tree-ring width, density, and

stable isotopes in combination as repeated measures of temperature variation (Gagen et al.

2006). When multiple proxies are used to make independent estimates of temperature, for

example, those resultant estimates can be averaged together. As long as the signal (e.g.

atmospheric temperature) is the same between the multiple proxies, but the noise is from

different sources, the noise will average out leaving a clearer picture of past temperature

fluctuations (McCarroll and Loader 2004).

379
Standard Procedures

In angiosperms, the earlywood vessels may have some isotopic input from the previous year’s

stored photosynthates, although the overall affect of stored reserves from the previous year’s

growth is still a contested issue. Therefore, working only with the latewood from each year is

likely to provide the best record of an individual year’s isotopic composition (McCarroll and

Loader 2004).

Most analysis of tree rings for stable isotopes starts with the extraction of holocellulose or

cellulose. This is a useful starting point for stable isotopic analysis because cellulose has a well

defined composition compared to whole wood, it is structurally bound in each year, and it does

not suffer from possible translocation (Leavitt and Danzer 1993). McCarroll and Loader (2004)

suggest, however, that whole wood analysis should be reexamined in greater detail because the

time constraints of cellulose extraction slow the processing of multiple samples in automated

analysis and the isotopic variability in whole wood seems to be comparable to that in the

cellulose.

Holocellulose is the total cellulose, containing cellulose and hemicellulose (non-glucose

celluloses), and is a product that is obtained after removal of lignin from the wood. Cellulose

(C6H10O5)n is a structural polymer of glucose (C6H12O6) which is the basic sugar that is

developed during photosynthesis. The typical cellulose polymer is a linear chain consisting of

thousands of glucose building blocks. Cellulose can naturally be found bundled together as

microfibrils (a bundle of approximately 50 cellulose molecules). Cellulose can be broken down

380
into three main components: alpha cellulose, beta cellulose, and gamma cellulose. The different

types of cellulose are defined by how they behave during a sodium hydroxide NaOH extraction.

Alpha cellulose is insoluble in strong NaOH (17.5%) and can be removed as a solid; Beta

cellulose is soluble in strong NaOH but precipitates after neutralization; Gamma cellulose

remains in solution after neutralization. Hemicellulose is another product from wood which is

made up of polysaccharides that coat the surface of cellulose microfibrils, running parallel with

their structure.

To isolate cellulose for analysis of carbon or oxygen stable isotopes, lignin is removed through

oxidation in acidified sodium chlorite (NaClO2) yielding holocellulose. The hemicellulose is

then removed by reaction with sodium hydroxide (NaOH) producing alpha-cellulose (Leavitt and

Danzer 1993, McCarroll and Loader 2004). Different procedures have been perfected for

different species, for example, resins need to be removed from pine trees prior to cellulose

extraction and can be done with a Soxhlet apparatus and a solvent of toluene:ethanol in a 2:1

mixture (McCarroll and Loader 2004). See the individual papers cited in the references section

for specific methods for each stable isotope.

Once the appropriate wood extraction has been accomplished, the sample is usually converted to

a gaseous form through combustion. The ionized particles are then driven down a flight tube by

high voltage electric charge and through a magnetic field of a mass spectrometer. The different

isotopes are separated by the pull of the magnet and are collected in Faraday cup detectors. I like

to think of this like a prism separating out the different wavelengths of white light into a

rainbow. The heavier ions follow an arc with a greater radius of curvature than the lighter

381
elements. The Faraday cup detector is a series of metal cups which build up charge depending

upon how many ions come in contact with the cup. Once this charge is measured, the mass

spectrometer can report how many isotopes were in each weight category and therefore the

isotopic composition of the sample.

Results from stable isotope analyses are reported as deviations from a known standard, which is

why most stable isotopes are reported as a δ (delta) value. For example, analysis of stable carbon

ratios might be reported as -20‰ δ13C, where the sample is 20 part per thousand more depleted

in 13C than the standard. Thus, the ‰ (permil) symbol means per 1000 units, just as a %

(percent) symbol means per 100 units. Concentrations of many of these stable isotopes are fairly

low so they are reported as number of atoms of the stable isotopes for every 1000 atoms in the

standard. The standard to which hydrogen and oxygen samples are compared is the standard

mean ocean water (SMOW). For carbon isotopes, the standard is a calcium carbonate fossil

belemnite from the Pee Dee formation in South Carolina and is thus called PDB (for Pee Dee

Belemnite) (Leavitt 1992, McCarroll and Loader 2004). The fossil carbonate that was originally

used as the PDB standard has since been used up and was replaced by the Vienna-PDB (VPDB)

standard. Likewise, the SMOW has been replaced by a Vienna-SMOW (VSMOW) (Coplen

1995).

Fractionation

The process of fractionation is how any ratio of stable isotopes changes from its source to where

it is later sampled or stored. When examining 18O/16O, SMOW is the standard or baseline and

therefore δ18O of standard sea water will be 0‰. But when water evaporates off of the ocean,

382
more 16O is likely to go into the atmosphere because it is lighter, producing a lighter fraction in

the atmosphere compared to the standard (i.e. -13‰ in the initial water vapor shown in Figure

11.3). As this water moves over land and rain precipitates out of the air mass, more 18O is lost,

creating an even lighter isotopic composition of the water remaining in atmospheric vapor. In

this example, if everything else is held constant, one can tell the distance to the water source

based on the 18O/16O in an air mass. The 18O/16O ratio in sea water also changes as a result of

this evaporation and storage on land, so that marine sediments will become enriched in 18O at the

same time glaciers expand with the 18O-depleted water (snow) derived from ocean evaporation.

Further fractionation occurs in the plant itself. Diffusion of carbon isotopes from the atmosphere

into the leaf of the plant causes a -4.4‰ reduction in 13C, because the lighter 12C isotope

diffuses more rapidly into the leaves (Figure 11.4). The process of carboxylation (the addition of

a COOH group to a molecule) within the plant further fractionates the 13C of air, about -27‰ on

average. Again, many things can control how carbon isotopes fractionate in the plant, such as

leaf morphology, irradiance, air humidity, root to leaf distance, root depth, temperature, amount

and seasonality of precipitation (Figure 11.4; Lambers et al. 1998, Anderson et al. 2003).

Feedback mechanisms also exist, where the plant changes its environment through taking up

certain isotopes more so than others. By removing more 12CO2 from the atmosphere during

photosynthesis, for example, the 13CO2 concentration in ambient air will increase, changing the

chemistry of the air around the plant. Also, evaporation of water on the ground or in the

atmosphere, temperature, humidity, and the amount of circulation in the atmosphere around the

383
 

Fractionation
Example 18O/16O in Precipitation

Controls

• Rain-out
4
• Temperature
• Amount
2

• Source Area

Figure 11. 3 An example of fractionation of 18O/16O from sea water to an inland site with two
rain events. Evaporation from the ocean causes the main fractionation event with more 16O evaporating off of the
ocean so that the ratio is -13‰. Each rain event removes more 18O so that the isotopic ratio drops by -2‰ for each
event (from Anderson et al. 2003 after Siegenthaler, 1979).

384
Figure 11. 4 Factors that control fractionation in a pine tree (from McCarroll and Loader 2004).

385
leaves will all affect the availability of H and O isotopes for assimilation into the plant (Figure

11.5; Lambers et al. 1998, Anderson et al. 2003).

Different tree species will take up stable isotopes in different concentrations. Because conifer

trees transport water much less efficiently than hardwoods resulting in more fractionation from

soil water, the gymnosperms will have heavier fractionations of O and H values than

angiosperms (McCarroll and Loader 2004). Leaf shape and size in angiosperms also controls

how trees fractionate stable isotopes as does the number of stomata on a leaf, which can change

through time. One evolutionary consequence of broader leaves is more area for evaporative

cooling, and therefore broadleaf trees are less coupled to atmospheric temperature than the

needles of a conifer. Rooting depth will control the source of water that trees take up; a shallow

rooted tree, for example, will have a greater inter-annual variability in its isotopic composition

because of its dependence on rainwater, as opposed to a tree with a deep taproot that can access

groundwater. Canopy dominance also controls isotopic fractionation because dominant trees

have more direct contact with moving air, which increases the coupling with open atmospheric

temperature, humidity, and sunlight, which in turn increases rates of photosynthesis. (McCarroll

and Loader 2004). Trees growing in the understory of a forest will experience lower

temperatures, high humidity, and less direct sunlight.

Because of isotopic variation within a single ring of a tree, Leavitt and Long (1984) suggest

pooling the wood from four radii from four trees (for a total of 16 cores) to obtain accurate

values of atmospheric δ13C for the site. McCarroll and Pawellek (1998) argued against this

method stating that it would be better to quantify the δ13C differences from one core to another

386
 

Figure 11. 5 Feedback mechanisms affecting fractionation in an oak tree (from McCarroll and
Loader 2004).

387
by keeping the individual samples separate. They noted that variability between trees was much

greater than the variation within a ring, and therefore recommend taking samples from more

trees rather than more cores from one tree to average for a more precise measure of the

atmospheric isotopic concentrations. They also outlined statistical measures on latewood δ13C to

quantify the error within and between sites and suggested that developing individual tree isotope

records was a better way to understand the noise in a chronology rather than pooling wood from

multiple trees. The cost and time it takes to process samples in an isotopic study often constrains

the number of replicate samples that can be obtained as well as the temporal and spatial

resolution and extent.

Other Usable Elements

Trees can be used as a measure of change in the nitrogen cycle due to alteration of land-use by

examining variations in 15N, according to a study by Bukata and Kyser (2005). They observed

an increase in 15N in white and red oak trees growing on the margin of sites that were clear cut

and experienced a permanent change in land-use. The timing and the magnitude of the events

was recorded in the nitrogen signal in the trees, although the event seemed to last longer (19

years) than previous studies of disturbance to the nitrogen cycle when foliage alone was sampled

would suggest. After 19 years, the nitrogen levels in the tree rings returned to pre-disturbance

levels. The authors did note that transport of fluids in the sapwood tended to skew the signal so

that some minor translocation likely occurred.

Yang et al. (1996) found that trees and shrubs in Death Valley, California accurately recorded

variations in soil sulfur. The authors measured levels of 34S in the wood of the plants and

388
examined variations in growth rings of Tamarix aphylla. They found that sulfur levels in the

plants varied along with sulfur levels in the source water and root growth.

English et al. (2001) used 87Sr/86Sr to determine the source area of wooden beams in structures

from Chaco Canyon, New Mexico that were built around A.D. 1000. They examined the ratio of
87
Sr to 86Sr in spruce and fir logs from the archaeological site and examined the strontium ratio in

the soil from three surrounding mountains at the spruce/fir zone. The authors determined that the

logs came from two of the mountains both approximately 75 km away, while a third mountain

source was avoided even though the mountain was the same distance. This example shows the

use of stable isotopes as a marker to determine where wood has come from on the landscape.

Conclusions

Stable isotopes provide another record that can be drawn from tree rings and give more

information about the environment such as paleo-humidity, summer rainfall, or drought

frequency (McCarroll and Pawellek 1998). The application of dendrochronology to isotopic

analysis is growing as the methodology becomes automated, reducing the time-consuming

attention that was needed in the past by a geochemistry expert. Smaller sample sizes are being

analyzed which enables greater replication, finer sampling resolution, and longer time series.

Dendrochemistry and stable isotopes are applications of dendrochronology that have recently

grown the most out of all dendrochronological applications, changing the face of contemporary

tree-ring research. In the next chapter, “Frontiers in Dendrochronology”, I explore new

applications, techniques, and methodologies of dendrochronology that are on the cutting edge of

research questions today.

389
Chapter 12: Frontiers in Dendrochronology

Introduction

Although people have long recognized the annual nature of temperate tree rings,

dendrochronology as a discipline remains a young science, less than a century old, with its first

laboratory founded in 1937 at the University of Arizona. It is growing quickly with 55 major

laboratories located around the world today according to Dr. Henri Grissino-Mayer’s Ultimate

Tree-Ring Webpages. Four large laboratories exist in the U.S. (Laboratory of Tree Ring

Research in Tucson, Arizona; Tree-Ring Laboratory at the Lamont-Doherty Earth Observatory

New Jersey; Tree-Ring Laboratory at University of Arkansas; Laboratory of Tree-Ring Science

at the University of Tennessee) with many smaller ones scattered throughout other American

states. Canada has many active laboratories and almost every country in the European Union has

at least one dendrochronology laboratory. Russia has many active researchers and China,

Australia, and Thailand all of active research programs. South America joined in

dendrochronology research about 30 years ago with some of the first research on that continent

conducted by researchers at the Laboratory of Tree-Ring Research in Arizona in the 1970s

(Lamarche et al. 1979c, 1979d). New applications of old techniques are being explored (such as

the identification of different insect outbreak systems) and new techniques are being perfected

(such as stable isotopic analysis and image analysis). Geographic frontiers still exist in

dendrochronology with the use of wood anatomy to explore and examine tropical trees; a greater

number of usable tropical species have been identified in the last 10 years than anyone had

anticipated. Some researchers are also extending dendrochronological techniques to perennial

weeds, shrubs and even to other organisms, such as fish, clams, and turtles. Dendrochronology

390
is a vibrant field with many active researchers contributing to important societal concerns (such

as climate change) and pushing the frontiers of our knowledge of the natural world.

Dendrochronologists continue to explore interrelationships between different organisms and

trees as they are recorded in tree rings, such as pandora moth (Coloradia pandora; Speer et al.

2001) or Dothistroma needle blight (Dothistroma septosporum; Welsh 2007). What used to be

considered “noise” in a tree-ring signal is now understood to be additional biological information

that can be isolated as data layers are peeled away. For example, synchronous fruiting history

(masting; Speer 2001) has recently been identified in five oak species. Dean et al. (1985)

continues to push the human behavioral interpretations that can be made from tree-ring data,

resulting in the field of dendrochronology being more widely used with greater importance to

human society. Along those lines, Speer and Hansen-Speer (2007) have outlined the

dendroecological records that can be applied to understand more about anthropogenic ecology

and resource availability for native cultures through food resource reconstructions, such as mast

and insects, and landscape modification, such as fire history and vegetation change.

Stable Isotopes

Stable isotopes are probably the application in dendrochronology that is most quickly adding to

our knowledge of past environments. In the last decade, isotope dendrochronologists have

overcome many of the hurdles of the time-consuming sample preparation and wet chemical

processing steps in their technique. Now, as these methods become streamlined and analytical

equipment improves, isotopic dendrochronologists are able to process hundreds of samples a day

391
in a laboratory, making annual resolution on long chronologies with good sample replication a

readily achievable goal (McCarroll and Loader 2004).

New techniques are being developed such as automated laser ablation of whole samples

connected directly to an inductively coupled plasma mass spectrometer (LAS ICP-MS) so that

sampling can be conducted much more quickly and whole core surveys can be quickly

accomplished. Furthermore, LAS ICP-MS removes small amounts of the wood such that the

technique does not completely destroy cores, enabling researchers to archive their samples for

future reanalysis if necessary (Watmough et al. 1998, Schulze et al. 2004). Another procedural

advance has been the achievement of subannual resolution by using robotic micromilling to mill

small sample aliquots from a core or slab (Wurster et al. 1999, Dodd et al. 2007). The masses

needed for analysis have also decreased with improvements made to mass spectrometers;

previously, samples as small as 0.3-0.15 micrograms (mg) depending on the isotope system have

been routinely analyzed and now the technology is pushing a useful sample size of only 0.02-

0.06 mg, again depending on the isotope of interest (Patterson personal communication). By

decreasing sample sizes and increasing our ability to isolate tiny aliquots of cellulose we can

achieve sample resolutions that represent a week or less for fast growing trees. Newer laser-

robotic coupled systems envisioned for the near future should reduce the time-resolution to days

or less (Patterson personal communication). The quick processing of whole samples will move

isotope dendrochronologists beyond single tree analysis to the investigation of many trees, such

that replication between years can be completed from stands of trees with good sample depth

back through time.

392
Isotope dendrochronology truly remains one of the new frontiers in dendrochronology as many

environments have not been examined to determine the trees’ response to climatic forcing factors

that affect stable isotopic fractionation. For example, in the first stable 13C analysis from a

subalpine zone, Treydte et al. (2001) found that the spruce trees were responding to late summer

temperature, precipitation, and relative humidity. The study of climatic responses at extreme

environments, such as mountain top sites, is important in the current condition of warming,

because many of the mountain top species may be stressed by warmer conditions yet restrained

from moving to higher elevations because they are already located at the peak of the mountain.

Multiple Proxies

Stable isotopes provide a wonderful new set of information from tree rings that have expanded

the information dendrochronologists can extract from tree rings. But, as was discussed in

Chapter 11, isotopic studies are limited by the many mechanisms that can control the

concentrations of a single isotope in the tree-ring record (McCarroll and Pawellek 2001). The

use of multiple proxies such as ring width, density, and a variety of stable isotopes, helps

dendrochronologists narrow down the main driving factors that control these variables in tree

rings making climatic reconstructions more accurate (Gagen et al. 2006).

Beyond the use of multiple proxies to examine the climatic records of the past, we can use

proxies of multiple disturbances to understand the interactions between climate, fire, and insect

outbreaks (Kulakowski and Veblen 2002, Kulakowski et al. 2003). Historically, each

disturbance would be examined in isolation to determine its influence on tree growth. Now we

393
are embracing the complexity of the natural system by examining all of the disturbances that

may occur on a site and determining how they influence each other.

Image Analysis of Reflected Light

Students new to dendrochronology often ask if there is an automated technique for measuring

and dating dendrochronological samples. Many attempts have been made to automate the

process, but in the end, none of them today are equal to manual observation of the tree rings

through a good microscope, although a few tools have been developed that do allow scanning

and automated ring boundary detection. So far, these techniques are still limited to non-porous

wood species (such as pine trees) that have clear ring boundaries and are not hindered by false

ring and micro ring structures. Some of these automated techniques such as Windendro (Guay et

al. 1992, Sheppard and Graumlich 1996) and LignoVision (Rinntech, Heidelberg, Germany) can

provide good results, but the automated technique are no substitution to quality controlled

crossdating and direct observation of the wood with a good quality binocular microscope. These

automated systems work from a scanned sample of wood and optical light reflectance to

determine ring boundaries. Both Windendro and Ligno Vision are expensive and limited by the

resolution of the image, but some laboratories have had regular success dating samples with

these systems.

Potentially, image analysis from reflected light has the capability to quickly provide many

different measures of a ring, such as whole ring width, earlywood width, latewood width, cell

lumen area, double wall thickness, and circularity index (of individual cells). These latter

measurements actually provide a measure of density throughout the tree rings, based on cell

394
lumen area and cell wall thickness (Jagels and Telewski 1990). Image analysis of tree-ring

samples started with the work of Telewski et al. (1983) and went through a series of advances

with the work of Jagels and Telewski (1990), Park and Telewski (1993), Munro et al. (1996),

Sheppard et al. (1996), and Sheppard and Wiedenhoeft (2007). Methodological issues such as

removing variations in color that do not relate to ring boundaries have hindered the widespread

applicability of this technique to tree-ring analysis. Recent advances, however, have been able to

correct for this color difference in the pine heartwood-to-sapwood transition, moving the

technique closer to general use (Sheppard and Wiedenhoeft 2007).

Wood Anatomy

Wood anatomy has been studied for hundreds of year, but it is taking on new energy in regards

to the applications of dendrochronology. Gärtner (2007b) demonstrates the usefulness of root

wood anatomy to determine burial depth of roots and how geomorphological events change that

burial depth. Dietz and Schweingruber (2001) are examining root wood anatomy in

dicotyledonous perennial herbs of genera never before considered by dendrochronologists to

determine the timing of past ecological events.

Efforts are being made to tease apart finer scale effects of climate on tree growth by examing

different climate responses from weekly and daily climate records and determining the effect on

the growth of individual cells in tree rings. Rossi et al. (2006) examined the effect of day length

and temperature on weekly xylem cell production in Picea, Pinus, Abies, and Larix, finding that

the trees were more likely to respond to day length than they were to temperature.

395
Vessel size in angiosperms is being used as another response to environmental factors. Fonti and

Garcia-Gonzolez (2004) used vessel size in a European chestnut (Castanea sativa) and found

that although variability was not great, earlywood vessel size responded to temperature at the end

of the growing season when carbon reserves are put aside for the following year’s growth and at

the beginning of the growing season. These are periods of time that are not usually recorded in

ring width or density and can provide a wider range of climatic data from a tree-ring series.

Fichtler et al. (2003) demonstrated the use of wood anatomy combined with radiocarbon dating

to determine the oldest age of a number of tropical tree species from Costa Rica, finding the most

ancient tree to be 530 years old based on a ring count. Close examination of the wood anatomy

enables researchers to recognize cell types that can be used to identify ring boundaries,

especially in angiosperm wood, that have complex wood structures. These early attempts at ring

identification need to be cross checked with other methods to determine their accuracy; in this

case radiocarbon dating was used to help verify the age of these trees. This study is the first step

to recognizing the annual rings in some of these tropical tree species through wood anatomy and

further work should be able to crossdate these genera to develop absolutely dated chronologies

for more extensive regions of the tropics.

Tropical Dendrochronology

Although dendrochronolgy in tropical environments had been conducted in the 19th century, it

had long been avoided by dendrochronologists during the 20th century, because conventional

wisdom said that there was not enough seasonality in temperature or precipitation to cause trees

to shut down on a regular basis forcing annual rings to form (Worbes 2007). Today however,

396
many locations have been found in the tropics where the annual seasonality is great enough or

trees are sensitive enough to even slight climatic variations to cause the cambium to form

different wood anatomical structures that become visible as rings and many genera of trees are

being investigated that do produce annual rings in the tropics (Worbes 1995, 2002, Fichtler et al.

2003, Fichtler et al. 2004). Through this work, dendrochronologists are pushing the geographic

frontiers of dendrochronology and covering the globe with a more uniform distribution of tree-

ring chronologies as can be seen by the holdings of the International Tree Ring Databank

(ITRDB; Figure 12.1).

Early dendrochronological research in new geographic locations generally starts with a close

examination of the wood anatomy of multiple species to demonstrate that these trees produce

annual rings (e.g. Villalba et al. 1985, Boninsegna et al. 1989). The next step is to test the

annual nature of the chronology and the reliability of ring development through crossdating

(Villalba and Boninsegna 1989, Stahle 1999, Speer et al. 2004). Stahle (1999) suggests a series

of tests to determine if trees have annual rings, based on the ability to crossdate the samples

between trees and across multiple sites, their correlation to climate, and through blind

crossdating tests of samples with known ages. By following these methods, researchers have

been able to document seasonal production of tree rings in Africa (Gourlay 1995, February and

Stock 1998, Stahle 1999, Fichtler et al. 2004), India (Bhattacharyya et al. 1992), Indonesia

(D’Arrigo et al. 1994), Java (Pumijumnong et al. 1995), Mexico (Stahle 1999), Brazilian

Amazon region (Worbes 1989, Vetter and Botosso 1989, Schöngart et al. 2004), Honduras

(Johnson 1980), and the Dominican Republic (Speer et al. 2004).

397
Figure 12. 1 The National Climatic Data Center (NCDC) runs the World Data Center (WDC) for
paleoclimatology which houses the tree-ring chronologies of the International Tree Ring
Databank (ITRDB). Note the prevalence of chronologies across North America, Europe, and
Siberia. There are a growing number of chronologies coming from South America and New
Zealand, but a lack of chronologies from Africa and the tropics.

398
Until the 21st century, the continent of Africa was largely unexplored for the occurrence of tree

species that develop annual rings. Gourlay (1995) was able to demonstrate that trees in the

genus Acacia did develop rings that were bounded by marginal parenchyma cells and that were

highly correlated with annual rainfall. February and Stock (1998) examined the potential of two

Podocarpus species in South Africa but were unable to date the rings due to poor circuit

uniformity and locally absent rings. Stahle et al. (1999) developed chronologies from

Pterocarpus angolensis in Zimbabwe at 18 degrees south latitude enabling the researchers to

inform the forest managers about the growth rates for this valuable timber species in tropical

Africa. Research throughout Africa continues to find older and datable species that can

contribute to the data gap on this continent (Eshete and Stahl 1999, Trouet et al. 2001, Worbes et

al. 2003, Fichtler et al. 2004, Schöngart et al. 2004, Verheyden et al. 2004, Couralet et al. 2005,

Schöngart et al. 2005, Verheyden 2005, Trouet et al. 2006, Therrell et al. 2006).

Unique techniques have also been used to determine the growth of tropical trees and to test for

the annual nature of tree rings in some locations. The artificial increase in 14C in the atmosphere

from atomic weapons testing caused what is known as the “Bomb Spike” which is an elevated

level of 14C that peaks in 1962 in all trees in the world (Worbes and Junk 1989). The Bomb

Spike can be used to find the 1962 ring and determine how much growth has occurred in the

intervening years. This information may be useful for forest managers so that they can, at least

roughly, determine the growth rate of tropical trees that are being harvested. Mariaux (1981)

developed an original way for determining the growth rate and ring production in tropical trees

by wounding the cambium and returning to see if rings were produced annually.

399
Unique Environments

Many forests in North America are limited in long-term tree-ring chronologies by extensive

logging that occurred around A.D. 1900 and removed many of the old trees, especially in the

eastern United States. Many sites have well-preserved old trees on sites that the loggers avoided

because the wood was not merchantable such as old lava flows, barrens, or swampy sites. This

can be overcome by finding unique sites to which loggers did not have access (Larson et al.

1999). For example, the Niagara Escarpment in southern Ontario has Eastern White Cedar

(Thuja occidentalis) growing on cliffs that obtain ages in excess of 1,000 years (Kelly et al.

1992). Research in an environment such as this requires a dendrochronologist who is also skilled

at rock climbing.

Submerged logs in anoxic environments (without oxygen) can be preserved for centuries,

millennia, and much more rarely, tens of millions of years. These trees have the potential to

produce long tree-ring chronologies (Larson and Melville 1996). Places such as the Great Lakes,

bog environments, lakes in Siberia, and debris piles at the base of cliffs have all produced

preserved logs. Sampling for submerged wood requires dendrochronologists that are interested

in scuba diving or have connections with scuba professionals. Submerged environments are

great preservation sites that have not yet been explored to their full potential.

Extremely long chronologies are now being developed from subfossil wood (usually buried

wood that has been preserved but not yet fossilized by the replacement of cellulose by minerals

such as quartz) that is being mined from stream banks on tributaries of the Missouri River

(Guyette and Stambaugh 2003). This wood has the potential to form the longest tree-ring

400
chronology in the world, extending back some 15,000 years, but work is slow and expensive.

Target dates for the samples are determined by the density of the wood (because the longer a

specimen is buried the more mass it loses even in anoxic conditions) and radiocarbon dates.

Once the sample is placed into a broad period of history, crossdating attempts can be more

productive when trying to date the sample against floating chronologies from that time period.

As in any tree-ring chronology, adequate sample depth throughout the series is necessary in

order to make reliable interpretations from the data; for this project, the task becomes significant

because of the length of the chronology.

Sclerochronology

Sclerochronology is the use of boney structures in a variety of organisms to determine the age of

the organism and to develop long term histories of the environment from those structures.

