PART II - DRAFT 03.10.2010 20th Century Climatic Change and the

advertisement
PART II - DRAFT 03.10.2010
20th Century Climatic Change and the Instrumental Record
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
Throughout the western U.S., complex biophysical conditions (e.g., topography) are important factors
influencing highly variable responses to broad-scale climatic conditions. As a consequence, climatic
conditions vary across elevational, latitudinal and longitudinal gradients and at different spatial and
temporal scales. While trends in natural climatic variability are evident within the different climatic
regions it is important to keep in mind that broad-scale patterns in temperature, precipitation,
snowpack and other conditions also vary greatly across smaller spatial scales. Additionally, as a direct
measurement of climate conditions for the past century, the instrumental record provides some of the
best data of past climates yet the data from these records represents only the locales where
instruments are located and are often geographically biased. Thus, instrumental records provide direct
measurements of past conditions where they are located and often better represent certain elevations
and locals better than others (e.g., mid and low elevations vs. alpine environments).
Temperature
For all climate regions in this study, 20th century climate change is characterized by high spatial and
temporal variability. At broad temporal and spatial scales, however, it is possible to summarize trends
in climatic conditions impacting large portions of the four climate regions. Since 1900, temperatures
have increased in most areas of the western U.S. from 0.5-2˚ C (Fig. 19, Pederson et al. 2009, Mote et al.
2003, Ray et al. 2008) although there are a few exceptions where cooling has occurred (e.g.,
southeastern Colorado, Ray et al. 2008, and central Idaho and northwestern Montana, CIG). The rate of
change varies by location and elevation but is typically 1˚ C from early 20th century to present (Hamlet et
al. 2007). For most of the northern portions of the study area, temperatures generally increased from
1900 to 1940, declined from 1940-1975 and have increased from 1975 to present (USGCRP 2005).
Similarly, in the southern U.S. Rockies, temperatures generally increased in the 1930s and 1950s with a
period of cooling in the 1960s-70s and a consistent increase to present (Ray et al. 2008). Compared to
temperature increased in the early 20th century, temperatures for much of the study area doubled from
the mid 20th century to present, largely associated with rapid warming occurring since 1975.
Figure 19. 1950–2007 trend in observed annual average North American surface temperature (°C, left) and the
time series of the annual values of surface temperature averaged over the whole of North America (right). Annual
anomalies are with respect to a 1971–2000 reference. The smoothed curve (black line) highlights low frequency
variations. A change of 1°C equals 1.8°F. (Data source: UK Hadley Center’s CRUv3 global monthly gridded
temperatures). Source: Ray et al. 2008
1
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
Temperature increases are more pronounced during the cool season (Hamlet and Lettenmaier 2007)
and in the northern U.S. Rockies, are roughly triple that for the global average (Hall and Fagre 2003), a
pattern that is evident at northern latitudes and higher elevation sites throughout the western U.S.
(USGCRP 2005). Mean regional spring and summer temperatures for 1987 to 2003 were 0.87˚ C higher
than those for 1970 to 1986, and were the warmest since 1895 (Westerling et al. 2006). Additionally,
Bonfils et al. (2008) and Barnett et al. (2008) found a strong anthropogenic signal in the warming
observed warming trends in mountain areas across the west, suggesting that a portion of recent
observed warming is attributable to human influenced changes in greenhouse gas and aerosol
concentrations.
Precipitation
Trends in precipitation for the study area are much less clear. Instrumental data for much of the
northwestern U.S. show modest increases in precipitation during the past century (Fig. 20, Mote et al.
1999, 2003,2005) whereas records from parts of the southern Rockies do not show trends in
precipitation for the past century (Ray et al. 2008). Natural variability in precipitation is evident in the
instrumental record for all of the climate regions and long-term drought conditions in the past century
impacted large areas of the study area although 20th century droughts were not as severe as those
evident at other periods during the past millennium (Cook et al. 2007, 2004). For example, two
significant droughts in the 1930’s and 1950’s impacted much of the study area. The 1930’s drought was
more widespread and pronounced in the northern climate regions while the 1950’s drought was
centered more on the southcentral and southwestern U.S. (Cook et al. 2007, Gray et al. 2004, Fye et al.
2003) and research suggests that climatic conditions that influenced the nature and location of these
droughts is likely linked to low-frequency oscillations in ocean-atmospheric interactions (McCabe et al.
2004, Gray et al. 2007, Hidalgo 2004, Graumlich et al. 2003).
60
61
62
63
64
65
66
67
68
69
70
71
72
73
74
75
76
77
78
79
80
81
82
Figure 20. Trends in 1 Apr SWE over the 1960–2002
period of record directly from snow course observations.