Sclerochronology has been applied to fish otoliths (“earstones” in the head of fish used for

balance and other sensory information), shells of clams, shells of turtles, and even in dinosaur

bones. Counting these increments to determine a best guess of the age of the organisms has been

done for some time. For example, Aristoltle discussed determination of a fishes age by counting

rings in an otolith. Dendrochronologists are now bringing the tool of crossdating to their field to

provide quality control and verification of these ages. With the benefit of crossdating, longer

chronologies with annual resolution can now be developed (Guyette and Rabeni 1995, Black et

al. 2005, Helama et al. 2006). During the North American Dendroecological Fieldweek in 2006

held at the Hatfield Marine Science Center, Bryan Black led a group that dated geoduck clams

from the Vancouver Island area. They found that the clams had better series intercorrelations

(~0.716) than any other chronology developed during the fieldweek and also had a significant

401
correlation (r=0.6) with January to March sea-surface temperature (Black et al. 2006). Black et

al. (2005) examined 50 rock fish (Sebastes diploproa) otoliths and found that they ranged from

30-84 years in age. The otoliths’ growth was significantly correlated with the Northern

Oscillation Index (r=0.51, p=0.0001), an upwelling index (r=0.40, p=0.002), and the Pacific

Decadal Oscillation (r=-0.29, p=0.007) (Black et al. 2005). These results have implications for

fisheries management, because many of these fish live much longer than previously expected

thereby reducing their rate of replacement. Perhaps the most tragic example of species depletion

relates to the age estimate of orange roughy to be an average of 15 years of age, when in reality

they were 150 years old. With longer lives and slower development rates, fish would have to be

taken less frequently to maintain viable populations.

Conclusion

I hope you have gleaned from this book that dendrochronology is a vibrant field of science that is

growing quickly today with many frontiers remaining to be explored. Researchers throughout

our discipline are investigating most areas of the natural sciences and touching on a wide variety

of fields. Science as a whole is becoming more interdisciplinary and dendrochronology is a tool

that can be applied to questions in ecology, archaeology, climatology, geology, hydrology, and

atmospheric sciences. These varied uses of the same basic skills make dendrochronology a

useful field of study which contributes to our knowledge of the natural world. It is a tool that can

be used by practitioners in different fields and is also a science with its own theories.

402
References

Abbe, C. 1893. Notes by the Editor. Monthly Weather Review 21:331–332.

Abrams, M.D 1985. Fire history of oak gallery forests in a northeast Kansas tallgrass prairie,

American Midland Naturalist 114: 188–191.

Abrams, M.D. 1992. Fire and development of oak forests. Bioscience 42(5): 346–353.

Abrams, M.D. 2000. Fire and the ecological history of oak forests in the Eastern United States.

In: Proceedings: Workshop on fire, people, and the central hardwoods landscape, ed. D.

A. Yaussy, pp. 46–55. USDA Forest Service, Northeastern Research Station. Gen. Tech.

Rep. NE –274, Richmond, KY.

Abrams, M.D. and Nowacki, G.J. 1992. Historical variation in fire, oak recruitment, and post-

logging accelerated successions in central Pennsylvania. Bulletin of the Torrey Botanical

Club 119(1): 19–28.

Abrams, M.D., Orwig, D.A., and Demeo, T.E. 1995. Dendroecological analysis of successional

dynamics for a presettlement-origin white-pine-mixed-oak forest in the southern

Appalachians, USA. Journal of Ecology 83(1): 123–133.

Agee, J.K. 1993. Fire Ecology of the Pacific Northwest Forests. Island Press, Washington, D.C.

Ahlstrom, R.V.N., Van West, C.R., Dean, J.S.. 1995. Environmental and Chronological Factors

in the Mesa Verde-Northern Rio Grande Migration. Journal of Anthropological

Archaeology 14:125–142.

Aldrich J.M. 1912. Larvae of a Saturniid moth used as food by California Indians. Journal of

the New York Entomological Society 20:28–31.

Aldrich, J.M. 1921. Coloradia pandora Blake, A moth of which the caterpillar is used as food

by Mono Lake Indians. Annuls of the Entomological Society of America 14:36–38.

403
Alestalo, J. 1971. Dendrogeomorphological interpretation of geomorphic processes. Fennia

105: 1–140.

Alfaro, R.I., Thomson, A.J., and Van Sickle, G.A. 1985. Quantification of Douglas-fir growth

losses caused by western spruce budworm defoliation using stem analysis. Canadian

Journal of Forest Research. 15:5–9.

Anderson, L., Carlson, C.E., and Wakimoto, R.H. 1987. Forest fire frequency and western

spruce budworm outbreaks in western Montana. Forest Ecology and Management 22:

251–260.

Anderson, W.T., Bernasconi, S.M., and McKenzie, J.A. 1998. Oxygen and carbon isotopic

record of climatic variability in tree ring cellulose (Picea abies): an example from central

Switzerland (1913–1995). Journal of Geophysical Research 103/D24, 31625–31636.

Anderson, W.T., Bernasconi, S.M., McKenzie, J.A., Saurer, M., and Schweingruber, F. 2002.

Model evaluation for reconstructing the oxygen isotopic composition in precipitation

from tree ring cellulose over the last century. Chemical Geology 182: 121–137.

Anderson, W.T., Evans, S.L., Hernadez, R., Pinzon, M.C., Kirby, M.E., Sternberg, L., and

Grissino-Mayer, H.D. 2003. Oxygen isotopic records in tree rings as indicators of the

climatic and atmospheric circulation changes from Europe, North America, and South

America. Oral presentation at the Association of America Geographers (AAG)

conference New Orleans. March 5 – 8, 2003.

Anonymous 1923. How a tree tells the story of forest fires. Scientific American 128(3): 183.

Arno, S.F. and Sneck, K.M. 1973. A method for determining fire history in conifer forests of the

mountain west. U.S. Department of Agriculture, Forest Service. GTR-INT-42. 26pp.

404
Asshof, R., Schweingruber, F.H., and Wermelinger, B. 1999. Influence of a gypsy moth

(Lymantria dispar L.) outbreak on radial growth and wood-anatomy of Spanish chestnut

(Castanea sativa Mill.) in Ticino (Switzerland). Dendrochronologia 16–17: 133–145.

Atwater, B.F. and Yamaguchi, D.K. 1991. Sudden, probably coseismic submergence of

Holocene trees and grass in coastal Washington State. Geology 19: 706–709.

Au, R. and Tardif, J.C. 2007. Allometric relationships and dendroecology of the dwarf shrub

Dryas integrifolia near Churchill, subarctic Manitoba. Canadian Journal of Botany

85(6): 585–597.

Ault, W.V., Senechal, R.G., and Erelebach, W.E. 1970. Isotopic composition as a natural tracer

of lead in our environment. Environmental Science and Technology 4: 305–313.

Babbage, C. 1838. Note M, On the Age of Strata, as Inferred from the Rings of Trees

Embedded in Them. In: The Ninth Bridgewater Treatise, a Fragment, 2nd edition. John

Murray, London. pp. 256–264.

Baes, C.F. and McLaughlin, S.B. 1984. Trace elements in tree rings: Evidence of recent and

historical air pollution. Science 224: 494–497.

Bailey, I.W. 1925a. The “spruce budworm” biocoenose. I. Frost rings as indicators of the

chronology of specific biological events. Botanical Gazette. 80: 93–101.

Bailey, I.W. 1925b. Notes on the “spruce budworm” biocoenose. II Structural abnormalities in

Abies balsamea. Botanical Gazette. 80: 300–310.

Baillie, M.G.L. 1982. Tree-Ring Dating and Archaeology. The University of Chicago Press,

Chicago. 274pp.

Baillie, M.G.L. 1995. A Slice Through Time: Dendrochronology and Precision Dating. B.T.

Batsford Ltd. London. 176pp.

405
Bailey, I.W. 1925. The "spruce budworm" biocoenose. 1. Frost rings as indicators of the

chronology of specific biological events. Botanical Gazette 80: 93–101.

Bannister, B. 1963. Dendrochronology. In: Brothwell, D. and Higgs, E. (eds.). Science in

Archaeology. New York, Basic Books. pp. 161–176.

Bannister, B. and Robinson, W.J. 1975. Tree-ring dating in archaeology. World Archaeology

7(2): 210–225.

Barber, V.A., Juday, G.P., Finney, B.P., and Wilmking, M. 2004. Reconstruction of summer

temperatures in interior Alaska from tree-ring proxies: Evidence for changing synoptic

climate regimes. Climatic Change 63(1-2): 91-120.

Barnston, A.G. and Livezey, R.E. 1987. Classification, seasonality and persistence of low-

frequency atmospheric circulation patterns. Monthly Weather Review 115: 1083–1126.

Bauch, J. and Eckstein, D. 1970. Dendrochronological dating of oak panels of Dutch

seventeenth century paintings. Studies in Conservation 15: 45–50.

Becker, B. 1979. Holocene tree-ring series from southern central Europe for archaeological

dating, radiocarbon calibration, and stable isotope analysis. In: R. Berger and H.E. Suess,

eds., Radiocarbon Dating. University of California Press, Berkeley, CA: 554–565 .

Becker, B. 1991. The history of dendrochronology and radiocarbon calibration. In: Taylor,

R.E. Long, A., and Kra, R.S. (eds.). Radiocarbon after four decades: An

interdisciplinary perspective. Spring Verlag, Heidelberg: 34–39.

Becker, B. 1993. An 11,000-year German oak and pine dendrochronology for radiocarbon

calibration. Radiocarbon 35: 201–213.

406
Becker, B., Delorme, A. 1978. Oak chronology for central Europe. The extension from medieval

to prehistoric times. In: Fletcher, J. (ed.) Dendrochronology in Europe. British

Archaeological Reports International Series 51: 59–64.

Becker, B., Kromer, B., and Trimborn, P. 1991. A stable isotope tree-ring timescale of the Late

Glacial/Holocene boundary. Nature 353: 647–649.

Begin, Y. 2000. Ice-push disturbances in high-Boreal and Subarctic lakeshore ecosystems since

A.D. 1830, northern Quebec, Canada. The Holocene 10(2): 179–189.

Bégin, Y., Bérubé, D. and Grégoire, M. 1993. Downward migration of coastal conifers as a

response to recent land emergence in eastern Hudson Bay, Québec. Quaternary Research

40: 81–88.

Begin, Y. and Filion, L. 1988. Age of landslides along the Grande Rivière de la Baleine estuary,

eastern coast of Hudson Bay, Quebec, Canada. Boreas 17: 289–299.

Bekker, M.F. 2004. Spatial variation in the response of tree rings to normal faulting during the

Hebgen Lake Earthquake, southwestern Montana, USA. Dendrochronologia 22(1): 53–

59.

Bennett, D.D., Schmid, J.M., Mata, S.A., Edminster, C.B. 1987. Growth Impact of the North

Kaibab Pandora Moth Outbreak. Research Note RM-474, pg. 1–4.

Bergeron, Y. 1991. The influence of island and mainland lakeshore landscapes on boreal forest

fire regimes. Ecology 72(6): 1980–1992.

Bergeron, Y. 2000. Species and stand dynamics in the mixed woods of Quebec's southern boreal

forest. Ecology 81(6): 1500–1516.

407
Bergeron, Y. and Archambault, S. 1993. Decreasing frequency of forest fires in the southern

boreal zone of Québec and its relation to global warming since the end of the “Little Ice

Age”. The Holocene 3(3): 255–259.

Bergeron, Y. and Brisson, J. 1990. Fire regime in red pine stands at the northern limit of the

species range. Ecology 71(4): 1352–1364.

Bert, D., Leavitt, S.W., and Dupouey, J.-L. 1997. Variations of wood d13C and water use

efficiency of Abies alba during the last century. Ecology 78(5): 1588–1596.

Bhattacharyya, A. Yadav, R.R., Borgaonkar, H.P., and Pant, G.B. 1992. Growth ring analysis

on Indian tropical trees: dendroclimatic potential. Current Science 62: 736–741.

Billamboz, A. 1992. Tree-ring analysis from an archaeodendrological perspective. The structural

timber from the southwest German lake dwellings. In: T.S. Bartholin, B.E. Berglund, D.

Eckstein, F.H. Schweingruber, and O. Eggertsson, eds., Tree Rings and Environment:

Proceedings of the International Symposium, Ystad, South Sweden, 3-9 September,

1990. Lundqua Report (Department of Quaternary Geology, Lund University, Sweden)

34: 34-40.

Billamboz, A. 2003. Tree rings and wetland occupation in Southwest Germany between 2000

and 500 BC: Dendrochronology beyond dating in tribute to F.H. Schweingruber. Tree-

Ring Research 59(1): 37-49.

Bindler, R., Renberg, I., Klaminder, J., and Emteryd, O. 2004. Tree rings as Pb pollution

archives? A comparison of 206Pb/207Pb isotope ratios in pine and other environmental

media. The Science of the Total Environment 319: 173–183.

Biondi, F. 1999. Comparing tree-ring chronologies and repeated timber inventories as forest

monitoring tools. Ecological Applications 9(1): 216–227.

408
Biondi, F., Gershunov, A., and Cayan, D.R. 2001. North Pacific decadal climate variability

since 1661. Journal of Climate 14(1): 5–10.

Biondi, F. and Waikul, K. 2004. DENDROCLIM2002: A C++ program for statistical

calibration of climate signals in tree-ring chronologies. Computers & Geosciences 30:

303–311.

Black, B.A., Allman, R., Campbell, B., Darbyshire, R., Klökler, D., Kormanyos, R., Munk, K.,

and Peterson, P. 2006. Application of dendrochronology techniques to a Pacific geoduck

clam (Panopea abrupta) in northern British Columbia, Canada. Final report of the 16th

North American Dendroecological Fieldweek (NADEF). Hatfield Marine Science Center

May 30 – June 7. 13pp.

Black, B.A., Boehlert, G.W., and Yoklavich, M.M. 2005. Using tree-ring crossdating

techniques to validate annual growth increments in long-lived fishes. Canadian Journal of

Fisheries and Aquatic Sciences 62(10): 2277–2284.

Blais, J.R. 1954. The recurrence of spruce budworm infestations in the past century in the Lac

Seul area of northwestern Ontario. Ecology 35: 62–71.

Blais, J.R. 1957. Some relationships of the spruce budworm to black spruce. Forestry

Chronicle 33: 364–372.

Blais, J.R. 1958a. Effects of defoliation by spruce budworm (Choristoneura fumiferana clem.)

on radial growth at breast height of balsam fir (Abies balsamea (L.) Mill.) and white

spruce (Picea glauca (Moench) Voss.). Forestry Chronicle 34: 39–47.

Blais, J.R. 1958b. Effects of 1956 spring and summer temperatures on spruce budworm

populations in the Gaspe Peninsula. Canadian Entomologist 90: 354–361.

409
Blais, J.R. 1961. Spruce budworm outbreaks and the climate of the boreal forest in eastern

North America. Report to the Quebec Society for the Protection of Plants (1959): 69–75.

Blais, J.R. 1962. Collection and analysis of radial-growth data from trees for evidence of past

spruce budworm outbreaks. Forestry Chronicle 38(4): 474–484.

Blais, J.R. 1965. Spruce budworm outbreaks in the past three centuries in the Laurentide Park,

Quebec. Forest Science 11: 130–138.

Blais, J.R. 1983. Trends in the frequency, extent, and severity of spruce budworm outbreaks in

eastern Canada. Canadian Journal of Forest Research 13: 539–547.

Blake, E. A., and Wagner, M. R. 1987. Collection and consumption of pandora moth,

(Coloradia pandora lindseyi Lepidoptera: Saturniidae), Larvae by Owens Valley and

Mono Lake Paiutes. Bulletin of the Entomological Society of America 33(1): 5p.

Blasing, T.J. and Fritts, H.C. 1976. Reconstructing past climatic anomalies in the north Pacific

and western North America from tree-ring data. Quaternary Research 6: 563–579.

Boninsegna, J.A., Villabla, R., Amarilla, L., and Ocampo, J. 1989. Studies of tree rings, growth

rates, and age-size relationships of tropical tree species in Misiones, Argentina.

International Association of Wood Anatomists (IAWA) Bulletin 10(2): 161–169.

Bonzani, R.M., Carlisle, R.C., and King, F.B. 1991. Dendrochronology of the Pennsylvania

Main Line Canal Lock Number Four, Pittsburgh. North American Archaeologist 12(1):

61–73.

Bortolot, Z.J., Copenheaver, C.A., Longe, R.L., Van Aardt, J.A.N. 2001. Development of a white

oak chronology using live trees and a post-Civil War cabin in south-central Virginia.

Tree-Ring Research 57(2): 197–203.

410
Bradley, R.S. 1999. Paleoclimatology: Reconstructing Climates of the Quaternary. Harcourt

Academic Press. Amsterdam. 613pp.

Briffa, K.R. 2000. Annual climate variability in the Holocene: interpreting the message of

ancient trees. Quaternary Science Reviews 19: 87-105.

Briffa, K.R., Bartholin, T.S., Eckstein, D. Jones, P.D., Karlen, W., Schweingruber, F.H. and

Zetterberg, P. 1990. A 1400-year tree-ring record of summer temperatures in

Fennoscandia. Nature 346, 434–439.

Briffa, K. and Jones, P.D. 1990. Basic chronology statistics and assessment. In: Cook, E.R. and

Kairiukstis, L.A. (eds.). Methods of dendrochronology: Applications in the

environmental sciences. Kluwer Academic Publishers, Dordrecht, The Netherlands. pp.

137–162.

Briffa, K.R., Jones, P.D., Bartholin, T.S., Eckstein, D., Schweingruber, F.H., Karlen, w.,

Zetterberg, P., and Eronen, M. 1992. Fennoscandian summers from AD 500:

temperature changes on short and long timescales. Climate Dynamics 7, 111–119.

Briffa, K.R., Jones, P.D., Schweingruber, F.H., and Osborn, T.J. 1998. Influence of volcanic

eruptions on Northern Hemisphere summer temperature over the past 600 years. Nature

393: 450-455.

Briffa, K.R., Osborn, T.J., Schweingruber, F.H., Harris, I.C., Jones, P.D., Shiyatov, S.G., and

Vaganov, E.A. 2001. Low-frequency temperature variations from a northern tree ring

density network. Journal of Geophysical Research 106(D3): 2929–2941.

Brown, P.M. 1996. OLDLIST: A database of maximum tree ages. In: Dean, J.S., Meko, D.M.,

and Swetnam, T.W. (eds.) Tree Rings, Environment, and Humanity. Radiocarbon 1996:

727–731.

411
Brown, P.M., Hughes, M.K., Baisan, C.H., Swetnam, T.H., and Caprio, A.C. 1992. Giant

sequoia ring-width chronologies from the central Sierra Nevada, California. Tree-Ring

Bulletin 52: 1–14.

Brown, P.M., Shepperd, W.D., Brown, C.C., Mata, S.A., and McClain, D.L. 1995. Oldest known

Engelmann spruce. USDA Forest Service Rocky Mtn. Forest and Range Exper. Sta. Res.

Note RM-RN-534. 6pp.

Brubaker, L.B. 1980. Spatial patterns of tree growth anomalies in the Pacific Northwest. Ecology

61(4): 798–807.

Brubaker, L.B., and Greene, S.K. 1979. Differential effects of Douglas-fir tussock moth and

western spruce budworm on radial growth of grand fir and Douglas-fir. Canadian Journal

of Forest Research 9:95–105.

Brunstein, F.C., and Yamaguchi, D.K. 1992. The oldest known Rocky Mountain bristlecone

pines (Pinus aristata Engelm.). Arctic and Alpine Research 24:253–256.

Buhay, W.M., Edwards, T.W.D. 1995. Climate in southwestern Ontario, Canada, between AD

1610 and 1885 inferred from oxygen and hydrogen isotopic measurements of wood

cellulose from trees in different hydrologic settings. Quaternary Research 44: 438–446.

Bukata, A.R. and Kyser, T.K. 2005. Response of the nitrogen isotopic composition of tree-rings

following tree-clearing and land-use change. Environmental Science and Technology 39:

7777–7783.

Burg, J. P. 1978. A new technique for time series data. In: D. G. Childers, editor. Modern

Spectrum Analysis. IEEE Press, New York, New York, USA. p 42–48.

Burk, R.L. and Stuiver, M. 1981. Oxygen isotope ratios in trees reflect mean annual temperature

and humidity. Science 211: 1417–1419.

412
Butler, D.R. 1987. Teaching general principles and applications of dendrogeomorphology.

Journal of Geological Education 35: 64–70.

Campbell, E.M., Alfaro, R.I., and Hawkes, B. 2007. Spatial distribution of mountain pine beetle

outbreaks in relation to climate and stand characteristics: A dendroecological analysis.

Journal of Integrative Plant Biology 49(2): 168–178.

Campbell, L.J. and Laroque, C.P. 2005. Dendrochronological Analysis of Endangered

Newfoundland Pine Marten Habitat: Decay Classification of Coarse Woody Debris in

Western Newfoundland. Mount Allison Dendrochronology Laboraotry, MAD Lab

Report 2005–07. 33pp.

Campbell, L.J. and Laroque, C.P. 2007. Decay progression and classification in two old-growth

forests in Atlantic Canada. Forest Ecology and Managemnt 238(1–3): 293–301.

Case, R.A. and MacDonald, G.M. 2003. Dendrochronological analysis of the response of

tamarack (Larix laricina) to climate and larch sawfly (Pristiphora erichsonii) infestations

in central Saskatchewan. Ecoscience 10(3): 380–388.

Chaloner, W.G., and Creber, G.T. 1973. Growth rings in fossil woods as evidence of past

climates. In: D.H. Tarling and S.K. Runcorn, eds., Implications of Continental Drift to

the Earth Sciences. Academic Press, London: 425–437.

Clark, N.E., Blasing, T.J., and Fritts, H.C. 1975. Influence of interannual climatic fluctuations

on biological systems. Nature 256(5515): 302–305.

Clements, F.E. 1910. The life history of lodgepole burn forests. USDA Forest Service Bulletin

79: 1–56.

Cochrane, J. and Daniels, L.D. 2008. Striking a balance: Safe sampling of partial stem cross-

sections in British Columbia. BC Journal of Ecosystems and Management 9(1):38–46.

413
Conkey, L.E., Keifer, M., and Lloyd, A.H. 1995. Disjunct jack pine (Pinus banksiana Lamb)

structure and dynamics, Acadia National Park, Maine. Ecoscience 2(2): 168–176.

Cook, E.R. 1985. A time series analysis approach to tree ring standardization, Dissertation, The

University of Arizona, Tucson.

Cook, E.R. 1992. A conceptual linear aggregate model for tree rings. In: Cook E.R. and

Kairiukstis, L.A. (eds.) Methods of dendrochronology: Applications in the

environmental sciences. Kluwer Academic Publishers. London. 98–104 pp.

Cook, E., Bird, T., Peterson, M., Barbetti, M., Buckley, B., D’Arrigo, R., and Francey, R. 1992.

Climatic change over the last millennium in Tasmania reconstructed from tree-rings. The

Holocene 2: 205–217.

Cook, E.R., Bird, T., Peterson, M., Barbetti, M., Buckley, B., D'Arrigo, R., Francey, R., Tans, P.

1991. Climatic change in Tasmania inferred from a 1089-year tree-ring chronology of

huon pine. Science 253: 1266–1268.

Cook, E. R., Briffa, K.R., and Jones, P.D. 1994. Spatial regression methods in

dendroclimatology: A review and comparison of two techniques. International Journal of

Climatology 14: 379-402.

Cook, E.R., Briffa, K.R., Meko, D.M., Graybill, D.A., and Funkhouser, G. 1995. The ‘segment

length curse’ in long tree-ring chronology development for paleoclimatic studies. The

Holocene 5(2): 229–237.

Cook E.R., Buckley B.M., D'Arrigo R.D., and Peterson, M.J. 2000. Warm-season temperatures

since 1600 B.C. reconstructed from Tasmanian tree rings and their relationship to large-

scale sea surface temperature anomalies. Climate Dynamics 16(2/3): 79–91.

414
Cook, E.R. and Cole, J. 1991. On predicting the response of forests in eastern North America to

future climatic change. Climatic Change 19: 271–282.

Cook, E.R. and Holmes, R.L. 1986. Users manual for program ARSTAN. In: Holmes, R.L.,

Adams, R.K., and Fritts, H.C. Tree-ring chronologies of western North America:

California, eastern Oregon and northern Great Basin. Chronology Series 6. Tucson:

Laboratory of Tree-Ring Research, University of Arizona: 50–56.

Cook, E. R., and Kairiukstis, L.A. 1990. Methods of Dendrochronology: Applications in the

environmental sciences. Kluwer Academic Publishers, Dordrecht, The Netherlands.

Cook, E.R., Jacoby, Jr., G.C. 1977. Tree-ring-drought relationships in the Hudson Valley, New

York. Science 198: 399–401.

Cook, E.R., Johnson, A.H., and Blasing, T.J. 1987. Forest decline: modeling the effect of

climate in tree rings. Tree Physiology 3: 27–40.

Cook, E.R., Meko, D.M., Stahle, D.W., and Cleaveland, M.K. 1999. Drought reconstructions

for the continental United States. Journal of Climate 12:1145–1162.

Cook, E.R., Woodhouse, C.A., Eakin, C.M., Meko, D.M., and Stahle, D.W. 2004. Long-term

aridity changes in the western United States. Science 306:1015–1018.

Copenheaver, C.A., Fuhrman, N.E., Gellerstedt, L.S., and Gellerstedt, P.A. 2004. Tree

encroachment in forest openings: a case study from Buffalo Mountain, Virginia. Castanea

69(4): 297–308.

Copenheaver, C.A., Pokorski, E.A., Currie, J.E., and Abrams, M.D. 2006. Causation of false

ring formation in Pinus banksiana: A comparison of age, canopy class, climate and

growth rate. Forest Ecology and Management 236(2–3): 348–355.

Coplen, T.B. 1995. The Discontinuance of SMOW and PDB. Nature 375: 285.

415
Corominas, J. and Moya, J. 1999. Reconstructing recent landslide activity in relation to rainfall

in the Llobregat River basin, Eastern Pyrenees, Spain. Geomorphology 30: 79–93.

Corona, E. 1986. Dendrocronologia: principi e applicazioni. In: Dendrocronologia: Principi e

Applicazioni. Atti del Seminario a Verona nei giorni 14 e 15 Novembre 1984. Istituto

Italiano di Dendrocronologia, Verona, Italy: 7–32.

Couralet, C., Sass-Klaassen, U., Sterck, F., Bekele, T., and Zuidema, P.A. 2005. Combining

dendrochronology and matrix modelling in demographic studies: An evaluation for

Juniperus procera in Ethiopia. Forest Ecology and Management 216(1-3): 317-330.

Craig, H., 1954. Carbon-13 variations in Sequoia rings and the atmosphere. Science 119, 141–

144.

Cronon, W. 1997. John Muir: Nature writings. The Library of America. New York. 888 pp.

Cropper, J.P. 1979. Tree-ring skeleton plotting by computer. Tree-Ring Bulletin 39: 47–60.

Cufar, K. 2007. Dendrochronology and past human activity – A review of advances since 2000.

Tree-Ring Research 63(1): 47–60.

Currey, D.R. 1965. An ancient bristlecone pine stand in eastern Nevada. Ecology 46(4): 564–

566.

Cutter, B.E. and Guyette, R.P. 1993. Anatomical, chemical, and ecological factors affecting tree

species choice in dendrochemistry studies. Journal of Environmental Quality 22(3): 611-

619.

D’Arrigo, R.D., Cook, E.R., Jacoby, G.C., and Briffa, K.R. 1993. NAO and sea surface

temperature signatures in tree-ring records from the North Atlantic sector. Quaternary

Science Reviews 12: 431–440.

416
D’Arrigo, R.D., Cook, E.R., Salinger, M.J., Palmer, J., Krusic, P.J., Buckley, B.M., and Villalba,

R. 1998. Tree-ring records from New Zealand: long-term context for recent warming

trend. Climate Dynamics 14: 191–199.

D’Arrigo, R.D. and Jacoby, G.C. 1991. A 1000-year record of winter precipitation from

northwestern New Mexico, USA: a reconstruction from tree-rings and its relation to El

Niño and the Southern Oscillation. The Holocene 1(2): 95–101.

D’Arrigo, R.D. and Jacoby, G.C. 1993. Tree growth-climate relationships at the northern boreal

forest tree line of North America: Evaluation of potential response to increasing carbon

dioxide. Global Biogeochemical Cycles 7(3): 525–535.

D’Arrigo, R.D., Jacoby, G.C., Krusic, P.J. 1994. Progress in dendroclimatic studies in

Indonesia. Terrestrial, Atmospheric and Oceanic Sciences 5: 349–363.

D’Arrigo, R.D., Malmstrom, C.M., Jacoby, G.C., Los, S.O., and Bunker, D.E. 2000.

Correlation between maximum latewood density of annual tree rings and NDVI based

estimates of forest productivity. International Journal of Remote Sensing 21(11): 2329–

2336.

Daniels, L.D. 2003. Western red cedar population dynamics in old-growth forests: Contrasting

ecological paradigms using tree rings. The Forestry Chronicle 79(3): 517–530.

Daniels, L.D., Dobry, J., Klinka, K., and Feller, M.C. 1997. Determining year of death of logs

and snags of Thuja plicata in southwestern coastal British Columbia. Canadian Journal

of Forest Research 27: 1132–1141.

Daniels, L.D. and Veblen, T.T. 2003. Regional and local effects of disturbance and climate on

altitudinal treelines in northern Patagonia. Journal of Vegetation Science 14: 733–742.

417
Daniels, L.D. and Veblen, T.T. 2004. Spatiotemporal influences of climate on atlitudinal

treeline in northern Patagonia. Ecology 85(5): 1284–1296.

Danzer, S.R. 1996. Rates of slope erosion determined from exposed roots of ponderosa pine at

Rose Canyon Lake, Arizona. In: Dean, J.S., Meko, D.M., and Swetnam, T.W. (eds.) Tree

Rings, Environment, and Humanity: Proceedings of the International Conference, Tucson

Arizona, 17–21 May 1994. Radiocarbon, Tucson. pp. 671–678.

Dean, J.S. 1978. Tree-ring dating in Archeology. University of Utah, Miscellaneous Paper

Number 24: 129–163.

Dean , J.S. 1997. Dendrochronology. In: Taylor and Aitken (eds.) Chronometric dating in

Archaeology. Plenum Press, New York. pp. 31–64.

Dean, J.S., Euler, R.C., Gumerman, G.J., Plog, F., Hevly, R.H, Karlstrom, T.N.V. 1985. Human

behavior, demography, and paleoenvironment on the Colorado plateaus. American

Antiquity 50: 537–554.

Dettinger, M. D., Ghil, M., Strong, C. M., Weibel, W., and Yiou, P. 1995. Software expedites

singular-spectrum analysis of noisy time series. Eos Transactions of the American

Geological Union 76(2):12, 14, 21.

Dey, D.C. and Guyette, R.P. 2000. Anthropogenic fire history and red oak forests in south-

central Ontario. Forestry Chronicle 76(2): 339–347.

Dieterich J. H. and Swetnam T.W. 1984. Dendrochronology of a fire-scarred pandora pine.

Forest Science 30(1): 238–247.

Dietz, H. and Schweingruber, F.H. 2001. Development of growth rings in roots of

dicotyledonous perennial herbs: experimental analysis of ecological factors. Bulletin of

the Geobotanical Institute ETH 67: 97–105.

418
Dodd, J.P., Patterson, W.P., Holmden, C., and Brasseur, J.M. 2007. Robotic micromilling of

tree-ring cellulose: a new tool for obtaining sub-seasonal environmental isotope records.

Chemical Geology special publication for The Stable Isotope Session from the 7th

International Conference on Dendrochronology, Beijing, China (in press).

Donnelly, J.R., Shane, J.B., and Schaberg, P.C. 1990. Lead mobility with the xylem of red

spruce seedlings: Implications for the development of pollution histories. Journal of

Environmental Quality 19: 268-271.

Douglass, A.E. 1909. Weather cycles in the growth of big trees. Monthly Weather Review 37:

225–237.

Douglass, A.E. 1914. A method of estimating rainfall by the growth of trees. Carnegie Institute

of Washington Publication No. 192: 101–121.

Douglass, A.E. 1917. Climatic records in the trunks of trees. American Forestry 23(288): 732–

735.

Douglass, A.E. 1920. Evidence of climatic effects in the annual rings of trees. Ecology 1(1): 24–

32.

Douglass, A.E. 1921. Dating our prehistoric ruins: how growth rings in trees aid in establishing

the relative ages of the ruined pueblos of the Southwest. Natural History 21(1): 27–30.

Douglass, A.E. 1929. The secret of the Southwest solved by talkative tree rings. National

Geographic Magazine. 56: 736–770.

Douglass, A.E. 1941. Crossdating in dendrochronology. Journal of Forestry 39(10): 825–831.

Drake, D.C., Naiman, R.J., and Helfield, J.M. 2002. Reconstructing salmon abundance in

rivers: An initial dendrochronological evaluation. Ecology 83(11): 2971–2977.

419
Du, S., Yamanaka, N., Yamamoto, F., Otsuki, K., Wang, S., and Hou, Q. 2007. The effect of

climate on radial growth of Quercus liaotungensis forest trees in Loess Plateau, China.

Dendrochronologia 25(1): 29-36.

Dubois, A.D. 1984. On the climatic interpretation of the hydrogen isotope ratios in recent and

fossil wood. Bulletin de la Societe Belge de Geologie 93: 267–270.

Duff, G.H., and Nolan, N.J. 1953. Growth and morphogenesis in the Canadian forest species.

Part I. The controls of cambial and apical activity in Pinus resinosa Ait. Canadian

Journal of Botany 31:471–513.

Duff, G.H., and Nolan, N.J. 1957. Growth and morphogenesis in the Canadian forest species.

Part II. Species increments and their relation to the quantity and activity of growth in

Pinus resinosa Ait. Canadian Journal of Botany 35:527–572.

Dupouey, J.-L., Leavitt, S.W., Choisnel, E., and Jourdain, S. 1993. Modeling carbon isotope

fractionation in tree-rings based upon effective evapotranspiration and soil-water status.

Plant, Cell and Environment 16: 939–947.

Duquesnay, A., Breda, N., Stievenard, M., and Dupouey, J.L. 1998. Changes of tree-ring d13C

and water-use efficiency of beech (Fagus sylvatica L.) in north-eastern France during the

past century. Plant, Cell and Environment 21: 565–572.

Duvick, D. N. and Blasing, T. J. 1981. A dendroclimatic reconstruction of annual precipitation

amounts in Iowa since 1680. Water Resources Research 17: 1183-1189.

Eckstein, D. 1972. Tree-ring research in Europe. Tree-Ring Bulletin 32: 1–18.

Eckstein, D. and Bauch, J. 1969. Beitrag zu Rationalisierung eines dendrochronologischen

Verfahrens und zu Analyse seiner Aussagesicherheit. Forstwissenschaftliches

Centralblatt 88: 230–250.

420
Eckstein, D. and Pilcher, J.R. 1990. Dendrochronology in Western Europe. In: Cook, E.R., and

Kairiukstis, L.A. (eds.). Methods of Dendrochronology: Applications in the

Environmental Sciences. pp 11–13.

Eckstein, D., Richter, K., Aniol, R.W., and Quiehl, F. 1984. Dendroclimatological

investigations of the beech decline in the southwestern part of the Vogelsberg (West

Germany). In German with English abstract. Forstwissenschaftliches Centralblatt 103:

274–290.

Eckstein, D., Wazny, T., Bauch, K., and Klein, P. 1986. New evidence for the

dendrochronological dating of Netherlandish paintings. Nature 320: 465–466.

Eckstein, D. and Wrobel, S. 2007. Dendrochronological proof of origin of historic timber –

Retrospective and perspectives. In: Haneca, K., Verheyden, A., Beekman, H., Gärtner,

H., Helle, G., and Schleser, G. TRACE – Tree Rings in Archaeology, Climatology and

Ecology. Proceedings of the DENDROSYMPOSIUM 2006. April 20th – 22nd 2006,

Tervuren, Belgium. Volume 5:8–20.

Edwards, T.W.D., Graf, W., Trimborn, P., Stichler, W., and Payer, H.D. 2000. d13C response

surface resolves humidity and temperature signals in trees. Geochimica et Cosmochimica

Acta 64: 161–167.

Egger, H., Gassmann, P., and Burri, N. 1985. Situation actuelle du travail au laboratoire de

dendrochronologie de Neuchatel. Dendrochronologia 3: 177–192.

Eisenhart, K.S. and Veblen, T.T. 2000. Dendroecological detection of spruce bark beetle

outbreaks in northwestern Colorado. Canadian Journal of Forest Research 30(11): 1788–

1798.

421
English, N.B., Betancourt, J.L., Dean, J.S., and Quade, J. 2001. Strontium isotopes reveal

distant sources of architectural timber in Chaco Canyon, New Mexico. Proceedings of

the National Academy of Science (PNAS) 98(21): 11891–11896.

Epstein, S. and Krishnamurthy, R.V. 1990. Environmental information in the isotopic record in

trees. Philosophical Transactions of the Royal Society 330A, 427–439.

Epstein, S. and Yapp, C.J. 1976. Climatic implications of the D/H ratio of hydrogen in C–H

groups in tree cellulose. Earth and Planetary Science Letters 30: 252–261.

Eshete, G. and Stahl, G. 1999. Tree rings as indicators of growth periodicity of acacias in the Rift

Valley of Ethiopia. Forest Ecology and Management 116(1-3): 107-117.

Esper J, Cook E.R., Krusic P.J., Peters K., and Schweingruber F.H. 2003a. Tests of the RCS

method for preserving low-frequency variability in long tree-ring chronologies. Tree-

Ring Research 59: 81–98.

Esper, J., Cook, E.R., and Schweingruber, F.H. 2002. Low-frequency signals in long tree-ring

chronologies and the reconstruction of past temperature variability. Science 295: 2250–

2253.

Esper, J. and Schweingruber, F.H. 2004. Large-scale treeline changes recorded in Siberia.

Geophysical Research Letters 31: 1–5.

Esper, J., Shiyatov, S.G., Mazepa, V.S., Wilson, R.J.S., Graybill, D.A., and Funkhouser, G.

2003b. Temperature-sensitive Tien Shan tree ring chronologies show multi-centennial

growth trends. Climate Dynamics 21(7-8): 699-706.

Falcon-Lang, H.J. 1999. The Early Carboniferous (Courceyan-Arundian) monsoonal climate of

the British Isles: evidence from growth rings in fossil woods. Geological Magazine

136(2): 177–187.

422
Falcon-Lang, H.J. 2005. Global climate analysis of growth rings in woods, and its implications

for deep-time paleoclimate studies. Paleobiology 31(3): 434–444.

Fantucci, R. and Sorriso-Valvo, M. 1999. Dendrogeomorphological analysis of a slope near

Lago, Calabria (Italy). Geomorphology 30: 165–174.

Farmer, J.G. and Baxter, M.S. 1974. Atmospheric carbon dioxide levels as indicated by the

stable isotope record in wood. Nature 247: 273–275.

Fastie, C.L. 1995. Causes and ecosystem consequences of multiple pathways of primary

succession at Glacier Bay, Alaska. Ecology 76(6): 1899–1916.

February, E.C. and Stock, W.D. 1998. An assessment of the dendrochronological potential of

two Podocarpus species. The Holocene 8(6): 747–750.

February, E.C. and Stock, W.D. 1999. Declining trends in the 13C/12C ratio of atmospheric

carbon dioxide from tree rings of South African Widdringtonia cedarbergensis.

Quaternary Research 52: 229–236.

Feng, X., Cui, H., Tang, K., and Conkey, L.E. 1999. Tree-ring delta-D as an indicator of Asian

monsoon intensity. Quaternary Research 51: 262–266.

Feng, X.H and Epstein, S. 1995a. Carbon isotopes of trees from arid environments and

implications for reconstructing atmospheric CO2 concentration. Geochimica et

Cosmochimica Acta 59: 2599–2608.

Feng, X.H and Epstein, S. 1995b. Climatic temperature records in dD data from tree rings.

Geochimica et Cosmochimica Acta 59: 3029–3037.

Ferguson, C.W. 1968. Bristlecone pine: Science and Esthetics. Science 159(3817): 839–846.

423
Ferguson, C.W., Lawn, B., and Michael, H.N. 1985. Prospects for the extension of the

bristlecone pine chronology: Radiocarbon analysis of H-84-1. Meteoritics 20(2): 415–

421.

Fichtler, E., Clark, D.A., and Worbes, M. 2003. Age and long-term growth of trees in an old-

growth tropical rain forest, based on analyses of tree rings and C-14. Biotropica 35(3):

306–317.

Fichtler, E., Trouet, V., Beeckman, H., Coppin, P., and Worbes, M. 2004. Climatic signals in tree

rings of Burkea africana and Pterocarpus angolensis from semiarid forests in Namibia.

Trees - Structure and Function 18(4): 442–451.

Filion, L., Payette, S., Delwaide, A., and Bhiry, N. 1998. Insect defoliators as major disturbance

factors in the high-altitude balsam fir forest of Mount Mégantic, southern Quebec.

Canadian Journal of Forest Research 28: 1832-1842.

Fletcher, J.M. 1976. A group of English royal portraits painted soon after 1513: A

dendrochronological study. Studies in Conservation 21(4): 171–178.

Fletcher, J.M. 1977. Tree-ring chronologies for the 6th to 16th centuries for oaks of southern and

eastern England. Journal of Archaeological Science 4(4): 335–352.

Food and Agriculture Organization (FAO). 1973. Inventario forestall. Inventario y fomento de

los recursos forestales: Republica Dominicana. Technical Report No. 3 FO:SF/DOM 8.

Rome: United Nationals Food and Agriculture Organization.

Fonti, P. and Garcia-Gonzalez, I. 2004. Suitability of chestnut earlywood vessel chronologies for

ecological studies. New Phytologist 163(1): 77–86.

Freyer, H.D. 1979a. On the 13C record in tree rings. Part 1. 13C variations in Northern

Hemispheric trees during the last 150 years. Tellus 31: 124–137.

424
Freyer, H.D. 1979b. On the 13C record in tree rings. Part 2. Registration of microenvironmental

CO2 and anomalous pollution effect. Tellus 31: 308–312.

Freyer, H.D., and Belacy, N. 1983. 12C/13C records in Northern Hemispheric trees during the

past 500 years: anthropogenic impact and climatic superpositions. Journal of Geophysical

Research 88: 6844–6852.

Friedrich, M., Kromer, B., Kaiser, K.F., Spurk, M., Hughen, K.A., and Johansen, S.J. 2001.

High-resolution climate signals in the Bolling-Allerod Interstadial (Greenland Interstadial

1) as reflected in European tree-ring chronologies compared to marine varves and ice-

core records. Quaternary Science Reviews 20(11): 1223-1232.

Friedrich, M., Kromer, B., Spurk, M., Hofmann, J., and Kaiser, K.F., 1999. Paleo-environment

and radiocarbon calibration as derived from Lateglacial/Early Holocene tree-ring

chronologies. Quaternary International 61: 27-39.

Friedrich, M., Remmele, S., Kromer, B., Hofmann, J., Spurk, M., Kaiser, K.F., Orcel, C., and

Kuppers, M. 2004. The 12,460-year Hohenheim oak and pine tree-ring chronology from

central Europe – A unique annual record for radiocarbon calibration and

paleoenvironment reconstructions. Radiocarbon 46(3): 1111–1122.

Fritts, H.C. 1971. Dendroclimatology and dendroecology. Quaternary Research 1: 419–449.

Fritts, H.C. 1976. Tree rings and climate. Academic Press. New York. 567p.

Fritts, H.C. 2001. Tree Rings and Climate. The Blackburn Press. New Jersey. 567p.

Fritts, H.C., Blasing, T.J., Hayden, B.P., and Kutzbach, J.E. 1971. Multivariate techniques for

specifying tree-growth and climate relationships and for reconstructing anomalies in

paleoclimate. Journal of Applied Meteorology 10(5): 845–864.

425
Fritts, H. C. and Dean, J. S. 1992. Dendrochronological modeling of the effects of climatic

change on tree-ring width chronologies from Chaco Canyon and environs. Tree Ring

Bulletin. 52:31–58.

Fritts, H.C. and Shao, X.M. 1992. Mapping climate using tree-rings from western North

America. In Bradley, R.S. and Jones, P.D. (eds.) Climate since A.D. 1500. Routledge,

London. pp: 269–295.

Fritts, H.C., Smith, D.G., Cardis, J.W., and Budelsky, C.A. 1965. Tree-ring characteristics

along a vegetation gradient in northern Arizona. Ecology 46(4): 393–401.

Fritts, H. C. and Swetnam, T. W. 1989. Dendroecology: a tool for evaluating variations in past

and present forest environments. In: Begon, M., Fitter, A. H., Ford, E. D. and Macfadyen,

A. (eds.). Advances in Ecological Research. Academic Press, London. 19:111–88.

Fule, P.R., Covington, W.W., and Moore, M.M. 1997. Determining reference conditions for

ecosystem management of southwestern ponderosa pine forests. Ecological Applications

7(3): 895–908.

Gagen, M., McCarroll, D., and Edourard, J.L. 2006. Combining ring width, density, and stable

carbon isotope proxies to enhance the climate signal in tree-rings: An example from the

southern French Alps. Climatic Change 78: 363–379.

Gärtner, H. 2003. The applicability of roots in dendrogeomorphology. In; Schleser, G.,

Winiger, M., Bräuning, A., Gärtner, H., Helle, G., Jansma, E., Neuwirth, B., and Treydte,

K. (eds.). Tree Rings in Archaeology, Climatology and Ecology, Volume 1. Proceedings

of the Dendrosymposium 2002. Schriften des Forschungszentrum Jülich, Reihe Umwelt

33: 120–124

426
Gärtner, H. 2007a. Glacial landorms, tree rings: Dendrogeomorphology. In: Elias, S.A. (ed.)

Encyclopedia of Quaternary Sciences. Volume 2. Elsevier, pp: 979-988.

Gärtner, H. 2007b. Tree roots – Methodological review and new development in dating and

quantifying erosive processes. Geomorphology 86: 243-251.

Gärtner, H., Schweingruber, F.H., Dikau, R. 2001. Determination of erosion rates by analyzing

structural changes in the growth pattern of exposed roots. Dendrochronologia 19(1): 81–

91.

Giardino, J.R., Shroder, J.F., and Lawson, M.P. 1984. Tree-ring analysis of movement of a

rock-glacier complex on Mount Mestas, Colorado, U.S.A. Arctic and Alpine Research

16(3): 299–309.

Girardclos, O., Lambert, G., and Lavier, C. 1996. Oak tree-ring series from France between

4000 B.C. and 8000 B.C. In: Dean, J.S., Meko, D.M., and Swetnam, T.W. (eds.). Tree

rings, environment, and humanity: proceedings of the international conference, Tucson,

Arizona, 17–21 May 1994. Tucson, Arizona, Radiocarbon. pp. 751–768.

Girardin, M.P., Tardif, J.C., Flannigan, M.D., and Bergeron, Y. 2006. Synoptic-scale

atmospheric circulation and boreal Canada summer drought variability of the past three

centuries. Journal of Climate 19(10): 1922–1947.

Glock, W.S. 1941. Growth Rings and Climate. The Botanical Review 7(12): 649–713.

Glock, W.S. 1951. Cambial frost injuries and multiple growth layers at Lubbock, Texas. Ecology

32(1): 28–36.

Gore, A.P., Johnson, E.A., and Lo, H.P. 1985. Estimating the time a dead tree has been on the

ground. Ecology 66(6): 1981–1983.

427
Gottesfeld, A.S. and Gottesfeld, L.M.J. 1990. Floodplain dynamics of a wandering river,

dendrochronology of the Morice River, British Columbia, Canada. Geomorphology 3:

159–179.

Gourlay, I.D. 1995. Growth ring characteristics of some African acacia species. Journal of

Tropical Ecology 11(1): 121–140.

Graumlich, L. 1993. A 1000-year record of temperature and precipitation in the Sierra Nevada.

Quaternary Research 39: 249-255.

Graumlich, L.J., Brubaker, L.B., and Grier, C.C. 1989. Long-term trends in forest net primary

productivity: Cascade Mountains, Washington. Ecology 70(2): 405–410.

Gray, S.T., Graumlich, L.J., Betancourt, J.L., and Pederson, G.T. 2004. A tree-ring based

reconstruction of the Atlantic Multidecadal Oscillation since 1567 A.D. Geophysical

Research Letters 31(12) Article Number L12205.

Gray, J. and Se, J.S. 1984. Climatic implications of the natural variation of D/H ratios in tree-

ring cellulose. Earth and Planetary Science Letters 70: 129–138.

Gray, J. and Thompson, P. 1976. Climatic information from 18O/16O ratios of cellulose in tree-

rings. Nature 262: 481–482.

Gray, J. and Thompson, P. 1977. Climatic information from 18O/16O analysis of cellulose,

lignin and wholewood from tree-rings. Nature 270: 708–709.

Graybill, D.A. and Shiyatov, S.G. 1992. Dendroclimatic evidence from the northern Soviet

Union. In: Bradley, R.S. and Jones, P.D. (eds.). Climate since A.D. 1500. London,

England, Routledge 393–414.

428
Greve, U., Eckstein, D., Aniol, R.W., and Scholz, F. 1986. Dendroclimatological investigations

on Norway spruce under different loads of air pollution. Allaemeine Forst- und

Jagdzeitung 157: 174–179.

Grissino-Mayer, H.D. 1995. Tree-ring reconstructions of climate and fire history at El Malpais

National Monument, New Mexico. Ph.D. Dissertation (Geography). The University of

Arizona, Tucson.

Grissino-Mayer, H.D. 1996. A 2129 year annual reconstruction of precipitation for

northwestern New Mexico, USA. In J.S. Dean, D.M. Meko, and T.W. Swetnam, eds.,

Tree Rings, Environment, and Humanity. Radiocarbon 1996, Department of Geosciences,

The University of Arizona, Tucson: 191–204.

Grissino-Mayer, H.D. 2001. Evaluating crossdating accuracy: A manual and tutorial for the

computer program COFECHA. Tree-Ring Research 57(2): 205–221.

Grissino-Mayer, H.D. 2003. A Manual and Tutorial for the Proper Use of an Increment Borer.

Tree-Ring Research 59(2): 63–79.

Grissino-Mayer, H.D., Blount, H.C., and Miller, A.C. 2001. Tree-ring dating and the

ethnohistory of the naval stores industry in southern Georgia. Tree-Ring Research 57(1):

3–13.

Grissino-Mayer, H.D., Cleavland, M.K., and Sheppard, P.R. 2002. Mastering the Rings. The

Strad 113:408–415.

Grissino-Mayer, H.D., Deweese, G.G., Williams, D.A. 2005. Tree-ring dating of the Karr-

Koussevitzky double bass: A case study in dendromusicology. Tree-Ring Research 61(2):

77–86.

429
Grissino-Mayer, H.D. and Fritts, H.C. 1997. The International Tree-Ring Data Bank: an

enhanced global database serving the global scientific community. The Holocene 7(2):

235–238.

Grissino-Mayer, H.D., Sheppard, P.R., and Cleaveland, M.K. 2003. Dendrochronological dating

of stringed instruments: A re-evaluation. Journal of the Violin Society of America 18(2):

127–174.

Grissino-Mayer, H.D., Sheppard, P.R., and Cleaveland, M.K. 2004. A dendroarchaeological re-

examination of the "Messiah" violin and other instruments attributed to Antonio

Stradivari. Journal of Archaeological Science 31(2): 167–174.

Grissino-Mayer, H.D., Swetnam, T.W., and Adams, R.K. 1997. The rare, old-aged conifers of

El Malpais – Their role in understanding climate change in the American Southwest.

New Mexico Bureau of Mines & Mineral Resources, Bulletin 156: 155–161.

Groven, R. and Niklasson, M. 2005. Anthropogenic impact on past and present fire regimes in a

boreal forest landscape of southeastern Norway. Canadian Journal 0f Forest Research

35: 2719–2726.

Guay, R., Gagnon, R., and Morin, H. 1992. MacDendro, a new automatic and interactive tree-

ring measurement system based on image processing. In: Bartholin, T.S., Berglund,

B.E., Eckstein, D., Schweingruber, F.H., and Eggertsson, O. (eds.) Tree rings and

Environment: Proceedings of the International Symposium, 3–9 Sept. 1990, Ystad,

Sweden. Department of Quaternary Geology, Lund University, Lund Sweden, Lundqua

Rep. No. 34: 128–131.

430
Guiot, J. 1990. Methods of calibration. In: Cook, E. R., and L.A. Kairiukstis (eds.). Methods of

Dendrochronology: Applications in the environmental sciences. Kluwer Academic

Publishers, Dordrecht, The Netherlands. pp 165–178.

Guiot, J. 1991. The bootstrapped response function. Tree-Ring Bulletin 51:39–41.

Guiot, J., Nicault, A., Rathgeber, C., Edouard, J.L., Guibal, E., Pichard, G., Till, C. 2005. Last-

millennium summer-temperature variations in western Europe based on proxy data. The

Holocene 15(4): 489-500.

Gutsell, S.L. and Johnson, E.A. 2002. Accurately ageing trees and examining their height-

growth rates: Implications for interpreting forest dynamics. Journal of Ecology 90: 153–

166.

Guyette, R.P., Cutter, B.E., and Henderson, G.S. 1989. Long-term relationships between

molybdenum and sulfur concentrations in redcedar tree rings. Journal of Environmental

Quality 18(3): 385-389.

Guyette, R.P., Cutter, B.E., and Henderson, G.S. 1991. Long-term correlations between mining

activity and levels of lead and cadmium in tree-rings of eastern red-cedar. Journal of

Environmental Quality 20(1): 146-150.

Guyette, R.P., Henderson, G.S., and Cutter, B.E. 1992. Reconstructing soil pH from manganese

concentrations in tree-rings. Forest Science 38(4): 727-737.

Guyette, R. and McGinnes, E.A. 1987. Potential in using elemental conectrations in radial

increment of old growth eastern red cedar to examine the chemical history of the

environment. In: Jacoby, G.C. and Hornbeck, J.W. (eds.) Proceedings of the

International Symposium on Ecological Aspects of Tree-Ring Analysis. U.S.

Department of Commerce, Washington, D.C. Publication No CONF-86081-44: 671–680.

431
Guyette, R.P., Muzika, R.M., and Dey, D.C. 2002. Dynamics of an anthropogenic fire regime.

Ecosystems 5: 472–486.