Positive trends are shown in blue and negative in red, by
the scale indicated in the legend. Source: Mote et al.
2005 Journal of Climate
Surface Hydrology
Since 1950 more precipitation has been falling as
rain than snow, snowmelt and peak runoff is
occurring earlier and river flows are decreasing
during summer months (Fig. 21, Pederson et al.
2010, 2009, Mote et al. 2006, Barnett et al. 2008).
Recent impacts on snowpack and surface hydrology
are strongly associated with more precipitation
falling as rain than snow and earlier snowmelt
(Knowles et al. 2006). Comparisons of paleodata
and instrumental records with changing climate
conditions suggest that cool season precipitation is
more strongly associated with natural multidecadal, decadal and inter-annual variability in
ocean-atmosphere conditions than rising
temperatures.
2
83
84
85
86
87
88
89
90
91
92
93
94
95
96
97
98
99
100
101
102
103
104
105
106
107
108
109
110
111
112
113
114
115
116
117
118
119
Despite modest increases in precipitation in parts of the central and northern portions of the study area
(modest declines for parts of the southern U.S. Rockies), significant declines in snowpack are evident in
much of the northwestern U.S., especially in the northern U.S. Rockies and parts of the Upper Columbia
Basin ((Hamlet and Lettenmaier 2007, Gray et al. 2004, 2007, Pederson et al. 2004, 2010, Selkowitz et al.
2002, Mote et al. 2005, Mote et al. 2008). In Glacier National Park glaciers decreased nearly 30% by
1993 from their areal extent in 1850 (Hall and Fagre 2003) and of the 150 glaciers present in 1910, only
27 still exist (Fagre 2002). Much of the changes to surface hydrology of the western U.S. since 1950 can
be attributed to human-caused climate change related to increases in greenhouse gas emissions and
aerosols (Barnett et al. 2008, Bonfils et al. 2008). During the past century trends in drought conditions
have been increasing in central and southern parts of the study area and decreasing in northern areas of
the study area (Andreadis and Lettenmaier 2006) although drought conditions are expected to increase
for much of the study area in the future (Hoerling and Eischeid 2007).
Figure 21. Average winter (Dec-Feb; top),
spring (Mar-May; middle), and annual
(bottom) minimum temperatures from
SNOTEL (water year Oct-Sep) and valley
MET (calendar year Jan-Dec) stations.
Source: Mote et al. 2006
Ocean-Atmosphere Interactions
Ocean-atmosphere interactions (e.g.,
Pacific Decadal Oscillation, Atlantic
Multidecadal Oscillation, and El Niño
Southern Oscillation) are important
drivers of multi-decadal, decadal and
inter-annual variability in temperature
and precipitation (Pederson et al. 2010,
2009, Gray et al. 2007, 2004, 2003,
McCabe et al. 2004, Hidalgo 2004) but
their impacts vary greatly across latitudinal, elevational and longitudinal gradients. The Pacific decadal
oscillation (PDO) exhibits a cool and warm phase that typically last for 20-30 year intervals (Mantua et
al. 1997). During the 20th century several switches occurred between warm and cool phases and the
magnitude of PDO phases increased in the latter half of the past century (Fig. 22, McCabe et al. 2004,
Mantua et al. 2002, 1997).
120
121
122
123
124
125
126
127
128
129
130
Figure 22. Time series of the annual PDO and AMO.
Shaded areas indicate combinations of positive (+) and
negative (-) PDO and AMO periods. Source: McCabe et
al. 2004 PNAS
3
131
Ocean-atmosphere interactions - PDO and ENSO.
132
133
134
135
136
The Pacific Decadal Oscillation (PDO) and El Niño-Southern Oscillation (ENSO) are the predominant source of interdecadal (PDO) and inter-annual (ENSO) climate variability for much of the study area and the potential for
temperature and precipitation extremes increases when ENSO and PDO are in the same phases (CIG, other cits.).
Natural variations in PDO and ENSO are characterized by changes in sea surface temperature, sea level pressure,
and wind patterns (Mantua 1997, Wolter and Timlin 1993).
137
138
139
140
141
142
143
144
145
146
The PDO is described as being in one of two phases: a warm phase (positive index value) and a cool phase
(negative index value). Figure 1 shows the sea surface temperature (SST) anomalies that are associated with the
warm phase of PDO. The spatial patterns are very similar: both favor anomalously warm sea surface temperatures
near the equator and along the coast of North America, and anomalously cool sea surface temperatures in the
central North Pacific. The cool phases for PDO and ENSO, which are not shown, have the opposite patterns of SST
anomalies: cool along the equator and the coast of North America and warm in the central north Pacific. During
the 20th century, each PDO phase typically lasted for 20-30 years (Figure X previous page). Studies indicate that
the PDO was in a cool phase from approximately 1890 to 1925 and 1945 to 1977. Warm phase PDO regimes
existed from 1925-1946 and from 1977 to (at least) 1998. Pacific climate changes in the late 1990's have, in many
respects, suggested another reversal in the PDO (from "warm" to "cool" phase and possibly back to “warm”).