Guyette, R.P. and Rabeni, C.F. 1995. Climate response among growth increments of fish and

trees. Oecologia 104: 272–279.

Guyette, R.P. and Stambaugh, M.C. 2003. The age and density of ancient and modern oak wood

in streams and sediments. IAWA Journal 24(4): 345–353.

Hadley, K.S. 1994. The role of disturbance, topography, and forest structure in the development

of a mountain forest landscape. Bulletin of the Torrey Botanical Club 121(1): 47–61.

Hall, G.S. 1987. Multielemental analysis of tree-rings by proton induced x-ray (PIXE) and

gamma ray emission (PIGE). In: Jacoby, G.C. and Hornbeck, J.W. (eds.) Proceedings of

the International Symposium on Ecological Aspects of Tree-Ring Analysis. U.S.

Department of Commerce, Washington, D.C. Publication No CONF-86081-44: 681–689.

Hantemirov, R.M., Gorlanova, L.A., and Shiyatov, S.G. 2004. Extreme temperature events in

summer in northwest Siberia since AD 742 inferred from tree rings. Palaeogeography,

Palaeoclimatology, Palaeoecology 209(1-4): 155-164.

Hart, E. 2002. Effects of woody debris on channel morphology and sediment storage in

headwater streams in the Great Smoky Mountains, Tennessee-North Carolina. Physical

Geography 23(6): 492–510.

Hart, E. 2003. Dead wood: Geomorphic effects of coarse woody debris in headwater streams,

Great Smoky Mountains. Journal of The Tennessee Academy of Science 78(2): 50–54.

Hartig, R. 1888. Das Fichten- und Tannenholz des bayerischen Waldes. Centralblatt f. das

gesamte Forstwesen 14: 357–364, 437–442.

432
Haury, E.W. 1962. HH-39: Recollections of a dramatic moment in Southwestern archaeology.

Tree-Ring Bulletin: 24(3–4): 11–14.

Heinselman, J.E. 1973. Fire in the virgin forests of the Boundary Water Canoe Area,

Minnesota. Quaternary Research 3: 329–382.

Heizer, R. F. 1954. The First Dendrochronologist. American Antiquity 22(2):186–188.

Helama, S., Schone, B.R., Black, B.A., Dunca, E. 2006. Constructing long-term proxy series

for aquatic environments with absolute dating control using a sclerochronological

approach: Introduction and advanced applications. Marine and Freshwater Research

57(6): 591–599.

Hemming, D.L., Switsur, V.R., Waterhouse, J.S., Heaton, T.H.E., Carter, and A.H.C. 1998.

Climate and the stable carbon isotope composition of tree ring cellulose: an

intercomparison of three tree species. Tellus 50B: 25–32.

Hessl, A.E. and Graumlich, L.J. 2002. Interactive effects of human activities, herbivory and fire

on quaking aspen (Populus tremuloides) age structures in western Wyoming. Journal of

Biogeography 29: 889-902.

Heyerdahl, E.K. and Card, V. 2000. Implications of paleorecords for ecosystem management.

Trends in Ecology and Evolution (TREE) 15(2): 49–50.

Heyerdahl, E.K. and McKay, S.J. 2001. Condition of live fire-scarred ponderosa pine trees six

years after removing partial cross sections. Tree-Ring Research 57(2): 131–139.

Hildahl, V., Reeks, W.A. 1960. Outbreaks of the forest tent caterpillar, Malacosoma disstria

Hbn., and their effects on stands of trembling aspen in Manitoba and Saskatchewan.

Canadian Entomologist 92: 199–209.

433
Hirschboeck, K.K., Ni, F., Wood, M.L., and Woodhouse, C.A. 1996. Synoptic

dendroclimatology: Overview and outlook. In: Dean, J.S., Meko, D.M., and Swetnam,

T.W. (eds.) Tree Rings, Environment, and Humanity. Radiocarbon 1996: 205–223.

Hoadley, R. B. 1990. Identifying Wood: Accurate results with simple tools. Newtown, CT: The

Taunton Press.

Holmes, R.L. 1983. Computer-assisted quality control in tree-ring dating and measurement.

Tree-Ring Bulletin 43: 69–78.

Holmes, R. L., and Swetnam, T. W. 1994a. Dendroecology program library: program

OUTBREAK user's manual. 5p. Unpublished document. On file with: Laboratory of

Tree-Ring Research. The University of Arizona. Tucson, AZ. 85721.

Holmes, R.L. and Swetnam, T.W. 1994b. Dendroecology program library: Program EVENT

User’s Manual superposed epoch analysis in fire history studies. 7p. Unpublished

document. On file with: Laboratory of Tree-Ring Research. The University of Arizona.

Tucson, AZ. 85721.

Hough, F.B. 1882. The Elements of Forestry. Robert Clarke and Company, Cincinnati. 381

pp.

Huang, J.G. and Zhang, Q.B. 2007. Tree rings and climate for the last 680 years in Wulan area of

northeastern Qinghai-Tibetan Plateau. Climatic Change 80: 369-377.

Huber, B. 1935. Die physiologische bedeutung der ring- und zerstreut- porigkeit. Ber. Deut.

Bot. Ges. 53:711–719.

Hupp, C.R. 1984. Dendrogeomorphic evidence of debris flow frequency and magnitude at

Mount Shasta, California. Environmental Geology 6(2): 121–128.

434
Hupp, C.R., Osterkamp, W.R., and Thornton, J.L. 1987. Dendrogeomorphic evidence and

dating of recent debris flows on Mount Shasta, Northern California. United States

Geological Survey. Professional Paper 1396-B. 45 pp.

Hurrell, J.W. 1995. Decadal trends in the North Atlantic Oscillation: Regional temperatures and

precipitation. Science 269: 676–679.

Jacoby, G.C. 1997. Application of tree ring analysis to paleoseismology. Reviews of Geophysics

35(2): 109–124.

Jacoby, G.C., Bunker, D.E. and Benson, B.E. 1997. Tree-ring evidence for an A.D. 1700

Cascadia earthquake in Washington and northern Oregon. Geology 25(11): 999–1002.

Jacoby, G.C. and D’Arrigo, R.D. 1999. Tree-ring indicators of climate change at Northern

Latitudes. World Resource Review 11(1): 21–29.

Jacoby, G.C., Sheppard, P.R. and Sieh, K.E. 1988. Irregular recurrence of large earthquakes

along the San Andreas Fault: Evidence from trees. Science 241: 196–199.

Jacoby, G.C., Ulan, L.D. 1983. Tree-ring indications of uplift at Icy Cape, Alaska, related to

1899 earthquakes. Journal of Geophysical Research 88(B11): 9305–9313.

Jacoby, G.C., Wiles, G., and D'Arrigo, R.D. 1996. Alaskan dendroclimatic variations for the past

300 years along a north-south gradient (transect). In: Dean, J.S., Meko, D.M., and

Swetnam, T.W. (eds.). Tree Rings, Environment, and Humanity. Radiocarbon 1996:

235–248.

Jacoby, G.C., Williams, P.L., and Buckley, B.M. 1992. Tree ring correlation between

prehistoric landslides and abrupt tectonic events in Seattle, Washington. Science 258

(5088): 1621–1623.

435
Jacoby, G.C., Workman, K.W., and D'Arrigo, R.D. 1999. Laki eruption of 1783, tree rings, and

disaster for northwest Alaska Inuit. Quaternary Science Reviews 18: 1365-1371.

Jagels, R. and Telewski, F.W. 1990. Computer-aided image analysis of tree rings. In: Cook, E.R.

and Kairiukstis, L.A. (eds.) Methods of Dendrochronology: Applications in the

Environmental Sciences. International Institute for Applied Systems Analysis, Kluwer

Academic Publishers, Boston, MA: 76–93.

Jain, S., Woodhouse, C.A., Hoerling, M.P. 2002. Multidecadal streamflow regimes in the interior

western United States: Implications for the vulnerability of water resources. Geophysical

Research Letters 29(21): 321–324.

Jansma, E. 1996. An 11,000-year tree-ring chronology of oak from the Dutch coastal region

(2258–1141 B.C.). In: Dean, J.S., Meko, D.M., and Swetnam, T.W. (eds.). Tree rings,

environment, and humanity: proceedings of the international conference, Tucson,

Arizona, 17–21 May 1994. Tucson, Arizona, Radiocarbon. pp. 769–778.

Jansma, E., Hanraets, E., and Vernimmen, T. 2004. Tree-ring research on Dutch and Flemish art

and furniture. In: E. Jansma, A. Bräuning, H. Gärtner, and G. Schleser, eds., Tree Rings

in Archaeology, Climatology and Ecology, Volume 2. Proceedings of the

Dendrosymposium 2003. Schriften des Forschungszentrum Jülich, Reihe Umwelt 44:

139-146.

Jedrysek, M.O., Krapiec, M., Skrzypek, G., Kaluzny, A., and Halas, S. 1998. An attempt to

calibrate carbon and hydrogen isotope ratios in oak tree rings cellulose: the last

millennium. RMZ Materials and Geoenvironment 45: 82–90.

Jenkins, S.E., R. Guyette, A. J. Rebertus. 1997. Vegetation-site relationships and fire history of

a savanna-glade-woodland mosaic in the Ozarks. In: Pallardy, S.G., Cecich, R.A.,

436
Garrett, H.E., Johnson,Proceedings, P.S. (eds.) 11th central hardwood forest conference,

1997 March 23–26, Columbia Missouri. USDA Forest Service. General Technical

Report NC-188, pp. 184–201.

Johnson, E.A. and Gutsell, S.L. 1994. Fire frequency models, methods and interpretations.

Advances in Ecological Research 25: 239-287.

Johnson, W.C. 1980. Dendrochronological sampling of Pinus oocarpa Shiede near Copan,

Honduras: A preliminary note. Biotropica 12(4): 315–316.

Jones, P.D., Briffa, K.R., Schweingruber, F.H. 1995. Tree-ring evidence of the widespread

effects of explosive volcanic eruptions. Geophysical Research Letters 22(11): 1333–

1336.

Jozsa, L. 1988. Increment core sampling techniques for high quality cores. Forintek Canada

Corporation Special Publication No. SP-30. 26pp.

Kaennel, M. and Schweingruber, F.H. 1995. Multilingual Glossary of Dendrochronology:

Terms and Definitions in English, German, French, Spanish, Italian, Portuguese, and

Russian. Paul Haupt Publishers. Berne. 467pp.

Kairiukstis, L. and Shiyatov, S. 1990. Dendrochronology in the USSR. In: Cook, E.R., and

Kairiukstis, L.A. (eds.). Methods of Dendrochronology: Applications in the

Environmental Sciences. pp 11–13.

Kapteyn, J.C. 1914. Tree-growth and meteorological factors. Rec. Trav. Bot. Neerl. 11: 70–93.

Kaye, M.W., T.W. Swetnam. 1999. An assessment of fire, climate, and Apache history in the

Sacramento Mountains, New Mexico. Physical Geography 20:305–330.

Kelly, P.E., Cook, E.R., and Larson, D.W. 1992. Constrained growth, cambial mortality, and

dendrochronology of ancient Thuja occidentalis on cliffs of the Niagara Escarpment: An

437
eastern version of Bristlecone pine? International Journal of Plant Science 153(1): 117–

127.

Kelly, P.E. and Larson, D.W. 1997. Dendroecological analysis of the population dynamics of an

old-growth forest on cliff-faces of the Niagara Escarpment, Canada. Journal of Ecology

85: 467–478.

Kemp, M. and Walker, M. 2001. Leonardo on Painting. An Anthology of Writings by

Leonardo da Vinci; With a Selection of Documents Relating to his Career as an Artist.

Yale University Press. 336p.

Kienast, F. 1982. Analytical investigations based on annual tree rings in damaged forest areas

of the Valais (Rhone Valley) endangered by pollution. In German with English

summary. Geographica Helvetica 3: 143–148.

Kitagawa, H., Matsumoto, E., 1995. Climatic implications of d13C variations in a Japanese cedar

(Cryptomeria japonica) during the last two millenia. Geophysical Research Letters 22,

2155–2158.

Kitzberger, T., Veblen, T.T., Villalba, R. 1995. Tectonic influences on tree growth in northern

Patagonia, Argentina: the roles of substrate stability and climatic variation. Canadian

Journal of Forest Research 25: 1684–1696.

Kneeshaw, D.D. and Bergeron, Y. 1998. Canopy gap characteristics and tree replacement in the

southeastern Boreal forest. Ecology 79(3): 783–794.

Koerber, T.W. and Wickman, B.E. 1970. Use of tree-ring measurements to evaluate impact of

insect defoliation. In: Smith, J. and Worrall J., eds., Tree-ring analysis with special

references to Northwestern America. University of British Columbia Faculty Forest

Bulletin No. 7:101–106.

438
Kozlowski, T.T. and Pallardy, S.G. 1997. Physiology of Woody Plants. Second Edition.

Academic Press, San Diego. 411pp.

Krapiec, M. 1996. Subfossil oak chronology (474 B.C. – A.D. 1529) from southern Poland. In:

Dean, J.S., Meko, D.M., and Swetnam, T.W. (eds.). Tree rings, environment, and

humanity: proceedings of the international conference, Tucson, Arizona, 17–21 May

1994. Tucson, Arizona, Radiocarbon. pp. 813–819.

Krause, C. and Eckstein, D. 1993. Dendrochronology of roots. Dendrochronologia, 11: 9–23.

Krause, C., and Gagnon, R. 2006. The relationship between site and tree characteristics and the

presence of wet heartwood in black spruce in the boreal forest of Quebec, Canada.

Canadian Journal of Forest Research 36: 1519–1526.

Krause, C. and Morin, H. 1999. Tree-ring patterns in stems and root systems of black spruce

(Picea mariana) caused by spruce budworms. Canadian Journal of Forest Research

29(10): 1583–1591.

Krause, C. and Morin, H. 2005. Adventive-root development in mature black spruce and

balsam fire in the boreal forests of Quebec, Canada. Canadian Journal of Forest Research

25: 2642–2654.

Krishnamurthy, R.V. 1996. Implications of a 400 year tree ring based 13C/12C chronology.

Geophysical Research Letters 23, 371–374.

Krishnamurthy, R.V. and Epstein, S. 1985. Treering D/H ratio from Kenya, East Africa and its

palaeoclimatic significance. Nature 317: 160–162.

Kromer, B. and Becker, B. 1993. German oak and pine 14C calibration, 7200-9439 BC.

Radiocarbon 35(1): 125-135.

439
Kromer, B. and Spurk, M. 1998. Revision and tentative extension of the tree-ring based 14C

calibration, 9200-11,855 cal BP. Radiocarbon 40(3): 1117-1125.

Krueger, K.W. and Trappe, J.M. 1967. Food reserves and seasonal growth of Douglas-fir

seedlings. Forest Science 13: 192-202.

Kuechler, J. 1859. Das Klima von Texas. Texas Staats-Zeitung. August 6. 1859, p.2. San

Antonio. (translated and reprinted in: Campbell, T. 1949. The pioneer tree-ring work of

Jacob Kuechler. Tree-Ring Bulletin 15: 16–19.

Kulakowski, D. and Veblen, T.T. 2002. Influences of fire history and topography on the pattern

of a severe wind blowdown in a Colorado subalpine forest. Journal of Ecology 90: 806–

819.

Kulakowski, D., Veblen, T.T., and Bebi, P. 2003. Effects of fire and spruce beetle outbreak

legacies on the disturbance regime of a subalpine forest in Colorado. Journal of

Biogeography 30: 1445–1456.

Kulman, H.M. 1971. Effects of insect defoliation on growth and mortality of trees. Annual

Review of Entomology 16:289–324.

Kuniholm, P.I. 2001. Dendrochronology and other applications of tree-ring studies on

archaeology. In: Browthwell, D.R. and Pollard, A.M. (eds.) Handbook of Archaeological

Sciences. John Wiley, New York. 35–46.

Kuniholm, P.I. 2003. Aegean Dendrochronology project December 2003 progress report.

Ithaca, New York, The Malcolm and Carolyn Wiener Laboratory for Aegean and Near

Eastern Dendrochronology, Cornell University.

440
LaMarche, V.C. Jr. 1968. Rates of slope degradation as determined from botanical evidence

White Mountains, California. United States Geological Survey Professional Paper 352–I.

45 pp.

LaMarche, V.C. Jr. 1973. Holocene climatic variations inferred from treeline fluctuations in the

White Mountains, California. Quaternary Research 3: 632–660.

LaMarche, V.C. Jr. 1974. Paleoclimatic inferences from long tree-ring records: Intersite

comparison shows climatic anomalies that may be linked to features of the general

circulation. Science 183 (4129): 1043–1048.

LaMarche, V.C. Jr. and Fritts, H.C. 1971. Anomaly patterns of climate over the western United

States, 1700–1930, derived from principal component analysis of tree-ring data. Monthly

Weather Review 99: 138–142.

LaMarche, V.C. Jr. and Hirschboeck, K.K. 1984. Frost rings in trees as records of major

volcanic eruptions. Nature 307: 121–126.

LaMarche, Jr., V.C., Holmes, R.L., Dunwiddie, P.W., and Drew, L.G. 1979a. Tree-ring

chronologies of the southern hemisphere: 5. South Africa. Chronology Series V .

Laboratory of Tree-Ring Research, University of Arizona, Tucson, AZ: 1–27.

LaMarche, Jr., V.C., Holmes, R.L., Dunwiddie, P.W., Drew, L.G. 1979b. Tree-ring chronologies

of the southern hemisphere: Australia. Chronology Series V 4. Laboratory of Tree-Ring

Research, University of Arizona, Tucson, AZ: 1–89.

LaMarche, Jr., V.C., Holmes, R.L., Dunwiddie, P.W., Drew, L.G. 1979c. Tree-ring chronologies

of the southern hemisphere: Chile. Chronology Series V 2. Laboratory of Tree-Ring

Research, University of Arizona, Tucson, AZ: 1–43.

441
LaMarche, Jr., V.C., Holmes, R.L., Dunwiddie, P.W., Drew, L.G. 1979d. Tree-ring chronologies

of the southern hemisphere: 1. Argentina. Chronology Series V . Laboratory of Tree-Ring

Research, University of Arizona, Tucson, AZ: 1–69.

LaMarche, Jr., V.C., Holmes, R.L., Dunwiddie, P.W., Drew, L.G. 1979e. Tree-ring chronologies

of the southern hemisphere: 3. New Zealand. Chronology Series V . Laboratory of Tree-

Ring Research, University of Arizona, Tucson, AZ: 1–77.

LaMarche, V.C. Jr. and Mooney, H.A. 1967. Altithermal timberline advance in western United

States. Nature 213(5080): 980–982.

LaMarche, V.C. Jr. and Stockton, C.W. 1974. Chronologies from temperature-sensitive

bristlecone pines at upper treeline in western United States. Tree-ring Bulletin 34: 21–

45.

LaMarche, V.C. Jr., Wallace, R.E. 1972. Evaluation of effects on trees of past movements on the

San Andreas Fault, northern California. Geological Society of America Bulletin 83(9):

2665–2676.

Lambers, H., Chapin, F.S. III., and Pons, T.L. 1998. Plant physiological ecology. Springer,

New York. 540pp.

Lambert, G.N., Bernard, V., Doucerain, C., Girardclos, O., Lavier, C., Szepertisky, B., and

Trenard, Y. 1996. French regional oak chronologies spanning more than 1000 years. In:

Dean, J.S., Meko, D.M., and Swetnam, T.W. (eds.). Tree rings, environment, and

humanity: proceedings of the international conference, Tucson, Arizona, 17–21 May

1994. Tucson, Arizona, Radiocarbon. pp. 821–832.

Landres, P.B., Morgan, P., and Swanson, F.J. 1999. Overview of the use of natural variability

concepts in managing ecological systems. Ecological Applications 9(4): 1179–1188.

442
Lara, A. and Villalba, R. 1993. A 3,620-year temperature record from Fitzroya cuppressoides

tree rings in South America. Science 260: 1104–1106.

Larocque, S. J. and Smith, D. J. 2005. ‘Little Ice Age’ proxy glacier mass balance records

reconstructed from tree rings in the Mt. Waddington area, British Columbia Coast

Mountains, Canada. The Holocene 15(5): 748–757.

Larson, D.W., Matthes, U., Gerrath, J.A., Gerrath, J.M., Nekola, J.C., Walker, G.L., Porembski,

S., and Charlton, A. 1999. Ancient stunted trees on cliffs. Nature 398: 382–383.

Larson, D.W. and Melville, L. 1996. Stability of wood anatomy of living and Holocene Thuja

occidentalis L. derived from exposed and submerged portions of the Niagara Escarpment.

Quaternary Research 45(2): 210–215.

Larson, P.R. 1994. The Vascular Cambium: Development and Structure. Springer-Verlag,

Berlin, New York. 725pp.

Lavier, C., Lambert, G. 1996. Dendrochronology and works of art. In: Dean, J.S., Meko, D.M.,

and Swetnam, T.W. (eds.). Tree rings, environment, and humanity: proceedings of the

international conference, Tucson, Arizona, 17–21 May 1994. Tucson, Arizona,

Radiocarbon. pp. 343–352.

Lawrence, D.B. 1950. Estimating dates of recent glacier advances and recession rates by

studying tree growth layers. Transactions of the American Geophysical Union 31: 243–

248.

Lawrence, J.R. and White, J.W.C. 1984. Growing season precipitation from D/H ratios of

Eastern White Pine. Nature 311: 558–560.

Laxton, R.R. and Litton, C.D. 1988. An East Midlands master tree-ring chronology and its use

for dating vernacular buildings. University of Nottingham Department of Classical and

443
Archaeological Studies (Archaeology Section) Monograph Series III. Nottingham,

England.

Leavitt, S.W. 1992. Isotopes and trace elements in tree rings. LUNDQUA Report 34: 182–190.

Leavitt, S.W. 1993. Environmental information from 13C/12C ratios of wood. Geophysical

Monographs 78: 325–331.

Leavitt, S.W. and Danzer, S.R. 1993. Method for batch processing small wood samples to

holocellulose for stable-carbon isotope analysis. Analytical Chemical 65: 87–89.

Leavitt, S.W. and Lara, A. 1994. South American tree rings show declining d13C trend. Tellus

46B: 152–157.

Leavitt, S.W. and Long, A. 1984. Sampling strategy for stable carbon isotope analysis of tree

rings in pine. Nature 311: 145–147.

Leavitt, S.W. and Long, A. 1985. An atmospheric 13C/12C reconstruction generated through

removal of climate effects from tree ring 13C/12C measurements. Tellus 35B: 92–102.

LeBlanc, D.C. 1990a. Relationships between breast-height and whole-stem growth indices for

red spruce on Whiteface Mountain, New York. Canadian Journal of Forest Research 20:

1399–1407.

LeBlanc, D.C. 1990b. Red spruce decline on Whiteface Mountain, New York. I. Relationship

with elevation, tree age, and competition. Canadian Journal of Forest Research 20:1408–

1414.

LeBlanc, D.C., Raynal, D.J., and White, E.H. 1987. Acidic deposition and tree growth: 1. The

use of stem analysis to study historical growth patterns. Journal of Environmental

Quality 16(4):325–333.

444
Lehtonen, H. and Huttunen, P. 1997. History of forest fires in eastern Finland from the fifteenth

century AD — the possible effects of slash-and-burn cultivation. Holocene 7: 223–228.

Lertzman, K.P., Sutherland, G.D., Inselberg, A., and Saunders, S.C. 1996. Canopy gaps and the

landscape mosaic in a coastal temperate rain forest. Ecology 77: 1254–1270.

Lewis, E. 1873. The longevity of trees. Popular Science Monographs 3: 321–334.

Lewis, M.A. 2002. Culturally modified trees as indicators of cultural activity in Northern

Temperate Rainforests, Maxwell Center for Anthropological Research, Newsletter

Number 1: 4–5.

Libby, L.M. and Pandolfi, L.J. 1974. Temperature dependence of, isotope ratios in tree rings.

Proceedings of the National Academy of Science 71: 2482–2486.

Libby, L.M., Pandolfi, L.J., Payton, P.H., Marshall III, J., Becker, B. and Giertz-Siebenlist, V.

1976. Isotopic tree thermometers. Nature 261: 284–290.

Liese, W. 1978. Bruno Huber : The pioneer of European dendrochronology. In: Fletcher, J.

(ed.) Dendrochronology in Europe: Principles, interpretations and applications to

archaeology and history. Based on the Symposium held at the National Maritime

Museum, Greenwich, July 1977. National Maritime Museum, Greenwich,

Archaeological Series No. 4, Research Laboratory for Archaeology and History of Art,

Oxford University, Publication No. 2, British Arcaheological Reports International Series

51. 1–10 pp.

Lin, A. and Lin, S. 1998. Tree damage and surface displacement: the 1931 M 8.0 Fuyun

earthquake. Journal of Geology 106(6): 751–757.

Linnaeus, C. 1745. Olandska och Gothlandska Resa, etc. 344p.

Linnaeus, C. 1751. Skanska Resa, etc. 434p.

445
Lipp, J. and Trimborn, P. 1991. Long-term records and basic principles of tree-ring isotope data

with emphasis on local environmental conditions. Pal.aoklimaforschung 6: 105–117.

Lipp, J., Trimborn, P., Fritz, P., Moser, H., Becker, B. and Frenzel, B. 1991. Stable isotopes in

tree ring cellulose and climatic change. Tellus 43B: 322–330.

Little, E.L., Jr., 1971, Atlas of United States trees, volume 1, conifers and important hardwoods:

U.S. Department of Agriculture Miscellaneous Publication 1146, 9 p., 200 maps.

Liu, Y., Wu, X., Leavitt, S.W., and Hughes, M.K. 1996. Stable carbon isotope in tree rings from

Huangling, China and climatic variation. Science in China D(39) 2: 152–161.

Lloyd, A.H. and Graumlich, L.J. 1997. Holocene dynamics of treeline forests in the Sierra

Nevada. Ecology 78(4): 1199–1210.

Loader, N.J. and Switsur, V.R. 1996. Reconstructing past environmental change using stable

isotopes in tree-rings. Botanical Journal of Scotland 48: 65–78.

Lomolino, M.V., Riddle B.R., and Brown, J. 2006. Biogeography, Third Edition. Sinauer

Associates Inc., Sunderland MA. 845pp.

Long, A. 1982. Stable isotopes in tree rings. In Hughes, M.K., Kelly, P.M., Pilcher, J.R., and

LaMarche, Jr. V.C. (eds.) Climate from Tree Rings. Cambridge University Press,

Cambridge pp. 13–18.

Lorimer, C.G., and Frelich, L.E. 1989. A method for estimating canopy disturbance frequency

and intensity in dense temperate forests. Canadian Journal of Forest Research 19: 651–

663.

Luckman, B.H. 1988. Dating the moraines and recession of Athabasca and Dome Glaciers,

Alberta, Canada. Arctic and Alpine Research 20: 40–54.

446
Luckman, B.H. 2003. Assessment of present, past and future climate variability in the Americas

from treeline environments. IAI CRN03 Annual Report 2003.

Lynch, A. M., and Swetnam T.W. 1992. Old-growth mixed-conifer and western spruce

budworm in the southern Rocky Mountains. U.S. Department of Agriculture, Forest

Service. GTR-RM-213. pp66–80.

MacDonald, G.M. and Case, R.A. 2005. Variations in the Pacific Decadal Oscillation over the

past millennium. Geophysical Research Letters 32(8): Article Number L08703.

Madany, M.H., Swetnam, T.W. and West, N.E. 1982. Comparison of two approaches for

determining fire dates from tree scars. Forest Science 28(4): 856–861.

Malmstrom, C.M., Thompson, M.V., Juday, G., Los, S.O., Randerson, J.T., and Field, C.B.

1997. Interannual variation in global-scale net primary production: Testing model

estimates. Global Biogeochemical Cycles 11: 367–392.

Mann, M. E., Bradley, R. S., and Hughes, M. K. 1998. Global-scale temperature patterns and

climate forcing over the past six centuries. Nature 392: 779–787.

Mantua, N.J., Hare, S.R., Zhang, Y., Wallace, J.M., and Francis, R.C. 1997. A Pacific

Interdecadal Climate Oscillation with Impacts on Salmon Production. Bulletin of the

American Meteorological Society, 78, pp. 1069–1079.

Marchand, P.J. 1984. Dendrochronology of a fir wave. Canadian Journal of Forest Research

14(1): 51–56.

Mariaux, A. 1981. Past efforts in measuring age and annual growth in tropical trees. Yale

University School of Forestry and Environmental Studies Bulletin 94: 20-30.

447
Mason, R.R. and Torgerson, T.R. 1987. Dynamics of a non-outbreak population of the

Douglas-fir tussock moth (Lepidoptera: Lymantriidae) in southern Oregon.

Environmental Ecology 16:1217–1227.

Mason, R.R., Wickman, B.E., and Paul, H.G. 1997. Radial growth response of Douglas-fir and

grand fir to larval densities of the Douglas-fir tussock moth and the western spruce

budworm. Forest Science 43:194–205.

Massey, C. L. 1940. The pandora moth, a periodic pest of western pine forests. U.S.

Department of Agriculture, Forest Service. Technical Bulletin 137. 20pp.

Matthews, J.A. 1977. Glacier and climate fluctuations inferred from tree-growth variations over

the last 250 years, central Norway. Boreas 6: 1–24.

McCarroll, D. and Loader, N.J. 2004. Stable isotopes in tree rings. Quaternary Science

Reviews 23: 771–801.

McCarroll, D. and Pawellek, F. 1998. Stable carbon isotope ratios of latewood cellulose in

Pinus sylvestris from northern Finland: Variability and signal-strength. The Holocene

8(6): 675–684.

McCarroll, D. and Pawellek, F. 2001. Stable carbon isotope ratios of Pinus sylvestris from

northern Finland and the potential for extracting a climate signal from long

Fennoscandian chronologies. The Holocene 11(5): 517–526.

McCarthy, D.P. and Luckman, B.H. 1993. Estimating ecesis for tree-ring dating of moraines:

A comparative study form the Canadian Cordillera. Arctic and Alpine Research 25: 63–

68.

448
McCarthy, D.P., Luckman, B.H., and Kelly, P.E. 1991. Sampling height-age error corrections

from spruce seedlings in glacial forefields, Canadian Cordillera. Arctic and Alpine

Research 23: 451–455.

McCord, V.A.S. 1996. Fluvial process dendrogeomorphology: Reconstruction of flood events

from the southwestern United States using flood-scarred trees. In: Dean, J.S., Meko,

D.M., and Swetnam, T.W. (eds.) Tree Rings, Environment, and Humanity: Proceedings

of the International Conference, Tucson Arizona, 17–21 May 1994. Radiocarbon,

Tucson. pp. 689–699.

McCormac, F.G., Baillie, M.G.L., Pilcher, J.R., Brown, D.M., and Hoper, S.T. 1994. d13C

measurement from the Irish oak chronology. Radiocarbon 36: 27–35.

McCullough, D.G., Werner, R.A., and Neumann, D. 1998. Fire and insects in northern and

boreal forest ecosystems of North America. Annual Review of Entomology 43:107–127.

McLaren, B.E. and Peterson, R.O. 1994. Wolves, moose, and tree rings on Isle Royale. Science

266(5190): 1555–1558.

Meisling, K.E. and Sieh, K.E. 1980. Disturbance of trees by the 1857 Fort Tejon earthquake,

California. Journal of Geophysical Research 85(B6): 3225–3238.

Meko, D. M. and Baisan, C. H. 2001. Pilot study of latewood-width of confers as an indicator

of variability of summer rainfall in the north American Monsoon region. International

Journal of Climatology 21: 697-708.

Meko, D.M., Cook, E.R., Stahle, D.W., Stockton, C.W., and Hughes, M.K. 1993. Spatial

patterns of tree-growth anomalies in the United States and southeastern Canada. Journal

of Climate 6: 1773–1786.

449
Meko, D. M., Stockton, C. W., and Boggess, W. R. 1980. A tree-ring reconstruction of drought

in southern California. Water Resources Bulletin 16(4): 594-600.

Meko, D.M. and Woodhouse, C.A. In press. Chapter 8: Applications of streamflow

reconstruction to water resources management. In: Hughes, M.K., Diaz, H.F., and

Swetnam, T.W. (eds.) Dendroclimatology: Progress and Prospects. Springer Verlag

series Developments in Paleoenvironmental Research (DPER) edited by Last, W.M. and

Smol, J.P.

Miles, D.H. and Worthington, M.J. 1998. Sonora Pass junipers from California USA:

construction of a 3,500-year chronology. In Stravinskiene, V. and Juknys, R.. (eds.).

Dendrochronology and Environmental Trends - Proceedings of the International

Conference 17–21 June 1998, Kaunas, Lithuania. Vytautas Magnas University

Department of Environmental Sciences, Kaunas.

Miller, G.T.Jr. 2005. Living in the environment: Principles, Connections, and Solutions. 14th

Edition. Brooks Cole Publishing, Pacific Grove. 720pp.

Mitchell, J.M., Jr., Stockton, C.W., and Meko, D.M. 1979. Evidence of a 22-year rhythm of

drought in the western United States related to the Hale Solar Cycle since the 17th

Century. In: McCormac, B.M. and Seliga, T.A. (eds.) Solar-Terrestrial Influences on

Weather and Climate, D. Reidel Publishing Company, Dordrecht, Holland., p. 125-143.

Moberg, A., Sonechkin, D.M., Holmgren, K., Datsenko, N.M., and Karlén, W. 2005. Highly

variable Northern Hemisphere temperatures reconstructed from low- and high-resolution

proxy data. Nature 433: 613–617.

Mobley, C.M. and Eldridge, M. 1992. Culturally modified trees in the Pacific Northwest. Arctic

Anthropology 29(2): 91–110.

450
Mora, C.I., Miller, D.L., and Grissino-Mayer, H.D. 2006. Tempest in a tree ring:

Paleotempestology and the record of past hurricanes. The Sedimentary Record 4(3): 4–8.

Morgan, P., Aplet, G.H., Haufler, J.B., Humphries, H.C., Margaret, M.M., and Wilson, W.D.

1994. Historical range of variability: A useful tool for evaluating ecosystem change. In:

Sampson, R.L. and Adams, D.L., eds., Assessing forest ecosystem health in the inland

west. Proceedings of the American Forests scientific workshop. The Hawthorn Press, Inc.

New York. 87–111.

Mott D. G., Nairn, L.D., and Cook, J.A. 1957. Radial growth in forest trees and effects of insect

defoliation. Forest Science 3(3):286–304.

Motyka, R.J. 2003. Little Ice Age subsidence and post Little Ice Age uplift at Juneau, Alaska,

inferred from dendrochronology and geomorphology. Quaternary Research 59(3): 300-

309.

Muir, J. 1911. My first summer in the Sierra. Houghton Mifflin Company. First Edition.

Munro, M.A.R., Brown, P.M., Hughes, M.K., and Garcia, E.M.R. 1996. Image analysis of

tracheid dimensions for dendrochronological use. In: Dean, J.S., Meko, D.M. and

Swetnam, T.W. (eds.) Tree Rings, Environment, and Humanity. Radiocarbon 1996: 843–

851.

Nash, S.E. 1997. A cutting-date estimation technique for ponderosa pine and Douglas fir wood

specimens. American Antiquity 62(2): 260–272.

Nash, S. E. 1999. Time, Trees, and Prehistory: Tree-ring dating and the development of North

American archaeology, 1914–1950. The University of Utah Press, Salt Lake City.

294pp.

451
Nials, F., Gregory, D., Graybill, D. 1989. Salt River streamflow and Hohokam irrigation

systems. In: Heathington, C., Gregory, G. The 1982–1984 excavations at Las Colinas:

Environment and subsistence. Archaeological Series 162, Arizona State University,

Tempe. pp. 59–78.

Nicolussi, K., Kaufmann, M., Patzelt, G., van der Plicht, J., and Thurner, A. 2005. Holocene

tree-line variability in the Kauner Valley, Central Eastern Alps, indicated by

dendrochronological analysis of living trees and subfossil logs. Vegetation History and

Archaeobotany 14(3): 221-234.

Niklasson, M. and Granström, A. 2000. Numbers and sizes of fires: Long-term spatially explicit

fire history in a Swedish Boreal landscape. Ecology 81(6): 1484–1499.

Nogler, P. 1981. Auskeilende und fehlende Jahrringe in absterbenden Tannen (Abies alba Mill.)

Allgemeine Forstzeitschrift 36(28): 709–711.

O’Neill, L.C. 1963. The suppression of growth rings in jack pine in relation to defoliation by

Swaine jack-pine sawfly. Canadian Journal of Botany 41:227–235.

Okada, N., Fujiwara, T., Ohta, S. and Matsumoto, E. 1995. Stable carbon isotopes of

Chamaecyparis obtusa grown at a high altitude region in Japan: within and among-tree

variations. In: Ohta, S., Fujii, T., Okada, N., Hughes, M.K., Eckstein, D. (Eds.), Tree-

Rings: From the Past to the Future. Proceedings of the International Workshop on Asian

and Pacific Dendrochronology. Forestry and Forest Products Research Institute Scientific

Meeting Report 1, pp. 165–169.

Orvis, K. H. and Grissino-Mayer, H. D. 2002. Standardizing the reporting of abrasive papers

used to surface tree-ring samples. Tree-Ring Research: 58(1): 47–50.

452
Orwig, D.A., Cogbill, C.V., Foster, D.R., and O’Keefe, J.F. 2001. Variations in old-growth

structure and definitions: Forest dynamics on Wachusett Mountain, Massachusetts.

Ecological Applications 11(2): 437–452.

Page, R. 1970. Dating episodes of faulting from tree rings. Effects of the 1958 rupture of the

Fairweather fault on tree growth. Geological Society of America Bulletin 81: 3085–3094.

Panshin, A.J., De Zeeuw, C., and Brown, H.P. 1964. Textbook of wood technology. Volume I:

Structure, identification, uses, and properties of the commercial woods of the United

States. McGraw-Hill Book Company. New York. 643p.

Park, W.-K. and Telewski, F.W. 1993. Measuring maximum latewood density by image analysis

at the cellular level. Wood and Fiber Science 25(4): 326–332.

Patterson, J. E. 1929. The pandora moth, a periodic pest of the western pine forests. U.S.

Department of Agriculture, Forest Service. Technical Bulletin 137. 20pp.

Payette, S. 1987. Recent porcupine expansion at tree line; a dendro-ecological analysis.

Canadian Journal of Zoology 65: 551–557.

Payette, S., Boudreau, S., Morneau, C., and Pitre, N. 2004. Long-term interactions between

migratory caribou, wildfires and nunavik hunters inferred from tree rings. Ambio 33(8):

482-486.

Payette, S., Morneau, C., Sirois, L., and Desponts, M. 1989. Recent fire history of the northern

Québec biomes. Ecology 70(3): 656–673.

Pearman, G.I., Francey, R.J., and Fraser, P.J.B. 1976. Climatic implications of stable carbon

isotopes in tree-rings. Nature 260: 771–773.

453
Pederson, N., Cook, E.R., Jacoby, G.C., Peteet, D.M., and Griffin, K.L. 2004. The influence of

winter temperatures on the annual radial growth of six northern range margin tree

species. Dendrochronologia 22(1): 7–29.

Pederson, N.A., Jones, R.H., Sharitz, R.R. 1997. Age structure and possible origins of old Pinus

taeda stands in a floodplain forest. Journal of the Torrey Botanical Society 124(2): 111–

123.

Perkins, D.L. and Swetnam, T.W. 1996. A dendroecological assessment of whitebark pine in the

Sawtooth - Salmon River region, Idaho. Canadian Journal of Forest Research 26(12):

2123–2133.

Pendall, E. 2000. Influence of precipitation seasonality on Pin˜ on pine cellulose dD values.

Global Change Biology 6: 287–301.

Phipps, R. L. 1985. Collecting, preparing, crossdating, and measuring tree increment cores.

U.S. Geological Survey. Water Resources Investigations Report 85–4148. 48p.

Phipps, R.L. 2005. Some geometric constraints on ring-width trend. Tree-Ring Research 61(2):

73–76.

Phipps, R.L., Ireley, D.L., and Baker, C.P. 1979. Tree rings as indicators of hydrologic change in

the Great Dismal Swamp, Virginia and North Carolina. US Geological Survey Water

Resources Investigations Report 78-136: 1–26.

Pickett, S.T.A. and White, P.S. 1985. The ecology of natural disturbance and patch dynamics.

Academic Press. San Diego. 472 pp.

Pilcher, J.R., Baillie, M.G.L., Schmidt, B., and Becker, B. 1984. A 7,272-year tree-ring

chronology for western Europe. Nature 312: 150–152.

454
Piovesan G., Di Filippo A., Alessandrini A., Biondi F. Schirone, E.B. 2005. Structure, dynamics

and dendroecology of an old-growth Fagus forest in the Apennines. Journal of

Vegetation Science 16: 13–28

Pollens, S. 1999. Le Messie. Journal of the Violin Society of America 16(1): 77–101.

Pollens, S. 2001. Messiah redux. Journal of the Violin Society of America 17(3): 159–179.

Presnall, C.C. 1933. Fire studies in the Mariposa Grove. Yosemite Nature Notes 12(3): 23–24.

Pumijumnong, N., Eckstein, D., and Sass, U. 1995. Tree-ring research on Tectona grandis in

northern Thailand. IAWA Journal 16:385–192.

Pyne, S.J. 1982. Fire in America: A cultural history of wildland and rural fire. Princeton

University Press, Princeton.

Ramesh, R., Bhattacharya, S.K., and Gopalan, K. 1985. Dendroclimatological implications of

isotope coherence in trees from Kashmir Valley, India. Nature 317: 802–804.

Ramesh, R., Bhattacharya, S.K., and Gopalan, K. 1986. Climatic correlations in the stable

isotope records of silver fir (Abies pindrow) tree from Kashmir, India. Earth and

Planetary Science Letters 79: 66–74.

Ratzeburg J.T.C. 1866. Die Waldverderbnis, oder dauernder Schaden, welcher durch

Insektenfrass, Schälen, Schlagen und Verbeissen an lebenden Waldbäumen entsteht.

Berlin, Nicolaische Verlagsbuchhandlung, 2 vol.

Reid, M. 1989. The response of understorey vegetation to major canopy disturbance in the

subalpine forests of Colorado. Masters Thesis, University of Colorado, Boulder.

Robertson, I., Rolfe, J., Switsur, V.R., Carter, A.H.C., Hall, M.A., Barker, A.C., and

Waterhouse, J.S. 1997a. Signal strength and climate relationships in 13C/12C ratios of

455
tree ring cellulose from oak in southwest Finland. Geophysical Research Letters 24:

1487–1490.

Robertson, I., Switsur, V.R., Carter, A.H.C., Barker, A.C., Waterhouse, J.S., Briffa, K.R. and

Jones, P.D. 1997b. Signal strength and climate relationships in 13C/12C ratios of tree

ring cellulose from oak in east England. Journal of Geophysical Research 102: 19507–

19519.

Robertson, I., Waterhouse, J.S., Barker, A.C., Carter, A.H.C., and Switsur, V.R. 2001. Oxygen

isotope ratios of oak in east England: implications for reconstructing the isotopic

composition of precipitation. Earth and Planetary Science Letters 191: 21–31.

Roig, F.A., Le-Quesne, C., Boninsegna, J.A., Briffa, K.R., Lara, A., Grudd, H., Jones, P.D.,

Villagran, C. 2001. Climate variability 50,000 years ago in mid-latitude Chile as

reconstructed from tree rings. Nature 410: 567–570.

Roig, F.A., Villalba, R., and Ripalta, A. 1988. Climatic factors in Discaria trinervis growth in

Argentine Central Andes. Dendrochronologia 6: 61–70.

Rossi, S., Deslauriers, A., Anfodillo, T., Morin, H., Saracino, A., Motta, R., and Borghetti, M.

2006. Conifers in cold environments synchronize maximum growth rate of tree-ring

formation with day length. New Phytologist 170(2): 301–310.

Rubner, K. 1910. Das Hungern des Cambiums und das Aussetzen der Jahresringe. Naturw.

Zeits. Forst – u. Landw. 8: 212–262.

Ruzhich, V.V., San'kov, V.A., Dneprovskii, Y.I. 1982. The dendrochronological dating of

seismogenic ruptures in the Stanovoi Highland. Soviet Geology and Geophysics 23(8):

57–63.

456
Ryerson, D.E., Swetnam, T.W., and Lynch, A.M. 2003. A tree-ring reconstruction of western

spruce budworm outbreaks in the San Juan Mountains, Colorado, U.S.A. Canadian

Journal of Forest Research 33(6): 1010–1028.

Salisbury, F.B. and Ross, C.W. 1992. Plant Physiology. Fourth Edition. Wadsworth Publishing

Company, Belmont. 682pp.

Salzer, M.W. 2000. Dendroclimatology in the San Francisco Peaks Region of northern Arizona,

USA. PhD Dissertatin, The University of Arizona. 211pp.

Sarton, G. 1954. When was tree-ring analysis discovered?. Isis 45(4): 383–384.

Saurer, M., Robertson, I., Siegwolf, R., and Leuenberger, M. 1998b. Oxygen isotope analysis of

cellulose: an interlaboratory comparison. Analytical Chemistry 70: 2074–2080.

Saurer, M., Schweingruber, F.H., Vaganov, E.A., Shiyatov, S.G., and Siegwolf, R. 2002. Spatial

and temporal oxygen isotope trends at northern tree-line Eurasia. Geophysical Research

Letters 29: 10–14.

Saurer, M. and Siegenthaler, U. 1989. 13C/12C isotope ratios in tree rings are sensitive to

relative humidity. Dendrochronologia 7: 9–13.

Saurer, M., Siegenthaler, U., and Schweingruber, F. 1995. The climate–carbon isotope

relationship in tree rings and the significance of site conditions. Tellus 47B: 320–330.

Saurer, M., Siegwolf, R., Borella, S., and Schweingruber, F. 1998a. Environmental information

from stable isotopes in tree rings of Fagus sylvatica. In: Beniston, M., Innes, J.L. (Eds.),

The Impacts of Climate Variability on Forests. Springer, Berlin, pp. 241–253.

Savage, M. and Swetnam, T.W. 1990. Early 19th century fire decline following sheep pasturing

in a Navajo ponderosa pine forest. Ecology 71(6): 2374–2378.

457
Schiegl, W.E. 1974. Climatic significance of deuterium abundance in growth rings of Picea.

Nature 251: 582–584.

Schleser, G.H., Frielingsdorf, J., and Blair, A. 1999. Carbon isotope behaviour in wood and

cellulose during artificial aging. Chemical Geology 158: 121–130.

Schöngart, J., Junk, W.J., Piedade, M.T.F., Ayres, J.M., Huttermann, A., and Worbes, M. 2004.

Teleconnection between tree growth in the Amazonian floodplains and the El Nino-

Southern Oscillation effect. Global Change Biology 10(5): 683-692.

Schöngart, J., Orthmann, B., Hennenberg, K.J., Porembski, S., and Worbes, M. 2006. Climate-

growth relationships of tropical tree species in West Africa and their potential for climate

reconstruction. Global Change Biology 12(7): 1139-1150.

Schöngart, J., Piedade, M.T.F., Wittmann, F., Junk, W.J., and Worbes, M. 2005. Wood growth

patterns of Macrolobium acaciifolium (Benth.) Benth. (Fabaceae) in Amazonian black-

water and white-water floodplain forests. Oecologia 145(3): 454-461.

Schulman, E. 1937. Some Early Papers on Tree-Rings: J.C. Kapteyn. Tree-Ring Bulletin 3:28–

29.

Schulman, E. 1938. Nineteen centuries of rainfall history in the Southwest. Bulletin of the

American Meteorological Society 19(5): 211–216.

Schulman, E. 1954. Longevity under adversity in conifers. Science 119: 396–399.

Schulman, E. 1956. Dendroclimatic changes in semiarid America. University of Arizona Press,

Tucson, AZ, USA. 142pp.

Schulze, B., Wirth, C., Linke, P., Brand, W.A., Kuhlmann, I., Horna, V., and Schulze, E.D.

2004. Laser ablation-combustion-GC-IRMS - a new method for online analysis of intra-

annual variation of delta C-13 in tree rings. Tree Physiology 24(11): 1193–1201.

458
Schweingruber, F. H. 1988. Tree Rings: Basics and Applications of Dendrochronology. D.

Reidel Publishing Company Dordrecht. 276pp.

Schweingruber, F. H. 1996. Tree Rings and Environment - Dendroecology. Haupt Press.

Berne. 609p.

Seager, R., M. Ting, I. Held, Y. Kushnir, J. Lu, G. Vecchi, H.-P. Huang, N. Harnik, A. Leetmaa,

N.-C. Lau, C. Li, J. Velez, and N. Naik. 2007. Model projections of an imminent

transition to a more arid climate in southwestern North America. Science 316(5828):

1181-1184.

Seckendorff A.F., 1881. Beiträge zur Kenntnis der Schwarzföhre. Mitteilung aus dem forstlichen

Versuchswesen Oesterreichs. Carl Gerold Verlag, Wien, 66 pp.

Sellards, E.H., Tharp, B.C., and Hill, R.T. 1923. Investigation on the Red River made in

connection with the Oklahoma-Texas boundary suit. University of Texas Bulletin No.

2327. 172pp.

Shah, S.K., Bhattacharyya, A., and Chaudhary, V. 2007. Reconstruction of June–September

precipitation based on tree-ring data of teak (Tectona grandis L.) from Hoshangabad,

Madhya Pradesh, India. Dendrochronologia 25(1): 57-64.

Shao, X.M., Wang, S.Z., Xu, Y., Zhu, H.F., Xu, X.G., and Xiao, Y.M. 2007. A 3500-year master

tree-ring dating chronology from the northeastern part of the Qaidam Basin. Quaternary

Sciences 27: 477–485.

Sheppard, P.R. and Graumlich, L.J. 1996. A reflected-light video imaging system for tree-ring

analysis of conifers. In: Dean, J.S., Meko, D.M., and Swetnam, T.W. (eds.) Tree rings,

environment, and humanity: Proceedings of the International Conference, 17–21 May

1994, Department of Geosciences, University of Arizona, Tucson. pp 879–889.

459
Sheppard, P.R., Graumlich, L.J., and Conkey, L.E. 1996. Reflected-light image analysis of

conifer tree rings for reconstructing climate. The Holocene 6(1): 62–68.

Sheppard, P.R. and Jacoby, G.C. 1989. Application of tree-ring analysis to paleoseismology:

Two case studies. Geology 17: 226–229.

Sheppard, P.R., and White, L.O. 1995. Tree-ring responses to the 1978 earthquake at Stephens

Pass, northeastern California. Geology 23(2): 109–112.

Sheppard, P.S. and Witten, M.L., 2005. Laser trimming tree-ring cores for dendrochemistry of

metals. Tree-Ring Research 62: 87-92.

Sheppard, P.R. and Wiedenhoeft, A. 2007. An advance ment in removing extraneous color from

wood for low-magnification reflected-light image analysis of conifer tree rings. Wood

and Fiber Science 39(1): 173–183.

Sherzer, W.H. 1905. Glacial studies in the Canadian Rockies and Selkirks. Smithsonian

Miscellaneous Collections 47: 453–496.

Sheu, D.D., Kou, P., Chiu, C.-H., and Chen, M.-J. 1996. Variability of tree-ring d13C in Taiwan

fir: growth effect and response to May–October temperatures. Geochimica et

Cosmochimica Acta 60: 171–177.

Shinn, D. A. 1978. Man and the land: an ecological history of fire and grazing on eastern

Oregon rangelands. M.A. thesis, Oregon Sate University, Corvallis.

Shore, T.L., Safranyik, L., Hawkes, B.C., and Taylor, S.W. 2006. Effects of the mountain pine

beetle on lodgepole pine stand structure and dynamics. In: Safranyik, L. and Wilson, B.

(eds.) The Mountain Pine Beetle: A sysntehsis of biology, management, and impacts on

lodgepole pine. Canadian Forest Service. 95–114 pp.

460
Show, S.B., Kotok, E.I. 1924. The role of fire in the California pine forests. USDA Department

Bulletin 1924: 1–80.

Shroder, J.F. Jr. 1978. Dendrogeomorphological analysis of mass movement on Table Cliffs

Plateau, Utah. Quaternary Research 9: 168–185.

Shroder, J.F. Jr. 1980. Dendrogeomorphology: Review and new techniques of tree ring dating.

Progress in Physical Geography 4: 161–188.

Shroder, J.F., Jr. and Butler, D.R. 1987. Tree-ring analysis in the Earth Sciences. In: Jacoby,

G.C. and Hornbeck, J.W. (eds.) Proceedings of the International Symposium on

Ecological Aspects of Tree-Ring Analysis. August 17–21, 1986. Marymount College,

Tarrytown, New York. Conf-8608144. pp. 186–212.

Shvedov, F. 1892. The tree as a chronicle of droughts. Meteorological Herald 5: 163–178 (in

Russian).

Sibold, J.S., Veblen, T.T., Chipko, K., Lawson, L., Mathis, E., and Scott, J. 2007. Influences of

secondary disturbances on lodgepole pine stand development in Rocky Mountain

National Park. Ecological Applications 17(6): 1638-1655.

Sibold, J.S., Veblen, T.T. and Gonzalez, M.E. 2006. Spatial and temporal variation in historic

fire regimes in subalpine forests across the Colorado Front Range in Rocky Mountain

National Park, Colorado, USA. Journal of Biogeography 33(4): 631-647.

Siegenthaler, U. 1979. Stable hydrogen and oxygen isotopes in the water cycle. In: Jäger, E.

and Hunziker, J.C. (eds.). Lectures in Isotope Geology Springer-Verlag, Berlin. pp. 264–

273

Sigafoos, R.S. 1964. Botanical evidence of floods and floodplain deposits. United States

Geological Survey. Professional Paper 485-A: 1–35.

461
Sigafoos, R.S. and Hendricks, E.L. 1961. Botanical evidence of the modern history of Nisqually

Glacier, Washington. United States Geological Survey Professional Paper 387-A.

Sigafoos, R.S. and Hendricks, E.L. 1972. Recent activity of glaciers of Mount Rainier,

Washington. United States Geological Survey Professional Paper 387-B.

Smiley, T.L. 1958. The geology and dating of Sunset Crater, Flagstaff, Arizona. In: Anderson,

R.Y. and Harshbarger, J.W. (eds.) Guidebook of the Black Mesa Basin, Northwestern

Arizona. Socorro, New Mexico, New Mexico Geological Society, Ninth Field

Conference: 186–190.