147
148
149
150
151
152
153
154
155
156
157
158
159
160
161
Figure 1. Warm Phase PDO and ENSO. The
spatial pattern of anomalies in sea surface
temperature (shading, degrees Celsius) and sea
level pressure (contours) associated with the
warm phase of PDO for the period 1900-1992.
Note that the main center of action for the
PDO (left) is in the north Pacific, while the main
center of action for ENSO is in the equatorial
Pacific (right). Contour interval is 1 millibar,
with additional contours drawn for +0.25 and
0.5 mb. Positive (negative) contours are
dashed (solid). Source: Climate Impacts Group,
University of Washington.
162
163
164
165
166
167
ENSO variations are commonly referred to as El Niño (the warm phase of ENSO) or La Niña (the cool phase of
ENSO). An El Niño is characterized by stronger than average sea surface temperatures in the central and eastern
equatorial Pacific Ocean, reduced strength of the easterly trade winds in the Tropical Pacific, and an eastward shift
in the region of intense tropical rainfall (Fig. 2). A La Niña is characterized by the opposite – cooler than average
sea surface temperatures, stronger than normal easterly trade winds, and a westward shift in the region of intense
tropical rainfall.
168
169
170
171
172
173
174
175
176
177
178
179
180
181
182
Natural variation in the strength of PDO and ENSO events impact climate regions in different ways. In the
northwestern U.S. and parts of the central U.S. Rockies, warm-phase PDO and El Niño winters tend to be warmer
and drier than average with below normal snowpack and streamflow whereas La Niña winters tend to be cooler
and wetter than average with above normal snowpack and streamflow (Graumlich et al. 2003, Cayan et al. 1999).
The southern U.S. Rockies and the southwestern U.S. respond differently, here warm-phase PDO and El Niño
winters tend to be wetter than average with above normal snowpack and streamflow and La Niña winters tend to
be drier than average with below normal snowpack and streamflow (Gray et al. 2004).
Figure 2. Multivariate ENSO index, 1950-2009.
Positive (red) index values indicate an El Niño event.
Negative (blue) values indicate a La Niña event
(Wolter and Timlin 1998, 1993).
4
183
184
185
186
187
188
189
190
191
192
193
194
195
196
197
198
199
200
201
202
203
204
205
206
207
208
209
210
211
212
213
214
215
216
217
218
219
220
221
222
223
224
225
226
227
228
229
230
231
As with other indices of ocean-atmosphere conditions, the PDO index influences precipitation differently
across the western U.S., with cool-season precipitation negatively correlated with a positive (warm)
phase PDO in the Upper Columbia Basin and northern U.S. Rockies and parts of the central U.S. Rockies
and GYA and positively correlated with the warm PDO in parts of the central and southern U.S. Rockies
(Fig. 23, Mote et al. 2005). Thus, the northwestern U.S. generally receives more winter precipitation
during when PDO index is low or negative and the southwestern U.S. receives more precipitation when
the PDO index is high or positive.
Fig 23. Relationships between two climate
indices, NPI and PDO, and 1 Apr SWE, over the
1960–2002 period of record. (a), (b) Correlations
are shown as red for negative and blue for
positive; circles indicate statistically significant
trends, and/or indicates insignificant trends. (c),
(d) The trend explained by regression with the
index, SWEX [see Eq. (3)], in units of cm as in Fig.
5. Source: Mote et al. 2005 Journal of Climate
Significant droughts appear to be linked to
complex interactions between PDO, AMO
and, to a lesser extent, variations in ENSO
although longer droughts likely result from
low-frequency oscillations in PDO and AMO.