Smith, D.J. and Laroque, C.P. 1996. Dendroglaciological dating of a Little Ice Age glacial

advance at Moving Glacier, Vancouver Island, British Columbia. Geographie physique

et Quaternaire 50: 47–55.

Smith, D.J., and Lewis, D. 2007. Dendroglaciology. Encyclopedia of Quaternary Science. Edited

by: S.A. Elias. Elsevier Scientific. Volume 2: 986–994.

Smith, K.T. and Sutherland, E. K. 1999. Fire-scar formation and compartmentalization in oak.

Canadian Journal of Forest Research 29(2): 166–171.

Smith, K.T., and Sutherland, E.K. 2001. Terminology and biology of fire scars in selected central

hardwoods. Tree-Ring Research 57(2): 141–147.

Solomon, S., D. Qin, M. Manning, R.B. Alley, T. Berntsen, N.L. Bindoff, Z. Chen, A.

Chidthaisong, J.M. Gregory, G.C. Hegerl, M. Heimann, B. Hewitson, B.J. Hoskins, F.

Joos, J. Jouzel, V. Kattsov, U. Lohmann, T. Matsuno, M. Molina, N. Nicholls, J.

Overpeck, G. Raga, V. Ramaswamy, J. Ren, M. Rusticucci, R. Somerville, T.F. Stocker,

P. Whetton, R.A. Wood and D. Wratt. 2007. Technical Summary. In: Solomon, S., D.

Qin, M. Manning, Z. Chen, M. Marquis, K.B. Averyt, M. Tignor and H.L. Miller (eds.).

462
2007. Climate Change 2007: The Physical Science Basis. Contribution of Working

Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate

Change. Cambridge University Press, Cambridge, United Kingdom and New York, NY,

USA.

Sonninen, E. and Jungner, H. 1995. Stable carbon isotopes in tree-rings of a Scots pine (Pinus

sylvestris L.) from northern Finland. Palaeoklimaforschung 15: 121–128

Speer, J.H. 1997. A Dendrochronological Record of Pandora Moth (Coloradia pandora, Blake)

Outbreaks in Central Oregon. MS Thesis. The University of Arizona. 159 pp.

Speer, J.H. 2001. Oak mast history from dendrochronology: A new technique demonstrated in

the southern Appalachian region. PhD Dissertation. The University of Tennessee,

Knoxville. 241 pp.

Speer, J.H. 2006. Experiential learning and exploratory research: The 13th Annual North

American Dendroecological Fieldweek (NADEF). Indiana State University, Department

of Geography, Geology, and Anthropology, Professional Paper Series No. 23.

Speer, J.H. and Hansen-Speer, K.H. 2007. Ecological applications of dendrochronology in

archaeology. Journal of Ethnobiology 27(1): 88–109.

Speer, J.H. and Holmes, R. 2004. Stem analysis on four ponderosa pine trees affected by

repeated pandora moth defoliation in central Oregon. Tree-Ring Research 60(2): 69–76.

Speer, J.H., Orvis, K.H., Grissino-Mayer, H.D., Kennedy, L.M., Horn, S.P. 2004. Assessing the

dendrochronological potential of Pinus occidentalis Swartz in the Cordillera Central of

the Dominican Republic. The Holocene 14(4): 563–569.

Speer, J.H., Swetnam, T. W, Wickman, B.E., Youngblood, A. 2001. Changes in pandora moth

outbreak dynamics during the past 622 years. Ecology 82: 679–697.

463
Spencer, D.A. 1964. Porcupine population fluctuations in past centuries revealed by

dendrochronology. The Journal of Applied Ecology 1(1): 127–149.

St. George, S. and Nielsen, E. 2002. Hydroclimatic change in southern Manitoba since A.D.

1409 inferred from tree rings. Quaternary Research 58(2): 103-111.

Stahle, D.W. 1979. Tree-ring dating of historic buildings in Arkansas. Tree-Ring Bulletin 39: 1–

28.

Stahle, D.W. 1996. Tree rings and ancient forest history. In: Eastern Old-Growth Forests, edited

by M.B. Davis, Island Press, Washington D.C., pp. 321–343.

Stahle, D.W. 1999. Effective strategies for the development of tropical tree-ring chronologies.

International Association of Wood Anatomists (IAWA) Journal 20: 249–253.

Stahle, D.W. and Cleaveland, M.K. 1993. Southern oscillation extremes reconstructed from tree

rings of the Sierra Madre Occidental and southern Great Plains. Journal of Climate 6:

129–140.

Stahle, D.W., Cleaveland, M.K., Blanton, D.B., Therrell, M.D., and Gay, D.A. 1998a. The lost

colony and Jamestown droughts. Science 280: 564–567.

Stahle, D.W., Cleaveland, M.K. and Hehr, J.G. 1988. North Carolina climate changes

reconstructed from tree rings: A.D. 372–1985. Science 240: 1517–1519.

Stahle, D.W., Cook, E.R., Cleaveland, M.K., Therrell, M.D., Meko, D.M., Grissino-Mayer,

H.D., Watson, E., and Luckman, B.H. 2000. Tree-ring data document 16th century

megadrought over North America, Eos, Transactions of the American Geophysical Union

81(12): 212,125. Reprinted by AGU in Earth in Space, March 2000.

Stahle, D.W., D'Arrigo, R.D., Krusic, P.J., Cleaveland, M.K., Cook, E.R., Allan, R.J., Cole, J.E.,

Dunbar, R.B., Therrell, M.D., Gay, D.A., Moore, M.D., Stokes, M.A., Burns, B.T.,

464
Villanueva-Diaz, J., and Thompson, L.G. 1998b. Experimental dendroclimatic

reconstruction of the Southern Oscillation. Bulletin of the American Meteorological

Society 79: 2137–2152.

Stahle, D.W., Mushove, P.T., Cleaveland, M.K., Roig, F., and Haynes, G.A. 1999. Management

implications of annual growth rings in Pterocarpus angolensis from Zimbabwe. Forest

Ecology and Management 124:217–229.

Stahle, D.W., Van Arsdale, R.B. and Cleaveland, M.K. 1992. Tectonic signal in baldcypress

trees at Reelfoot Lake, Tennessee. Seismological Research Letters 63(3): 439–447.

Stallings, W.S. Jr. 1937. Some Early Papers on Tree-Rings: J. Kuechler. Tree-Ring Bulletin 3:

27–28.

Stallings, W.S. Jr. 1949. Dating prehistoric ruins by tree-rings. Revised Edition. Laboratory of

Tree-Ring Research. Tucson.

Stewart, O.C. 1936. Cultural element distributions: XIV Northern Paiute. Anthropological

Records 4.

Stockton, C.W. and Jacboy, G.C. 1976. Long-term surface-water supply and streamflow trends

in the Upper Colorado River Basin. Lake Powell Research Project Bulletin 18. 79 pp.

Stockton, C. W. and Meko, D.M. 1975. A long-term history of drought occurrence in western

United States as inferred from tree rings, Weatherwise 28(6): 245-249.

Stoeckhardt, A. 1871. Untersuchungen über die schädliche Wirkung des Hütten- und

Steinkohlenrauches auf das Wachstum der Pflanzen, insbesondere der Fichte und Tanne.

Tharandter Forstliches Jahrbuch 21: 218–254.

465
Stoffel, M., Lievre, I., Conus, D., Grichting, M.A., Raetzo, H., Gartner, H.W., Monbaron, M.

2005. 400 years of debris-flow activity and triggering weather conditions: Ritigraben,

Valais, Switzerland . Arctic, Antarctic, and Alpine Research 37(3): 387–395.

Stokes, M. A. and Smiley, T. L. 1968. An introduction to tree-ring dating. The University of

Chicago Press. Tucson. 73p.

Stokes, M.A. and Smiley, T.L. 1996. An introduction to tree-ring dating. The University of

Arizona Press. Tucson, 73p.

Stuart J.D., Agee J.K., and Gara R.I. 1989. Lodgepole pine regeneration in an old, self-

perpetuating forest in south Oregon. Canadian Journal of Forest Research 19:1096–104

Studhalter, R.A. 1955. Tree growth: Some historical chapters. The Botanical Review 21(1–3):

1–72.

Studhalter, R.A. 1956. Early history of crossdating. Tree-Ring Bulletin 21(1–4): 31–35.

Stuiver, M. and Braziunas, T.F. 1987. Tree cellulose 13C/12C isotope ratios and climate change.

Nature 328: 58–60.

Stuiver, M., Kromer, B., Becker, B., and Ferguson, C.W. 1986. Radiocarbon age calibration back

to 13,300 years BP and the 14C age matching of the German oak and US bristlecone pine

chronologies. Radiocarbon 28(2B): 969-979.

Sutherland, E.K. 1997. History of fire in a southern Ohio second-growth mixed-oak forest. .

In: Pallardy, S.G., Cecich, R.A., Garrett, H.E., Johnson,Proceedings, P.S. (eds.) 11th

central hardwood forest conference, 1997 March 23–26, Columbia Missouri. USDA

Forest Service. General Technical Report NC-188, pp. 172–183.

Swanson, F.J., Jones, J.A., Wallin, D.O., and Cissel, J.H. 1994. Natural variability-implications

for ecosystem management. In: Jensen, M.E. and Bourgeron, P.S. tech., eds., Volume II:

466
Ecosystem management: principles and applications. Gen. Tech. Rep. PNW-GTR-318.

Portland, OR: U.S. Department of Agriculture, Forest Service, Pacific Northwest

Research Station: 89–103.

Swetnam, T.W. 1984. Peeled ponderosa pine trees: a record of inner bark utilization by Native

Americans. Journal of Ethnobiology 4(2): 177–190.

Swetnam, T. W. 1990. Fire History and climate in the Southwestern United States. In

Proceedings of Symposium on Effects of Fire in Management of Southwestern U. S.

Natural Resources, November 15–17, 1988, Tucson, Arizona, ed. J. S. Krammes, , pp. 6–

17. General Technical Report, RM-GTR-191, USDA Forest Service.

Swetnam, T.W., Allen, C.D., and Betancourt, J.L. 1999. Applied historical ecology: Using the

past to manage for the future. Ecological Applications 9(4): 1189–1206.

Swetnam, T.W. and Baisan, C. H. 1996. Historical fire regime patterns in the southwestern

United States since A.D. 1700. In: Allen, C.D. (ed.) Fire effects in southwestern forests:

proceedings of the second La Mesa fire symposium. Los Alamos, New Mexico. March

29–31, 1994. U.S. Department of Agriculture, Forest Service. General Technical Report

RM-GTR-286: 11–32.

Swetnam, T. W. and Betancourt, J.L. 1990. Fire-southern oscillation relations in the

southwestern United States. Science 249:1017–1020.

Swetnam, T. W. and Lynch, A. M. 1989. A Tree-Ring reconstruction of western spruce

budworm outbreaks in the Southern Rocky Mountains. Forest Science 35(4):962–986.

Swetnam, T. W., and Lynch, A. M. 1993. Multi-century, regional-scale patterns of western

spruce budworm history. Ecological Monographs 63(4):399–424.

467
Swetnam, T. W., Thompson, M. A., and Sutherland, E. K. 1985. Using dendrochronology to

measure radial growth of defoliated trees. USDA Forest Service, Agriculture Handbook

639. 39 p.

Swetnam, T. W., Wickman, B. E., Paul, H. G., and Baisan, C. H. 1995. Historical patterns of

western spruce budworm and Douglas-fir tussock moth outbreaks in the Northern Blue

Mountains, Oregon since A.D. 1700. Research Paper PNW-RP-484. Portland, OR: U. S.

Department of Agriculture, Forest Service, Pacific Northwest Research Station. 27 p.

Switsur, V.R., Waterhouse, J.S., Field, E.M., and Carter, A.H.C. 1996. Climatic signals from

stable isotopes in oak tree-rings from East Anglia, Great Britain. In: Dean, J.S., Meko,

D.,M., Swetnam, T.W. (Eds.), Tree Rings, Environment and Humanity Radiocarbon,

Radiocarbon, Arizona. pp. 637–645.

Switsur, V.R., Waterhouse, J.S., Field, E.M.F., Carter, A.H.C., Hall, M., Pollard, M., Robertson,

I., Pilcher, J.R., and Heaton, T.H.E. 1994. Stable isotope studies of oak from the English

Fenland and Northern Ireland. In: Funnell, B.M., Kay, R.L.F. (Eds.), Palaeoclimate of the

Last Glacial/Interglacial Cycle. Natural Environment Research Council Special

Publication 94/2: 67–73.

Tang, K., Feng, X., and Ettle, G.J. 2000. The variations in dD of tree rings and the implications

for climatic reconstruction. Geochimica et Cosmochimica Acta 64: 1663–1673.

Tang, K., Feng, X., and Funkhouser, G. 1999. The d13C of tree rings in full-bark and strip-bark

bristle cone pine trees in the White Mountains of California. Global Change Biology 5:

33–40.

Tans, P., and Mook, W.G. 1980. Past atmospheric CO2 levels and the 13C/12C ratios in tree

rings. Tellus 32: 268–283.

468
Tardif, J. and Bergeron, Y. 1999. Population dynamics of Fraxinus nigra in response to flood-

level variations, in northwestern Quebec. Ecological Monographs 69(1): 107–125.

Tardif, J.C., Conciatori, F., Nantel, P., Gagnon, D. 2006. Radial growth and climate responses of

white oak (Quercus alba) and northern red oak (Quercus rubra) at the northern

distribution limit of white oak in Quebec, Canada. Journal of Biogeography 33(9): 1657–

1669.

Tarr, R.S. and Martin, L. 1914. Alaskan glacier studies of the National Geographic Society in

the Yakutat Bay, Prince William Sound and Lower Copper River Regions. National

Geographic Society. 498pp.

Taylor, S.W., Carroll, A.L., Alfaro, R.I., and Safranyik, L. 2006. Forest, Climate, and Mountain

pine beetle outbreak dynamics in western Canada. In: Sasfranyik, L. and Wilson, B.

(eds.) The mountain pine beetle: A synthesis of biology, management, and impacts on

lodgepole pine. Canadian Forest Service, Pacific Forestry Center 67–94pp.

Telewski, F.W., Wakefield, A.H., and Mordecat, J.J. 1983. Computer-assisted image analysis

of tissues of ethrel-treated Pinus taeda seedlings. Plant Physiology 72: 177–181.

Therrell, M.D., Stahle, D.W., Ries, L.P., and Shugart, H.H. 2006. Tree-ring reconstructed rainfall

variability in Zimbabwe. Climate Dynamics 26(7-8): 677-685.

Thompson, D.R. 2005. Fine scale disturbance and stand dynamics in mature spruce-subalpine

fir forests of central British Columbia. M.S. Thesis. University of Northern British

Columbia.

Thompson, D.R., Daniels, L.D., and Lewis, K.J. 2007. A new dendroecological method to

differentiate growth responses to fine-scale disturbance from regional-scale

environmental variation. Canadian Journal of Forest Research 37: 1034-1043.

469
Topham, J. and McCormick, D. 1997. The ring saga. The Strad 108: 404–411.

Topham, J. and McCormick, D. 1998. A dendrochronological investigation of stringed

instruments of the Cremonese School (1666–1757) including "The Messiah" violin

attributed to Antonio Stradivari. Journal of Archaeological Science 27(3): 183–192.

Topham, J. and McCormick, D. 2001. The dating game. The Strad 112: 846–851.

Torrence, C. and Compo, G.P. 1998. A Practical Guide to Wavelet Analysis. Bull. Amer.

Meteor. Soc., 79: 61–78.

Touchan, R., Akkemik, U., Hughes, M.K., and Erkan, N. 2007. May-June precipitation

reconstruction of southwestern Anatolia, Turkey, during last 900 years from tree rings.

Quaternary Research 68: 196-202.

Touchan, R., Garfin, G.M., Meko, D.M., Funkhouser, G., Erkan, N., Hughes, M.K., and Wallis,

B.S. 2003. Preliminary reconstructions of spring precipitation in southwestern Turkey

from tree-ring width. International Journal of Climatology 23: 157-171.

Towner, R. H., L. Sesler, T. Hovezak. 1999. Navajo culturally modified trees in the Dinétah. In

Diné Bíkéyah: Papers in Honor of David M. Brugge, eds. M.S. Duran, D.T. Kirkpatrick,

pp. 195–209. The Archaeological Society of New Mexico 24, Albuquerque, 1999.

Treydte, K., Schleser, G.H., Helle, G., Frank, D.V., Winiger, M., Haug, G.H., and Esper, J.

2006. The twentieth century was the wettest period in northern Pakistan over the past

millennium. Nature 440: 1179–1182.

Treydte, K., Schleser, G.H., Schweingruber, F.H., Winiger, M. 2001. The climatic significance

of δ13C in subalpine spruces (Lötschental, Swiss Alps): A case study with respect to

altitude, exposure and soil moisture. Tellus. Series B, Chemical and physical

meteorology 53(5): 593–611.

470
Trouet, V., Coppin, P., and Beeckman, H. 2006. Annual growth ring patterns in Brachystegia

spiciformis reveal influence of precipitation on tree growth. Biotropica 38(3): 375-382.

Trouet, V., Haneca, K., Coppin, P., and Beeckman, H. 2001. Tree ring analysis of Brachystegia

spiciformis and Isoberlinia tomentosa: Evaluation of the ENSO-signal in the miombo

woodland of eastern Africa. IAWA Journal 22(4): 385-399.

Twining, A.C. 1833. On the growth of timber. American Journal of Science and Arts 24: 391–

393.

Vaganov, E.A., Hughes, M.K., Kirdyanov, A.V., Schweingruber, F.H., and Silkin, P.P. 1999.

Influence of snowfall and melt timing on tree growth in subarctic Eurasia. Nature 400:

149-151.

Vaganov, E.A., Hughes, M.K., Silkin, P.P., and Nesvetailo, V.D. 2004. The Tunguska event in

1908: Evidence from tree-ring anatomy. Astrobiology 4(3): 391-399.

Vaganov, E.A., Naurazhaev, M.M., Schweingruber, F.H., Briffa, K.R., and Moell, M. 1996. An

840-year tree-ring width chronology for Taimir as an indicator of summer temperature

changes. Dendrochronologia 14: 193-205.

Vale T.R. 2002. Fire, Native Peoples, and the Natural Landscape. Island Press, Washington.

Van Arsdale, R.B., Stahle, D.W., Cleaveland, M.K. and Guccione, M.J. 1998. Earthquake

signals in tree-ring data from the New Madrid seismic zone and implications for

paleoseismicity. Geology 26(6): 515–518.

van der Burgt, X.M. 1997. Determination of the age of Pinus occidentalis in La Celestina,

Dominican Republic, by the use of growth rings. International Association of Wood

Anatomists Journal 18: 139–146.

471
van West, C.R., Dean, J.S. 2000. Environmental Characteristics of the A.D. 900–1300 Period in

the Central Mesa Verde Region. Kiva 66:19–44.

Vautard, R., and Ghil, M. 1989. Singular spectrum analysis in nonlinear dynamics, with

applications to paleoclimatic time series. Physica D 35: 395–424.

Veblen, T.T., Hadley, K.S., Nel, E.M., Kitzberger, T. Reid, M., and Villabla, R. 1994.

Disturbance regime and disturbance interactions in a Rocky Mountain Subalpine Forest.

The Journal of Ecology 82(1): 125–135.

Veblen, T.T., Kitzberger, T. and Lara, A. 1992. Disturbance and forest dynamics along a transect

from Andean rain forest to Patagonian shrubland. Journal of Vegetation Science 3(4):

507–520.

Verheyden, A. 2005. Rhizophora mucronata wood as a proxy for changes in environmental

conditions. New Phytologist 167(2): 425-435.

Verheyden, A., Kairo, J.G., Beeckman, H., and Koedam, N. 2004. Growth rings, growth ring

formation and age determination in the mangrove Rhizophora mucronata. Annals of

Botany 94: 59-66.

Vetter, R.E. and Botosso, P.C. 1989. Remarks on age and growth rate determination of

Amazonian trees. IAWA Bulletin 10(2): 133-145.

Villalba, R. and Boninsegna, J.A. 1989. Dendrochronological studies of Prospois flexuosa DC.

International Association of Wood Anatomists (IAWA) Journal: 10(2): 155–160.

Villalba, R. Boninsegna, J.A., and Holmes, R.L. 1985. Cedrela angustifolia and Juglans

australis: Two new tropical species useful in dendrochronology. Tree-Ring Bulletin 45:

25–35.

472
Villalba, R., Cook, E.R., Jacoby, G.C., D’Arrigo, R.D., Veblen, T.T., and Jones, P.D. 1998a.

Tree-ring based reconstructions of northern Patagonia precipitation since A.D. 1600. The

Holocene 8(6): 659–674.

Villalba, R., Grau, H.R., Boninsegna, J.A., Jacoby, G.C., and Ripalta, A. 1998b. Tree-ring

evidence for long-term precipitation changes in subtropical South America. International

Journal of Climatology 18: 1463–1478.

Vittoz, P., Stewart, G.H., Duncan, R.P. 2001. Earthquake impacts in old-growth Nothofagus

forests in New Zealand. Journal of Vegetation Science 12(3): 417–426.

Vroblesky, D.A. and Yanosky, T.M. 1990. Use of tree-ring chemistry to document historical

ground-water contamination events. Ground Water 28(5): 677-684.

Vroblesky, D.A., Yanosky, T.M., and Siegel, F.R. 2005. Increased concentrations of potassium

in heartwood of trees in response to groundwater contamination. Environmental Geology

19(2): 71–74.

Wagner, G. 2003. Eastern Woodlands Anthropogenic Ecology. In People and Plants in Ancient

Eastern North America, ed. P.E. Minnis, pp. 126–171 Smithsonian Books, Washington.

Wallace, R.E. and LaMarche, Jr., V.C. 1979. Trees as indicators of past movements on the San

Andreas Fault. Earthquake Information Bulletin 2(4): 127–131.

Waterhouse, J.S., Barker, A.C., Carter, A.H.C., Agafonov, L.I., and Loader, N.J. 2000. Stable

carbon isotopes in Scots pine tree rings preserve a record of the flow of the river Ob.

Geophysical Research Letters 27: 3529–3532.

Watmough, S.A., Hutchinson, T.C., and Evans, D.R. 1998. The quantitative analysis of sugar

maple tree rings by laser ablation in conjunction with ICP-MS. Journal of Environmental

Quality 27(5): 1087–1094.

473
Webb, G.E. 1983. Tree rings and telescopes. The scientific career of A.E. Douglass. University

of Arizona Press, Tucson, 242p.

Webb, G.E. 1986. Solar physics and the origins of dendrochronology. Isis 77: 291–301.

Weber, U.M. and Schweingruber F. H. 1995. A dendroecological reconstruction of western

spruce budworm outbreaks (Choristoneura occidentalis) in the Front Range, Colorado,

from 1720 to 1986. Trees 9: 204–213.

Wells, A., Duncan, R.P. and Stewart, G.H. 1998. Forest dynamics in Westland, New Zealand:

the importance of large, infrequent earthquake-induced disturbance. Journal of Ecology

89(6): 1006–1018.

Welsh, C. 2007. The relationship between climate and outbreak dynamics of Dothistroma

needle blight in northwest British Columbia, Canada. M.S. Thesis. University of

Northern British Columbia. 187 pp.

Westphal, T. 2003. High-medieval urban development between the middle Elbe and the lower

Oder based on dendrochronological data. In: G. Schleser, M. Winiger, A. Bräuning, H.

Gärtner, G. Helle, E. Jansma, B. Neuwirth, and Kerstin Treydte (eds.) Tree Rings in

Archaeology, Climatology and Ecology, Volume 1: Proceedings of the

Dendrosymposium 2002. Schriften des Forschungszentrum Jülich, Reihe Umwelt 33: 20-

22.

Wickman, B.E. 1963. Mortality and growth reduction of white fir following defoliation by the

Douglas-fir tussock moth. U.S. Department of Agriculture, Forest Service Research

Paper. PSW-7. 14pp.

Wickman, B.E. 1980. Increased growth of white fire after a Douglas-fir tussock moth outbreak.

Journal of Forestry 78:31–33.

474
Wickman, B.E., Mason, R.R., and Swetnam, T.W. 1994. Searching for long-term patterns of

forest insect outbreaks. Pages 251–261 In Leather, S.R., Walters, K.F.A., Mills, N.J., and

Watt, A.D. (eds.) Individuals, populations, and patterns in ecology. Intercept, Andover,

UK.

Wigley, T. M. L., Briffa, K.R., and P.D. Jones. 1984. On the average value of correlated time

series, with applications in dendroclimatology and hyrometeorology. American

Meteorological Society 23:201–213.

Wiles, G.C., Barclay, D.J., Calkin, P.E., and Lowell, T.V. 2008. Century to millennial-scale

temperature variations for the last two thousand years indicated from glacial geologic

records of southern Alaska. Global and Planetary Change 60: 115-125.

Wiles, G.C., Calkin, P.E., and Jacoby, G.C. 1996. Tree-ring analysis and Quaternary geology:

Principles and recent applications. Geomorphology 16: 259–272.

Wiles, G.C., Post, A., Muller, E.H., and Molnia, B.F. 1999. Dendrochronology and Late

Holocene history of Bering Piedmont Glacier, Alaska. Quaternary Research 52: 185–

195.

Wilkinson, M.C. 1997. Reconstruction of historical fire regimes along an elevation and

vegetation gradient in the Sacramento Mountains, New Mexico. M.S. thesis, University

of Arizona.

Wimmer, R. 2001. Arthur Freiherr von Sechendorff-Gudent and the early history of tree-ring

crossdating. Dendrochronologia 19(1): 153–158.

Woodhouse, C.A. 1999. Artificial neural networks and dendroclimatic reconstructions: An

example from the Front Range, Colorado, USA. The Holocene 9(5): 521-529.

475
Woodhouse, C.A. 2001. A tree-ring reconstruction of streamflow or the Colorado Front Range.

Journal of the American Water Resources Association 37(3): 561–569.

Woodhouse C. A., Gray S. T., and Meko D. M. 2006. Updated streamflow reconstructions for

the Upper Colorado River basin. Water Resources Research 42: W05415,

doi:10.1029/2005WR004455.

Worrall, J. 1990. Subalpine larch: oldest trees in Canada? The Forestry Chronicle: 478–479.

Worbes, M. 1989. Growth rings, increment and age of trees in inundation forests, savannas and a

mountain forest in the Neotropics. IAWA Bulletin 10(2): 109-122.

Worbes, M. 1995. How to measure growth dynamics in tropical trees: A review. IAWA Journal

16(4): 227–351.

Worbes, M. 2002. One hundred years of tree-ring research in the tropics -- a brief history and an

outlook to future challenges. In: P. Cherubini, ed., Tree rings and people. Conference

Proceedings, Davos, Switzerland, September 2001. Dendrochronologia 20 (1-2 (special

issue)): 217-231.

Worbes, M. and Junk, W.J. 1989. Dating tropical trees by means of 14C from bomb tests.

Ecology 70(2): 503–507.

Worbes, M., Staschel, R., Roloff, A., and Junk, W.J. 2003. Tree ring analysis reveals age

structure, dynamics and wood production of a natural forest stand in Cameroon. Forest

Ecology and Management 173(1-3): 105-123.

Wurster, C.M., Patterson, W.P., and Cheatham, M.M. 1999. Advances in micromilling

techniques: A new apparatus for acquiring high-resolution oxygen and carbon stable

isotope values and major/minor elemental ratios from accretionary carbonate. Computers

and Geosciences 25: 1159–1166.

476
Yadav, R.R. and Kulieshius, P. 1992. Dating of earthquakes: Tree ring responses to the

catastrophic earthquake of 1887 in Alma-Ata Kazakhstan. The Geographical Journal

158(3): 259–299.

Yamaguchi, D.K. 1983. New tree-ring dates for recent eruptions of Mount St. Helens.

Quaternary Research 20: 246–250.

Yamaguchi, D.K. 1991. A simple method for cross-dating increment cores from living trees.

Canadian Journal of Forest Research 21: 414–416.

Yamaguchi, D. K. and Hoblitt, R.P. 1995. Tree-ring dating of pre-1980 volcanic flowage

deposits at Mount St. Helens, Washington. GSA Bulletin 107(9): 1077–1093.

Yang, W., Spencer, R.J., and Krouse, H.R. 1996. Stable sulfur isotope hydrogeochemical

studies using desert shrubs and tree rings, Death Valley, California, USA. Gechemica et

Cosmochimica Acta 60: 3015–3022.

Yanosky, T.M. Hansen, B.P., and Schening, M.R. 2001. Use of tree rings to investigate the

onset of contamination of a shallow aquifer by chlorinated hydrocarbons. Journal of

Contaminant Hydrology 50: 159-173.

Yanosky, T.M., Hupp, C.R., and Hackney, C.T. 1995. Chloride concentrations in growth rings

of Taxodium distichum in a saltwater-intruded estuary. Ecological Applications 5(3):

785–792.

Yanosky, T.M. and Jarrett, R.D. 2001. Dendrochronologic evidence for the frequency and

magnitude of paleofloods. In: Ancient Floods, Modern Hazards: Principles and

Applications of Paleoflood Hydrology. Water Science and Application 5: 77–89.

477
Yapp, C.J. and Epstein, S. 1982. A re-examination of cellulose carbonbound hydrogen dD

measurements and some factors affecting plant–water D/H relationships. Geochimica et

Cosmochimica Acta 46: 955–965.

Yarnell, S.L. 1998. The southern Appalachians: A history of the landscape. USDA Forest

Service, Southern Research Station GTR-SRS-18. 45 pp.

Zeuner, F.E. 1958. Dating the Past: An introduction to geochronology. Fourth Edition.

Methuen and Co. Ltd. London. 491p.

Zetterberg, P. 1990. Dendrochronological dating of a wooden causeway in Finland. Norwegian

Archaeological Review 23(1-2): 54-59.

Zhang, Q.B., and Alfaro, R.I. 2002. Periodicity of two-year cycle spruce budworm outbreaks in

central British Columbia: A dendro-ecological analysis. Forest Science 48(4): 722–731.

Zimmermann, B., Schleser, G.H., and Brauning, A. 1997. Preliminary results of a Tibetan stable

C-isotope chronology dating from 1200 to 1994. Isotopes in Environmental and Health

Studies 33: 157–165.

478
Appendix A: Tree and Shrub Species that have been used by
dendrochronologists

How to use this list


The table has four columns: CDI (Crossdating Index), Species Code, Genus, Species, and
Common name(s).

 The CDI is the Crossdating Index, where


o "0" indicates the species does not crossdate, or no crossdating information is available
o "1" indicates a species known to crossdate within and between trees (minor importance to
dendrochronology)
o "2" indicates a species known to crossdate across a region
(major importance to dendrochronology).
 The "Code" is the standard four-letter abbreviation assigned by the International Tree-Ring Data Bank for
archiving purposes.
 An asterisk (*) beside the four-letter code indicates this species has tree-ring measurements and
chronologies held in the International Tree-Ring Data Bank.
 The common names are compiled from various sources.

From Ultimate Tree Ring Webpages http://web.utk.edu/~grissino/species.htm

CDI Species
Code ID Genus Species Common Name
2 ABAL* Abies alba silver fir, European fir
1 ABAM* Abies amabilis Pacific silver fir
2 ABBA Abies balsamea balsam fir
1 ABBO* Abies borisii-regis Bulgarian fir, King Boris fir
0 ABBN Abies bornmuelleriana Bornmueller's fir
0 ABBR Abies bracteata bristlecone fir
1 ABCE* Abies cephalonica Greek fir
0 ABCH Abies chensiensis Chensien fir
1 ABCI Abies cilicica Cilician fir
2 ABCO* Abies concolor white fir
0 ABEQ Abies equi-trojani
0 ABFX Abies faxoniana Faxon fir
0 ABFI Abies firma Japanese fir, Momi fir
1 ABFO Abies forestii Chinese fir
1 ABFR Abies fraseri Fraser fir
1 ABGR Abies grandis grand fir, giant fir
0 ABHO Abies holophylla Manchurian fir
1 ABKA Abies kawakamii Taiwan fir

479
1 ABKO Abies koreana Korean fir
2 ABLA* Abies lasiocarpa subalpine fir, corkbark fir
1 ABMA* Abies magnifica California red fir
1 ABMR Abies mariessi Marie's fir
1 ABMC Abies marocana Moroccan fir
0 ABNB Abies nebrodensis Sicilian fir
0 ABNE Abies nephrolepis East Siberian fir
1 ABNO* Abies nordmanniana Caucasian fir
1 ABNU Abies numidica Algerian fir
1 ABPI* Abies pindrow Himalayan silver fir
1 ABPN* Abies pinsapo Spanish fir
1 ABPR* Abies procera noble fir
1 ABRC Abies recurvata Min fir
0 ABRE Abies religiosa Mexican fir, sacred fir
0 ABSA Abies sachalinensis Sachalin fir, todo
0 ABSI Abies sibirica Siberian fir
1 ABSB* Abies spectabilis silver fir, East Himalayan fir
0 ABSQ Abies squamata flaky fir
0 ABVI Abies vietchii Vietch's silver fir
0 ACAL Acacia alpina
0 ACCA Acacia catechu cutch, Indian acacia
0 ACGI Acacia giraffae camel thorn
0 ACHO Acacia hotwittii
0 ACME Acacia melanoxylon blackwood
0 ACNI Acacia nilotica gum arabic tree
0 ACRA Acacia raddiana Israelian acacia
1 ACCA Acer campestre hedge maple, field maple
0 ACNE Acer negundo boxelder, ash-leaved maple
0 ACMO Acer mono maple
1 ACOP Acer opalus Italian maple
0 ACPE Acer pensylvanicum striped maple
1 ACPL Acer platanoides Norway maple
1 ACPS Acer pseudoplatanus sycamore maple, plane tree
1 ACRU Acer rubrum red maple
0 ACSA* Acer saccharinum silver maple
2 ACSH* Acer saccharum sugar maple
0 ACSC Acer spicatum mountain maple
0 ACTU Acer turkestanica
0 ADDI Adansonia digitata baobab, monkey bread tree
0 ADFA Adenostoma fasciculatum chamise, greasewood
1 ADHO* Adesmia horrida
1 ADUS* Adesmia uspallatensis
0 AEHI Aesculus hippocastanum horse chestnut
0 AEPU Aextoxicon punctatum olivillo, tique
0 AFAF Afzelia africana afzelia, apa, doussie, alinga, papao
0 AFQU Afzelia quanzensis afzelia, mambokofi, chanfuta
2 AGAU* Agathis australis kauri pine
0 AGMA Agathis macrophylla Fijian kauri
0 AGMO Agathis moorei kauri
0 AGOV Agathis ovata kauri

480
0 AGPA Agathis palmerstoni North Queensland kauri
2 AGRO Agathis robusta kauri pine, Queensland kauri
0 AGVI Agathis vitiensis
0 AIAL Ailanthus altissima Tree of Heaven
0 ALVE Allocasuarina verticillata
1 ALGL Alnus glutinosa common alder, European alder
0 ALHI Alnus hirsuta
1 ALIN Alnus incana grey alder, white alder
0 ALMA Alnus maximowiczii
0 ALRH Alnus rhombifolia white alder
0 ALRU Alnus rubra red alder
0 ALRG Alnus rugosa speckled alder, rough alder
0 ALSE Alnus serrulata hazel alder
0 ALSI Alnus sinuata Sitka alder
1 ALVI Alnus viridis green alder
0 ALCR Alnus viridis American green alder
1 AMSP Amelanchier Medik. serviceberry
0 AMOV Amelanchier ovalis
0 AMLU Amomyrtus luma luma
0 ANCO Andira coriacea Saint Martin rouge
1 ANSP Annona spraguei araucaria
1 ARAN Araucaria angustifolia Parana araucaria, Parana pine
2 ARAR* Araucaria araucana monkey puzzle, araucaria, pehuen
0 ARBI Araucaria bidwilli bunya pine, bunya
0 ARCU Araucaria cunninghamii hoop pine, Moreton bay pine
0 ARHE Araucaria heterophylla Norfolk Island pine
0 ARHU Araucaria hunsteinii pine
0 ARGL Arctostaphylos glauca bigberry manzanita
1 ARTR Artemisia tridentata big sagebrush
2 ATCU* Athrotaxis cupressoides pencil pine, smooth Tasmanian cedar
2 ATSE* Athrotaxis selaginoides King Billy pine
0 AUKL Aucoumea klaineana okoume
2 AUCH* Austrocedrus chilensis Chilean cedar, cipres de la cordillera,
0 BAAE Balanites aegyptiaca Jericho balsam, heglig
0 BLTA Beilschmiedia tawa Kirk tawa
0 BBVU Berberis vulgaria common barberry
0 BTEX Bertholletia excelsa Brazil nut, yuvia, turury, para nut tree
0 BEAB Betula albosinensis Chinese birch
1 BEAL Betula alleghaniensis yellow birch
0 BEER Betula ermanii Japanese birch, dakekaba
0 BEGL Betula glandulosa bog birch, dwarf birch
1 BEGR Betula grossa Japanese cherry birch
0 BELE Betula lenta sweet birch, black birch
0 BENI Betula nigra river birch
1 BEPA Betula papyrifera paper birch
0 BEAK Betula papyrifera
1 BEPE Betula pendula silver birch, European white birch
0 BEPL Betula platyphylla jagjag-namu, Japanese birch
0 BEPO Betula populifolia gray birch
1 BEPU Betula pubescens downy birch, mountain birch

481
1 BEUT Betula utilis Himalayan birch
1 BEVE Betula verrucosa silver birch, European white birch
0 BOQU Bombacopsis quinata
0 BOMA Bombax malabaricum semul, ngiu, ngiew, gon run do
1 BUGR Bursera graveolens palo santo
0 BUSI Bursera simaruba gumbo-limbo, West-Indian birch
1 BUSE Buxus sempervirens common box, boxwood
0 CACO Callitris columellaris cypress pine
0 CAIN Callitris intratropica cypress pine
0 CAMA Callitris macleayana brush cypress pine
1 CAPR* Callitris preissii Rottnest Island pine
1 CARO* Callitris robusta
1 CADE Calocedrus decurrens California incense cedar
1 CABU* Canthium burttii canthium
0 CASC Capparis scabrida sapote
0 CAPC Carapa procera carapa
0 CPBE Carpinus betulus hornbeam
0 CYCO Carya cordoformis bitternut hickory
1 CYGL Carya glabra pignut hickory
1 CYIL Carya illinoensis pecan
0 CYOV Carya ovata shagbark hickory
0 CYTO Carya tomentosa mockernut hickory
0 CAGL Caryocar glabrum chawari
0 CACR Castanea crenata Japanese chestnut
1 CADN Castanea dentata American chestnut
1 CASA Castanea sativa sweet chestnut, European chestnut
0 CSLI Casuarina litoralis black she-oak
0 CTSP Catalpa speciosa northern catalpa
0 CNCR Ceanothus crassifolius hoaryleaf ceanothus
1 CEAN* Cedrela angustifolia cedro salteno
0 CEFI Cedrela fissilis central American cedar
1 CELI* Cedrela lilloi cedro salteno
0 CEOD Cedrela odorata
0 CETO Cedrela toona Harms red cedar, Australian cedar,
2 CDAT Cedrus atlantica Atlantic cedar, Atlas cedar
1 CDBR* Cedrus brevifolia
1 CDDE Cedrus deodara deodar cedar, Himalayan cedar
1 CDLI* Cedrus libani Cedar of Lebanon
1 CLAU Celtis australis southern nettle tree, hackberry
0 CLCA Celtis caucasica Caucasian nettle tree
0 CLLA Celtis laevigata sugarberry
1 CLOC Celtis occidentalis hackberry
1 CLRE Celtis reticulata netleaf hackberry
0 CEOC Cephalanthus occidentalis buttonbush
0 CEMI Cercidium microphyllum yellow paloverde
0 CRBE Cercocarpus betuloides birchleaf mountain-mahogany
0 CRLE Cercocarpus ledifolius curlleaf mountain-mahogany
1 CRMO Cercocarpus montanus alderleaf cercocarpus
1 CHFO Chamaecyparis formosensis Formosan false cypress
2 CHNO Chamaecyparis nootkatensis Alaska yellow-cedar, Nootka cypress

482
1 CHOB Chamaecyparis obtusa hinoki cypress, Formosan cypress
1 CHPI Chamaecyparis pisifera sawara cypress
0 CHTH Chamaecyparis thyoides Atlantic white-cedar
0 CLEX Chlorophora excelsa iroko, kambala, mvule
0 CHSP Chorisia speciosa paneira
0 CIFR Citharexylum fruticosum Florida fiddlewood
0 COCO Copaifera coleosperma Rhodesian copalwood, mehibi
1 COAL Cordia alliodora laurel corriente, lauro amarillo, ajo ajo
0 COAP Cordia apurensis
0 COBI Cordia bicolor
0 COEL Cordia elaeagnoides
0 COTR Cordia trichotoma lauro pardo, peterebi
0 COFL Cornus florida flowering dogwood
0 COSA Cornus sanguinea
0 COAV Corylus avellana common hazel
0 COSI Corylus sieboldiana blume hazel
0 CTCO Cotinus coggygria European smoketree
0 CTSP Cotoneaster Medik. cotoneaster
0 CRAZ Crataegus azarolus azarole
0 CRMO Crataegus monogyna
2 CMJA* Cryptomeria japonica Japanese cedar, sugi, cryptomeria
1 CUAZ Cupressus arizonica Arizona cypress
0 CUAT Cupressus atlantica Atlas cypress
0 CUDU Cupressus dupreziana
1 CUGI Cupressus gigantea
0 CUGL Cupressus glabra smooth Arizona cypress
0 CULU Cupressus lusitanica Mexican cypress
2 CUSE Cupressus sempervirens Italian cypress, Mediterranean cypress
0 CYRA Cyrilla racemiflora swamp cyrilla, leatherwood
0 DADA Dacrycarpus dacrydioides kahikatea, white pine
1 DABD Dacrydium bidwillii New Zealand mountain pine
1 DABI* Dacrydium biforme
1 DACO* Dacrydium colensoi
2 DACU Dacrydium cupressinum rimu, red pine
1 DAFR Dacrydium franklinii Huon pine
0 DIGU Dicorynia guianensis angelique
0 DSVI Diospyros virginiana common persimmon
0 DITO Discaria toumatou matagouri, tumatu-kuru, wild Irishman
1 DITR Discaria trinervis
0 DRLA Dracophyllum latifolium neinei
0 DRWI Drimys winteri canelo, winter bark
1 DUVI Duschenkia viridis .
0 DYMA Dysoxylum malabaricum Bombay white cedar
0 ELGL Elaeoluma glabrascens rΘv.
1 EMRU Empetrum rubrum murtilla
1 ENCA Enkianthus campanulatus
0 ENAN Entandrophragma angolense gedu nohor, kalungi, tiama, edinam
0 ENCA Entandrophragma candollei kosipo, omu
0 ENCY Entandrophragma cylindricum sapeli, sapele, sapelli, assi
0 ENUT Entandrophragma utile sipo, utile EPSP Ephedra L. ephedra

483
0 EUCA Eucalyptus camaldulensis river red gum
0 EUDE Eucalyptus delegatensis alpine ash
0 EUGL Eucalyptus globulus Tasmanian bluegum
0 EUMA Eucalyptus marginata jarrah
0 EUMI Eucalyptus miniata Darwin woolybutt
0 EUNE Eucalyptus nesophila Melville Island bloodwood
0 EUOR Eucalyptus oreades Blue Mountains ash
0 EUPA Eucalyptus pauciflora snow gum, cabbage gum
0 EUST Eucalyptus stellulata black salee
0 EUTE Eucalyptus tetradonta Darwin stringybark
0 EUVI Eucalyptus viminalis ribbongum
0 EUCO Eucryphia cordifolia ulmo, muermo
0 EUJA Eugenia jambolana jaman, kelat eugenia
0 EXCU Exocarpus cuppressiforme native cherry
1 FAGR* Fagus grandifolia American beech
1 FAOR Fagus orientalis Oriental beech, eastern beech
2 FASY* Fagus sylvatica European beech, common beech
2 FICU* Fitzroya cupressoides alerce, Patagonian cypress
1 FRAM Fraxinus americana white ash
0 FRCA Fraxinus caroliniana Carolina ash
0 FACR Fagus crenata bunya beech
1 FREX* Fraxinus excelsior European ash, common ash
0 FRMA Fraxinus mandshurica Manchurian ash, yachidamo
1 FRNI* Fraxinus nigra black ash
0 FRPE Fraxinus pennsylvanica green ash, red ash
1 FRSP Fraxinus spaethiana ash
1 FRVE Fraxinus velutina velvet ash
0 GEAV Gevuina avellana avellano
0 GIBI Gingko biloba maidenhair tree, gingko
0 GLTR Gleditsia triacanthos honey locust
0 GMAR Gmelina arborea gumari, gumbar, yemane, gmelina
0 GOLA Gordonia lasianthus loblolly-bay
0 GOGL Goupia glabra goupia
0 GRVI Grevillea victoriae
0 GUCE Guarea cedrata bosse, guarea, white guarea
0 HABD Halocarpus bidwillii bog pine
1 HABI* Halocarpus biformis pink pine
0 HAKI Halocarpus kirkii manoao
0 HAVI Hamamelis virginiana witch hazel
0 HEAN Hedycaria angustifolia native mulberry
0 HEAR Heteromeles arbutifolia toyon
1 HEBR Hevea brasiliensis
0 ILAQ Ilex aquifolium English holly
0 ILCA Ilex cassine dahoon, dahoon holly
0 ILCO Ilex coriacea large gallberry, sweet gallberry
0 ILGL Ilex glabra inkberry, gallberry
0 ILIN Ilex inundata
0 ILOP Ilex opaca American holly
0 JACO Jacaranda copaia copaia, gobaja, futui, caroba
1 JGAU* Juglans australis Argentine walnut

484
0 JGCI Juglans cinerea butternut
0 JGNI Juglans nigra black walnut
0 JGRE Juglans regia common walnut
0 JUCH Juniperus chinensis Chinese juniper
1 JUCO Juniperus communis common juniper
0 JUDE Juniperus deppeana alligator juniper
1 JUDR Juniperus drupacea Syrian juniper
1 JUEX Juniperus excelsa Greek juniper, Grecian juniper
1 JUFO Juniperus foetidissima stinking juniper
0 JUMA Juniperus macropoda Himalayan pencil pine
0 JUMO Juniperus monosperma one-seed juniper
2 JUOC* Juniperus occidentalis western juniper
1 JUOS Juniperus osteosperma Utah juniper
1 JUOX Juniperus oxycedrus prickly juniper
1 JUPH Juniperus phoenicea Phoenicean juniper
0 JUPI Juniperus pinchotii redberry juniper, Pinchot juniper
0 JUPC Juniperus procera Uganda juniper, African pencil cedar,
1 JUPR Juniperus przewalskii Qilianshan juniper
1 JURE Juniperus recurva drooping juniper
2 JUSC* Juniperus scopulorum Rocky Mountain juniper
1 JUSM Juniperus semiglobosa
1 JUSE Juniperus seravschanica
0 JUTH Juniperus thurifera Spanish juniper
1 JUTU Juniperus turkestanica Turkestan juniper
2 JUVI* Juniperus virginiana eastern red-cedar
0 KHGR Khaya grandifolia acajou, Benin mahogany
0 KRDR Krenevaja drevesina
0 KUER Kunzea ericoides kanuka, white tea tree
0 LBGL Labatia glomerata
0 LBAN Laburnum anagyroides common laburnum
1 LGCO* Lagarostrobus colensoi
1 LGFR Lagarostrobus franklinii huon pine
0 LSFL Lagerstroemia flos-reginae pyinma, banaba, banglang, jarul
0 LSPA Lagerstroemia parviflora lendia
0 LSLA Lagerstroemia lanceolata benteak, nana
2 LADE* Larix decidua European larch
1 LAGM* Larix gmelinii Dahurian larch
1 LAGR Larix griffithiana Himalayan larch
1 LAJA Larix japonica Japanese larch
2 LALA* Larix laricina tamarack, eastern larch
2 LALY* Larix lyalli subalpine larch
2 LAOC* Larix occidentalis western larch
1 LAPO Larix potanini Chinese larch
2 LASI* Larix sibirica Siberian larch
0 LAPH Laurelia philippiana tepa
0 LASE Laurelia sempervirens laurelia, Chilean laurel, huahuan
0 LAHU Laxopterigium huasango haltaco
0 LECO Lecythis corrugata angelique
0 LEIN Lepidothamnus intermedius yellow-silver pine
0 LEFL Leptospermum flavescens tea tree

485
0 LESC Leptospermum scoparium manuka, red tea tree, black manuka
2 LIBI* Libocedrus bidwillii New Zealand cedar, pahautea
0 LIPL Libocedrus plumosa kawaka, plume incense cedar
0 LGVU Ligustrum vulgare
1 LIST Liquidambar styraciflua sweetgum
1 LITU* Liriodendron tulipifera tuliptree, yellow-poplar, tulip-poplar
0 LOFR Lomatia fraseri silky lomatia, tree lomatia
0 LOHI Lomatia hitsuta radal
0 LOXY Lonicera xylosteum
0 LOTR Lovoa trichilioides dibetou
0 MAAC Magnolia accuminata cucumbertree
0 MAGR Magnolia grandiflora southern magnolia
0 MAVI Magnolia virginiana sweetbay, swampbay
0 MASY Malus sylvestris apple tree
0 MABI Manilkara bidentata balata franc
0 MICH Michelia champaca champak
0 MINI Michelia niligirica pilachampa, champak
0 MOCO Moronobea coccinea manil montagne, mountain manil
0 MOAL Morus alba white mulberry
0 MORU Morus rubra red mulberry
0 MYCE Myrica cerifera southern bayberry, bayberry
0 MYGA Myrica gale sweet gale, bog myrtle
0 NEAM Nectandra amazonum
0 NTLO Notelaea longifolia large mock-olive
0 NOAL Nothofagus alpina rauli
1 NOAN Nothofagus antarctica Antarctic beech, nirre
1 NOBE* Nothofagus betuloides coihue de Magallanes, guindo
0 NOCU Nothofagus cunninghamii Australian nothofagus, myrtle beech
0 NODO Nothofagus dombeyi coihue, Dombey's southern beech
0 NOFU Nothofagus fusca red beech, New Zealand red beech
1 NOGU* Nothofagus gunnii tanglefoot beech
2 NOME* Nothofagus menziesii silver beech, Menzies's red beech
0 NONE Nothofagus nervosa rauli
0 NONI Nothofagus nitida roble chicote
1 NOOB Nothofagus obliqua southern beech, roble
1 NOPU* Nothofagus pumilio lenga
2 NOSO* Nothofagus solandri mountain beech, black beech
0 NYOG Nyssa ogechee Ogeechee tupelo
0 NYSY Nyssa sylvatica black tupelo, blackgum
0 OCUS Ocotea usambarensis ocotea, camphor
0 OSCA Ostrya carpinifolia hop hornbeam
0 OXAR Oxydendrum arboreum sourwood
0 PARI Parapiptadenia rigida
0 PAAU Parkia auriculata
0 PATO Paulownia tomentosa empress tree
0 PECA Peronema canescens sunkai, koeroes
0 PEBO Persea borbonia redbay, shorebay
0 PELI Persea lingue lingue
0 PELN Petrophile linearis pixie mops
0 PBPO Phoebe porfiria

486
1 PHAL* Phyllocladus alpinus mountain toatoa, alpine celery top pine
1 PHAS* Phyllocladus aspleniifolius
1 PHGL* Phyllocladus glaucus toatoa
1 PHTR* Phyllocladus trichomanoides tanekaha, celery pine
2 PCAB* Picea abies Norway spruce
0 PCAS Picea asperata dragon spruce
1 PCBA Picea balfouriana
1 PCBR Picea brachytyla
1 PCCA Picea cajanensis
1 PCCH Picea chihuahuana chihuahua spruce
2 PCEN* Picea engelmannii Engelmann spruce
2 PCGL* Picea glauca white spruce
1 PCGN Picea glehnii Sakhalin spruce
0 PCJE Picea jezoensis Yezo spruce, Hondo spruce
1 PCLI Picea likiangensis Likiang spruce
2 PCMA* Picea mariana
1 PCOM* Picea omorika Serbian spruce, Pancic spruce
1 PCOR* Picea orientalis eastern spruce, Oriental spruce
2 PCPU* Picea pungens blue spruce, Colorado spruce
1 PCPR Picea purpurea
2 PCRU* Picea rubens red spruce
1 PCSH Picea shrenkiana Shrenk's spruce
2 PCSI* Picea sitchensis Sitka spruce
1 PCSM Picea smithiana Himalayan spruce
1 PCTI Picea tienschanica Tien-shan spruce
2 PLUV* Pilgerodendron uviferum cipres de las Guaytecas
2 PIAL* Pinus albicaulis whitebark pine
2 PIAR* Pinus aristata Rocky Mountain bristlecone pine
1 PIAM* Pinus armandii David's pine, Armand's pine
2 PIBA* Pinus balfouriana foxtail pine
2 PIBN* Pinus banksiana jack pine
1 PIBR* Pinus brutia Calabrian pine, brutia pine, see kiefer
0 PIBU Pinus bungeana lacebark pine
0 PICN Pinus canariensis Canary Island pine
0 PICA Pinus caribaea Caribbean pine, Cuban pine
2 PICE* Pinus cembra Swiss stone pine, Arolla pine
2 PICM* Pinus cembroides Mexican pinyon, Mexican nut pine
1 PICH Pinus chihuahuana chihuahua pine
2 PICO* Pinus contorta lodgepole pine
0 PICL Pinus coulteri Coulter pine, bigcone pine
1 PIDN Pinus densata
1 PIDE* Pinus densiflora Japanese red pine
2 PIEC* Pinus echinata shortleaf pine
2 PIED* Pinus edulis pinyon, Colorado pinyon
1 PIEL Pinus elliottii slash pine
1 PIEN Pinus engelmannii Apache pine
2 PIFL* Pinus flexilis
1 PIGE Pinus gerardiana chilgoza pine, Gerard's pine
2 PIHA* Pinus halepensis Aleppo pine, Jerusalem pine
1 PIHE Pinus heldreichii Heldreich's pine, panzer fohre

487
2 PIJE* Pinus jeffreyi Jeffrey pine
1 PIKE Pinus kesiya Khasi pine
1 PIKO Pinus koraiensis Korean pine
1 PILG Pinus lagunae laguna pinyon
2 PILA* Pinus lambertiana sugar pine
2 PILE* Pinus leucodermis Bosnian pine, greybark pine
2 PILO* Pinus longaeva Intermountain bristlecone pine
1 PIMA Pinus massoniana Masson pine
1 PIMK Pinus merkusii Merkus pine, mindoro pine
1 PIME Pinus mesogeensis cluster pine
2 PIMO* Pinus monophylla singleleaf pinyon
1 PIMZ Pinus montezumae Montezuma pine
1 PIMC Pinus monticola western white pine
1 PIMU* Pinus mughus krummholz pine
1 PIMG Pinus mugo mountain pine, stone pine
0 PIMR Pinus muricata bishop pine
2 PINI* Pinus nigra Austrian pine, black pine
0 PIOC* Pinus occidentalis West Indian pine
0 PIOO Pinus oocarpa Nicaraguan pitch pine, ocote pine
1 PIPA* Pinus palustris longleaf pine
0 PIPT Pinus patula Mexican weeping pine
1 PIPE* Pinus peuce Macedonian pine, Balkan pine
1 PIPI* Pinus pinaster maritime pine, cluster pine
2 PIPN* Pinus pinea Italian stone pine, umbrella pine
2 PIPO* Pinus ponderosa ponderosa pine, western yellow pine
1 PIPM Pinus pumila dwarf Siberian pine
1 PIPU* Pinus pungens Table Mountain pine
1 PIQU Pinus quadrifolia Parry pinyon
1 PIRA Pinus radiata Monterrey pine
2 PIRE* Pinus resinosa red pine
1 PIRI* Pinus rigida pitch pine
1 PIRO Pinus roxburghii chir pine
1 PISI Pinus sibirica Siberian stone pine
2 PISF* Pinus strobiformis southwestern white pine
2 PIST* Pinus strobus eastern white pine, Weymouth pine
2 PISY* Pinus sylvestris Scots pine, Scotch pine
1 PITB Pinus tabulaeformis Chinese pine
2 PITA* Pinus taeda loblolly pine
0 PITH Pinus thunbergii Japanese black pine
1 PITO Pinus torreyana Torrey pine
2 PIUN Pinus uncinata mountain pine
1 PIVI Pinus virginiana Virginia pine, scrub pine
1 PIWA Pinus wallichiana Himalayan pine, kail pine, blue pine
0 PSGR Pisonia grandis
0 PTAT Pistacia atlantica Atlas pistache, betoum
0 PTKH Pistacia khinjuk kakkar
0 PTPA Pistacia palaestina Israelian pistache
0 PTVE Pistacia vera green mastic, real mastictree
0 PLAC Platanus acerifolia London plane tree
1 PLOC Platanus occidentalis American sycamore

488
0 PLOR Platanus orientalis Oriental plane tree
0 PLIN Platonia insignis parcouri
1 PLOR Platyeladus orientalis Chinese pine
0 PYSA Polyscias sambucifolius elderberry panax, elderberry ash
0 POFA Podocarpus falcatus yellowwood, oteniqua
0 POHA Podocarpus hallii Hall's totara
0 POLA Podocarpus lawrencei Tasmanian podocarpus
1 PONE Podocarpus neriifolius thitmin
0 PONI Podocarpus nivalis snow totara
1 PONU Podocarpus nubigensus manio de hojas punzantes
0 POPA Podocarpus parlatorei
0 POTO Podocarpus totara totara
1 PPAL Populus alba white poplar
0 PPAN Populus angustifolia narrowleaf cottonwood
1 PPBA Populus balsamifera balsam poplar
0 PPDE Populus deltoides eastern cottonwood
1 PPEU Populus euphratica charab poplar, Indian poplar
0 PPFA Populus fastigiata
1 PPFR Populus fremontii Fremont cottonwood
1 PPGR Populus grandidentata bigtooth aspen
1 PPNI Populus nigra lombardy poplar, black poplar
1 PPSI Populus sieboldii Japanese aspen
1 PPTR Populus tremuloides quaking aspen
0 PPTC Populus trichocarpa black cottonwood
1 PRMX* Premna maxima muchichio
1 PRFL Prosopis flexuosa
0 PRGL Prosopis glandulosa honey mesquite
0 PMAN Prumnopitys andina lleuque
0 PMFE Prumnopitys ferruginea miro
0 PMTA Prumnopitys taxifolia matai, black pine
0 PNAM Prunus americana American plum
0 PNAV Prunus avium wild cherry
0 PNIL Prunus ilicifolia
0 PNMA Prunus mahaleb
0 PNPE Prunus pennsylvanica pin cherry
1 PNSE Prunus serotina black cherry
0 PNSP Prunus spinosa
0 PSMU Pseudobombax munguba muguba, huira
1 PSSE Pseudobombax septenatum
1 PSJA Pseudotsuga japonica Japanese Douglas-fir
1 PSMA* Pseudotsuga macrocarpa bigcone Douglas-fir
2 PSME* Pseudotsuga menziesii Douglas-fir
0 PSAX Pseudowintera axillaris
0 PSCO Pseudowintera colorata mountain horopito, pepper tree
0 PSXA Pseudoxandra polyphleba
0 PTAN* Pterocarpus angolensis Muninga, bloodwood
0 PTVE Pterocarpus vernalis
0 PTRH Pterocarya rhoifolia Japanese wing nut
0 PTPA Pteronia pallens
1 PUTR Purshia tridentata bitter brush

489
0 QUAC Quercus acutissima
0 QUAF Quercus afares
2 QUAL* Quercus alba white oak
0 QUBI Quercus bicolor swamp white oak
0 QUBO Quercus boissieri Israelian oak
1 QUBR Quercus brantii
0 QUCL Quercus calliprinos Kermes oak, Israelian oak
1 QUCA Quercus canariensis Mirbeck's oak, Algerian oak
1 QUCE Quercus cerris Turkey oak, Austrian oak
1 QUCO Quercus coccinea scarlet oak
0 QUCP Quercus copeyensis
0 QUCR Quercus costaricensis
1 QUDE Quercus dentata kashiwa oak, Daimio oak
1 QUDG* Quercus douglasii blue oak
1 QUDS Quercus dschoruchensis
1 QUEL Quercus ellipsoidalis northern pin oak
0 QUEM Quercus emoryi Emory oak
0 QUEN Quercus engelmannii Engelmann oak
1 QUFG Quercus faginea Portuguese oak
1 QUFA Quercus falcata southern red oak
1* QUFR Quercus frainetto Hungarian oak
2 QUGA Quercus gambelii Gambel oak
0 QUGY Quercus garryana Oregon white oak
1 QUGR Quercus grisea gray oak
1 QUHA Quercus hartwissiana
0 QUIL Quercus ilex holm oak, holly oak
0 QUIT Quercus ithaburensis Mt. Tabor oak
0 QUKE Quercus kelloggii California black oak
1 QULA Quercus laurifolia laurel oak
1 QULO Quercus lobata valley oak
0 QULU Quercus lusitanica oak
1 QULY* Quercus lyrata overcup oak
1 QUMA* Quercus macrocarpa bur oak
0 QUMC Quercus macrolepis Valonia oak
0 QUML Quercus marilandica blackjack oak
0 QUMI Quercus michauxii swamp chestnut oak
0 QUMO Quercus mongolica Mongolian oak
0 QUGS Quercus mongolica
0 QUMU Quercus muehlenbergii chinkapin oak
1 QUNI Quercus nigra water oak
0 QUPA Quercus palustris pin oak
2 QUPE* Quercus petraea durmast oak, sessile oak
1 QUPO Quercus pontica Armenian oak
1 QUPR* Quercus prinus chestnut oak
2 QUPU Quercus pubescens downy oak, pubescent oak
0 QUPY Quercus pyrenaica Pyrenean oak
2 QURO* Quercus robur English oak
1 QURU* Quercus rubra red oak
1 QUSH Quercus shumardii Shumard oak
2 QUST* Quercus stellata post oak

490
0 QUSU Quercus suber cork oak, cork tree
1 QUVE* Quercus velutina black oak
0 QUAC Quintinia acutifolia Westland quintinia
0 RAGU Rapanea guianensis guiana rapanea
0 RESP Recordoxylon speciosum wacapou guitin
0 RHCA Rhamnus caroliniana Carolina buckthorn
0 RHCT Rhamnus cathartica
0 RHCR Rhamnus crocea hollyleaf buckthorn
0 RHOV Rhus ovata sugar sumac
1 RONE Robinia neomexicana New Mexico locust
0 ROPS Robinia pseudoacacia black locust
0 SBPI Sabina pingu
0 SBRE Sabina recurva
1 SBSA Sabina saltuaria
1 SBTI Sabina tibetica
1 SBWA Sabina wallichiana
0 SAAC Salix acutifolia pointed-leaved willow
1 SAAL Salix alba white willow
0 SAAM Salix amygdalina almond-leaved willow
0 SAAD Salix amygdaloides peachleaf willow
0 SAAR Salix arbusculoides littletree willow
0 SAAT Salix arctica Arctic willow
0 SABA Salix babylonica weeping willow
0 SACN Salix candida sage-leaf willow, silver willow
0 SACA Salix caprea pussy willow, goat willow
0 SACR Salix caroliniana Coastal Plain willow
0 SADI Salix discolor pussy willow, glaucous willow
0 SAEL Salix elaeagnos hoary willow
0 SAEX Salix exigua sandbar willow
0 SAGL Salix glauca grayleaf willow
0 SALA Salix lanata Richardson's willow
0 SALS Salix lasiolepis arroyo willow, white willow
0 SAMY Salix myrsinifolia
0 SAPH Salix phylicifolia tea-leaf willow
0 SAPL Salix planifolia sandbar willow, lakeshore willow,
0 SAPU Salix purpurea purple willow, purple osier
0 SAVI Salix viminalis basket willow, common osier
0 SNAL Santalum album sandalwood, santalin, chandal
0 SSAL Sassafras albinum sassafras
0 SSAL Sapium styllare
1 SACO Saxegothaea conspicua Prince Albert's yew
0 SCTR Schleichera trijuga ta-kro, kusum, kusamo
0 SCMI Schleronema micranthum cordeiro, scleronema
1 SCVE Sciadopitys verticillata
1 SESE Sequoia sempervirens coast redwood
2 SEGI Sequoiadendron giganteum giant sequoia
0 SHRO Shorea robusta sal
0 SIAM Simarouba amara simarouba
0 SOAM Sorbus americana mountain ash
0 SOAR Sorbus aria whitebeam

491
0 SOAU Sorbus aucuparia mountain ash, rowan
1 SOTE Sorbus torminalis chequer tree, wild service tree
0 SODU Sorocea duckei
0 SWLA Swartzia laevicarpa saboarana
0 SWMC Swietenia macrophylla
0 SWMA Swietenia mahagoni West Indies mahogany
0 SYGL Symphonia globulifera manil
0 TABA Tabebuia barbata Igapo-tree
0 TMAP Tamarix aphylla dur
1 TMCH Tamarix chinensis tamarisk, salt cedar
0 TMJO Tamarix jordanis
0 TPGU Tapirira guianensis tapirira, cedroi, jobo
0 TMXE Tasmannia xerophila
0 TAAS Taxodium ascendens pond cypress
2 TADI* Taxodium distichum baldcypress
2 TAMU* Taxodium mucronatum Montezuma cypress
1 TABA Taxus baccata common yew, English yew
1 TACU Taxus cuspidata Japanese yew
1 TEGR Tectona grandis teak
0 TEBR Terminalia brownii
0 TEGU Terminalia guianensis
0 TETO Terminalia tomentosa Indian laurel, taukkyan, sain
1 TEAR Tetraclinis articulata Arar tree, African thuya
2 THOC* Thuja occidentalis northern white-cedar
0 THOR Thuja orientalis Chinese arborvitae, Oriental arborvitae
1 THPL* Thuja plicata western redcedar, giant arborvitae
1 THST Thuja standishii Japanese arborvitae
1 THDO Thujopsis dolabrata hiba arborvitae
1 THHO Thujopsis dolabrata asunaro arborvitae
1 TIAM Tilia americana American basswood
1 TICO Tilia cordata littleleaf linden, winter linden,
1 TIPL Tilia platyphyllos broad-leaved linden, summer linden
1 TOCA Torreya californica California nutmeg
0 TRSC Triplochiton schleroxylon abachi, obeche, wawa, arere
0 TRCO Tristania conferta Queensland box tree
2 TSCA* Tsuga canadensis eastern hemlock
1 TSCR* Tsuga caroliniana Carolina hemlock
0 TSCH Tsuga chinensis Chinese hemlock
0 TSDI Tsuga diversifolia Japanese hemlock
1 TSDU Tsuga dumosa East Himalayan hemlock
2 TSHE* Tsuga heterophylla western hemlock
2 TSME* Tsuga mertensiana mountain hemlock
0 TSSI Tsuga sieboldii southern Japanese hemlock
1 ULGL Ulmus glabra Wych elm, Scots elm, mountain elm
1 ULLA Ulmus laevis European white elm
1 ULMI Ulmus minor smooth-leaved elm, field elm
0 ULPU Ulmus pumila Siberian elm
1 ULRU Ulmus rubra slippery elm
0 VBLA Vibernum lantana
0 VIME Virola melinonii mountain yayamadou

492
1 VIKE* Vitex keniensis moru, moru oak
0 VOAM Vouacapoua americana wacapou
0 WERA Weinmannia racemosa kamahi
0 WETR Weinmannia trichosperma tineo, tenio, palo santo
1 WICE* Widdringtonia cedarbergensis Clanwilliam cedar
0 ZISP Ziziphus spina-christi Judas tree, Christ thorn
0 ZYDU Zygophyllum dumosum

493
Appendix B: Age of the oldest trees per species.

This list is a compilation of Peter Brown’s OLDLIST and Neil Pederson’s Eastern OLDLIST for the eastern United
States. Those two lists have been combined here, organized by the oldest age of the trees, and filtered so that only
the oldest individual is represented for each species.

The table includes genus, species, age, type (see below), sample identification number, location of the sample, and
the collector’s information or a reference where the tree is mentioned.

Four types of ages are recognized in the database:


XD: crossdated
RC: ring counted
EX: extrapolations (based on ring measurements usually)
HI: historic record

Brown, P. 1996. Oldlist: A Database of Maximum. In: Dean, J.S., Meko, D.M., and Swetnam, T.W. (eds.)
Proceedings of the International Conference on Tree Ages. Tree Rings, Environment, and Humanity:
Relationships and Processes, 17-21 May, 1994, Tucson, Arizona. Radiocarbon 1996: 727-731.

Collector(s),
Dater(s),
Genus Species Age Type ID Location Reference

Wheeler Peak,
Pinus longaeva 4844 RC WPN-114 Nevada, USA Currey 1965

Lara and Villalba


Fitzroya cupressoides 3622 XD Chile 1993

M. Hughes, R.
Sierra Nevada, Touchan, E.
Sequoiadendron giganteum 3266 XD CBR26 California, USA Wright

Scofield Sierra Nevada, Miles and


Juniperus occidentalis 2675 XD juniper California, USA Worthington 1998

494
Central Colorado, Brunstein and
Pinus aristata 2435 XD CB-90-11 USA Yamaguchi 1992

Ficus religiosa 2217 HI Sri Lanka Anonymous

Northern California,
Sequoia sempervirens 2200 RC USA E. Fritz

Sierra Nevada,
Pinus balfouriana 2110 XD SHP 7 California, USA A. Caprio

Kananaskis, Alberta,
Larix lyalli 1917 EX Canada Worrall 1990

Northern New H. Grissino-Mayer,


Juniperus scopulorum 1889 XD CRE 175 Mexico, USA R. Warren

Northern New T. Swetnam, T.


Pinus flexilis 1670 XD ERE Mexico, USA Harlan

Sierra Nevada,
Pinus balfouriana 1666 XD RCR 1 California, USA A. Caprio

Kelly and Larson


Thuja occidentalis 1653 XD FL117 Ontario, Canada 1997

Vancouver Island,
Chamaecyparis nootkatensis 1636 RC? Canada L. Jozsa

495
Bladen County, Stahle, Cleaveland,
Taxodium distichum 1622 XD BLK 69 North Carolina, USA Hehr 1988

Northern New
Psuedotsuga menziesii 1275 XD BIC 63 Mexico, USA H. Grissino-Mayer

Perkins and
Pinus albicaulis 1267 XD RRR15 Central Idaho, USA Swetnam 1996

Lagarostrobus franklinii 1089 XD Tasmania, Australia Cook et al. 1991

Pinus edulis 973 XD SUN 2522 Northeast Utah, USA Schulman 1956

Wah Wah Mtns,


Pinus ponderosa 929 XD Utah, USA S. Kitchen

Central Colorado,
Picea engelmannii 911 XD FCC 23 USA Brown et al. 1995

Pine Grove Hills, F. Biondi, S.


Pinus monophylla 888 XD PGH-02 Nevada, USA Strachan

Juniperus virginiana 860 XD -- MO R. Guyette

Nachin, B.
Buckley, N.
Larix siberica 750 XD OVL-5N Ovoont, Mongolia Pederson

496
Dan Sperduto
Nyssa sylvatica 679 XD -- NH P. Krusic
Luckman 2003 (B.
Luckman, R. van
Dorp, D.
Klauane Lake, Youngblut, M.
Picea glauca 668 XD Yukon, Canada Masiokas)

Tarvagatay Pass, G. Jacoby, Nachin,


Pinus siberica 629 XD TPX-16 Mongolia D. Frank

Truckee, California, F. Biondi, S.


Pinus jeffreyi 626 XD TGS-02 USA Strachan

San Mateo Mtns, H. Grissino-Mayer,


Pinus strobiformis 599 XD VPK02 New Mexico, USA J. Speer, K. Morino

E. Cook; Cook and


Tsuga canadensis 555 XD 39021 Tionesta, PA Cole, 1991

Abruzzi Nat'l Park, Piovesan et al.


Fagus sylvatica 503 XD 1012306F Italy 2005

Luckman 2003 (B.


Southern Yukon, Luckman, M.
Abies lasiocarpa 501 XD Canada Kenigsberg)

Granite Lake,
Kenora, Ontario S. St. George;
Pinus resinosa 500 RC -- Canada Ontario's Old Trees

Bavarian Forest,
Picea abies 468 XD LBG Germany R. Wilson

497
E. Cook;
Quercus alba 464 XD 85141 Buena Vista, VA N. Pederson

Sierra Nevada,
Torreya californica 455 XD California, USA A. Caprio

Fundy Escarpment,
New Brunswick,
Picea rubens 445 RC 05BCL901a Canada B. Phillips

Fundy Escarpment,
Picea rubens 445 XD 05BCL901a NB, Cana. B. Phillips

Great Smoky
Mountains National
Liriodendron tulipifera 434 RC -- Park W. Blozan

D.Stahle,
M.Therrell,
Guadalupe Mtns. D.Griffin,
Quercus muehlenbergii 429 XD PSC23 Nat. Park, TX D.(Daniel)Stahle

E. Cook; N.
Pederson; Pederson
Quercus montana 427 XD LBC25 Uttertown, NJ et al., 2004

R. Guyette, M.
Platanus occidentalis 412 RC BHY001 MO Stambaugh

Swan Lake,
Algonquin Park, Ont. R.P. Guyette & B.
Pinus strobus 408 XD sww51 Canada Cole; ITRDB

North central
Quercus gambelli 401 XD Arizona, USA F. Biondi

498
Quercus stellata 395 XD KEY13 Osage County, OK D. Stahle; ITRDB

Vasiliauskas, S.
Algonquin Park, A.,; Ontario's Old
Betula alleghaniensis 387 RC -- Ontario, Canada Trees

Pinus rigida 375 XD -- Mohonk Lake, NY E. Cook; ITRDB

E. Cook; N.
Pederson & H.M.
Hopton; Pederson
Betula lenta 361 XD STE03 New Paltz, NY et al. in press

N. Pederson; A.
Curtis; Pederson et
Carya ovata 354 XD WFS08a Fiddler's Green, VA al. in press
Sprewell Bluff
Wildlife
Management Area,
Meriwether County,
GA (on the
Pinus palustris 354 XD SPB35 Piedmont) T. Knight

N. Pederson; H.M.
Hopton; Pederson
Magnolia acuminata 348 XD MDC02b Fiddler's Green, VA et al. in press

R. Guyette, M.
Quercus macrocarpa 343 XD BHY002 MO Stambaugh

Sleeping Giant Prov.


Park, Ontario, Girardin et. al,
Picea mariana 330 XD -- Canada 2006

Wachusett Mountain,
Quercus rubra 326 XD hem79 Massachusetts, USA Orwig et al. 2001

499
Lac Duparquet, Tardif & Bergeron,
Fraxinus nigra 319 XD -- Quebec 1999

Saline County,
Pinus echinata 315 XD LAW38 Arkansas, USA D. Stahle

E. Cook; Cook and


Tsuga caroliniana 307 XD 101231 Kelsey Tract, NC Cole, 1991
P. Sheppard & C.
Canham; P.
Sheppard;
Pederson et al. in
Acer rubrum 300 XD CATB142 Catskill Mtns, NY press

Catkill, New York, D & N. Pederson,


Quercus bicolor 285 XD RHS01a USA M. Hopton

Peter's Woods, Martion & Martin,;


Acer saccharum 280 RC -- Ontario, Canada Ontario's Old Trees

Greenbriar, Great
Smoky Mountains, J. Young, W.
Castanea dentata 270 XD GB204B TN Blozan; ITRDB

N. Pederson;
Mohonk Preserve, Pederson et al.,
Carya glabra 265 XD BCV16a NY 2004

Pisgah Nat'l Forest,


Quercus prinus 265 XD GL18 North Carolina, USA Speer 2001

R. Guyette, M.
Acer nigrum 247 RC BHY038 MO Stambaugh

500
Blue Lake, Ontario, Girardin et. al,
Pinus banksiana 246 XD -- Canada 2006

C. Krause, H.
Abies balsamea 245 XD -- Lac Liberal, Canada Morin; ITRDB

N. Pederson; T.
Congaree Swamp Doyle; Pederson et
Pinus taeda 241 XD -- National Park, SC al., 1997

Rainbow falls Prov.


Park, Ontario, Girardin et. al,
Betula papyrifera 240 XD -- Canada 2006

Rambulette Creek,
Quecus margaretta 234 XD RCP41 Taylor County, GA T. Knight

G. DeWeese, H.
Griffith Knob, Grissino-Mayer, C.
Pinus pungens 232 XD GKA111 Virginia, USA Lafon

Vasiliauskas, S.
Algonquin Park, A.,; Ontario's Old
Ostrya virginiana 230 RC -- Ontario, Canada Trees

Cotinus obovatus 221 RC -- Arkansas R. Guyette

Alley Spring,
Shannon County,
Quercus velutina 219 XD 293503 Missouri S. Voelker

Desha County
Quercus lyrata 218 XD SNA7 Arkansas, USA D. Stahle

501
Lake Abitibi Model
Forest, Ontario, Lefort, P.;
Populus tremuloides 213 RC -- Canada Ontario's Old Trees

Vasiliauskas, S.
A.,; Ontario's Old
Populus balsamifera 207 RC -- Ontario, Canada Trees

Backus Woods, B. Larson;


Fagus grandifolia 204 RC -- Ontario, Canada Ontario's Old Trees

Xanthorrhoea preissii 200 XD Tree 41 Western Australia D. Ward

R. Guyette, M.
Fraxinus quadrangulata 194 RC BHY012 MO Stambaugh

Rocky Creek,
Shannon County,
Ulmus alata 186 RC -- Missouri S. Voelker

Carya tomentosa 169 XD PCT2025 Fentress Co., TN J. Hart

J. Hart and S. van


Quercus falcata 141 XD -- Knox County, TN de Gevel

Coweeta Hyrdologic
Fraxinus americana 136 XD -- Laboratory, NC S. Butler 2006

MOFEP Site 5,
Shannon County
Quercus coccinea 124 XD 511068 Missouri S. Voelker

502
Good Harbor Plains, T.C. Wyse, P.C.
Populus grandidentata 113 XD 83-2 MI Goebel; ITRDB

Dirca palustris 44 RC -- MO M. Stambaugh

503
Appendix C: Pith Indicators

504
Appendix D: Field Note Cards

Introduction
Various note cards can be used for efficient data collection in the field. Note cards are useful because they remind

the researcher of the variety of information that can be collected from a site for a particular project and the enable

uniform data collection for different research projects. The following pages present a variety of useful field note

cards that can be photocopied or edited for personal use. I recommend printing these on card stock paper, so that

they can stand up to hard use in the field and remain a more permanent collection of that field data. Many of the

cards are made to be printed as half page cards with information provided for the front and the back. These cards

are meant to be the starting point for your own cards that hold the information that you need for your particular

project. The Core Collection chart and Fire History Sample Cards are modified from formats used by the University

of Arizona Laboratory of Tree-Ring Research and the Dendrogeomorphology Sample Card is modified from

Shroder 1978 and was used at the University of Nebraska.

505
Core Collections

Site :___________________ Date :_____________ Page : ________________________


Field Crew : ___________________________________________________________________
Site Description:

Sample Coring
ID X Coord Y Coord Height Species Notes

506
Dendrogeomorphology Sample Cards

Species _________________________________ Specimen Number ______________________________

Collection Date _________________________ Site ID ______________________________________

Crown Density ___________________________ Field ________________________________________

Lean Direction __________________________ Lab __________________________________________

Lean Degree _____________________________ Geomorphic Feature ___________________________

Lean Characteristics ____________________ Collector ____________________________________

Height of Samples above base Height of Tree _______________________________

A _____________ B _____________ Species Density ______________________________

C _____________ D _____________ Slope Direction ______________________________

Diameter at sample location Slope Degree _________________________________

A _____________ B _____________ Photo Data ___________________________________

C _____________ D _____________ ______________________________________________

Section Orientation (azimuth, base side,

crown side, etc. ) ______________________

Drawing of tree (show sample locations and height Core or section data
of measurements).
A

507
Appendix E: Web Resources.

Please note that web address to frequently change. All attempts were made to make sure that all links in this

appendix were operational when this book went to print. When encountering a bad address, please search the site

name in a web search engine to find the page.

Site Name Web Address


Bibliography of http://www.wsl.ch/dbdendro/ April 30, 2009
Dendrochronology
Dencroclim2002 Program http://dendrolab.org/dendroclim2002.htm April 30, 2009
Eastern OLDLIST http://people.eku.edu/pedersonn/oldlisteast/ April 30, 2009
Henri Grissino-Mayer’s Ultimate http://web.utk.edu/~grissino/ April 30, 2009
Tree Ring Webpages
International Dendroecological http://www.wsl.ch/fieldevent/ April 30, 2009
Fieldweek
North American http://dendrolab.indstate.edu/nadef/index.htm April 30, 2009
Dendroecological Fieldweek
OLDLIST http://www.rmtrr.org/oldlist.htm April 30, 2009
PRECON Program http://www.ltrr.arizona.edu/webhome/hal/precon.html April 30, 2009
PRISM Data Set http://www.prism.oregonstate.edu/ April 30, 2009
Voortech Company http://www.voortech.com/projectj2x/docs/userGuide.htm April 30,
2009

508
 
aspen, 5 Tunguska, 5
Messiah, 5
 

509

You might also like