Research indicates a link between the warmphase (positive) of AMO and past drought
with positive and negative phases of PDO
moderating the geographic center of
droughts (centered more in northwestern
U.S. when PDO is positive and southwestern
U.S. when PDO is negative (Fye 2003, Hidalgo
2004). For example, the dust bowl drought
was associated with a positive AMO and a
positive PDO and was centered more over
the southwestern U.S. whereas the 1950’s drought (positive AMO and PDO) was centered more over
Wyoming, Montana and the Canadian Rockies (Gray et al. 2004, Fye et al. 2003). Drought conditions in
the interior western U.S. are strongly associated with low-frequency variations in AMO and to a lesser
extent PDO (McCabe et al. 2004, Hidalgo 2004, Graumlich et al. 2003, Enfield et al. 2001) and these
variations appear more pronounced in the northern and central U.S. Rockies than parts of the
southwestern U.S. (Hidalgo 2004). In the coastal Pacific Northwest and southwestern U.S., variations in
precipitation and warm-season water availability appear more sensitive to low-frequency ENSO
variations than PDO and AMO although different combinations of these phases tend to amplify or
dampen ENSO signals in climatic and hydrologic records (Gray et al. 2007, 2004, McCabe et al. 2004,
Hidalgo 2004). While ocean-atmospheric interactions including ENSO and PDO are partially responsible
for variations in climatic conditions across the climate regions that are the focus of this synthesis,
research suggests that since the late 20th century, human influences, via increased greenhouse gas and
aerosol concentrations, are amplifying, dampening and, in some cases, overriding the influence of
natural variability of these phenomena (Barnett et al. 2008, Bonfils et al. 2008, McCabe et al. 2008, Gray
et al. 2006, 2003).
5
232
233
234
235
236
237
238
239
240
241
242
243
244
245
246
247
248
249
250
251
252
253
254
255
256
257
258
259
260
261
262
263
264
265
266
267
268
269
270
271
272
273
274
275
276
277
278
279
Changes in storm track and circulation patterns
Simulations of 21st century climate suggest a northward movement of the storm-track influencing
precipitation for much of the western U.S. (Yin 2005, Lorenz and DeWeaver 2007) which has the
potential to reduce precipitation for large parts of the study area. McAfee and Russell (2008) show that
a strengthening of the Northern Annual Mode, an index of sea level pressure poleward of 20˚ N, which
results in a poleward displacement of the Pacific Northwest stormtrack, increased zonal (west to east)
flow and reduced spring precipitation west of the Rocky Mountains and increased spring precipitation
east of the Rocky Mountains (McAfee and Russell 2008). This shift in the storm track is expected to
persist well into the future and may reduce the length of the cool-season, when circulation patterns
provide the bulk of precipitation for large areas of the central and northern parts of the study area
(McAfee and Russell 2008). If this becomes a more permanent shift in the storm-track position, this
phenomena could increase warm season (i.e., warm and dry) conditions for the Upper Columbia Basin,
northern U.S. Rockies and parts of the central U.S. Rockies whereas the central and southern U.S.
Rockies could receive increased spring season precipitation east of the Rockies. Changes in storm-track
position and circulation patterns will be superimposed on current trends, amplifying or dampening
changing conditions (e.g, reduced snowpack, earlier snowmelt and peak flows, diminished summer
flows, increased evapotranspiration) depending on location.
Ecological Impacts
Recent changes in climate conditions such as warming temperatures and associated declines in
snowpack and surface-hydrology are already influencing ecosystem dynamics. Several examples
illustrate the types of ecological impacts that have been observed in the past century including: earlier
spring bloom and leaf out of plant species, forest infill at and near treeline, and increased impact of
disturbances such as wildfire and insect outbreaks, all of which are likely to continue with increased
warming. Throughout much of the western U.S. spring bloom of a number of plant species has occurred
earlier, in some cases as much as several weeks (Cayan et al. 2001, Schwartz and Reiter 2000). In the
northern U.S. Rockies, increased density of trees at or near treeline has been observed at a number of
sites (Fagre et al. 2004). This “infill” phenomenon is not uncommon in the western U.S. and is predicted
to continue where minimum temperatures rise, snowpack in high-snowfall areas decreases and
moisture is not limiting (Graumlich et al. 2005, Lloyd and Graumlich 1997, Fagre et al. 2004, Rochefort et
al. 1994). While evidence for “infill” is widespread, upslope movement in treeline position is much more
variable and research suggests that upslope movement will be characterized by a high degree of spatial
heterogeneity in relation to variables that control treeline (Graumlich et al. 2005, Lloyd and Graumlich
1997).
Changing climate conditions are also influencing important disturbance processes that regulate
ecosystem dynamics. Warming temperatures, earlier snowmelt and increased evapotranspiration are
increasing moisture stress on forest species making them more susceptible to insect invasions. An
increase in the extent, intensity and synchronicity of mountain pine beetle attacks in the western U.S.
and Canada have been linked to forests stressed by dry intervals and are less able to resist beetle
infestations (Bentz et al. 2009, Hicke et al. 2006, Romme et al. 2006, Logan et al. 2003, Carroll et al.
2004, Breshears et al. 2005). Warming temperatures have also influenced bark beetle population
dynamics though reduced winter kill and have helped facilitate the reproduction and spread of these
insects (Black et al. 2010, Carroll et al. 2004). In some cases, past forest management (e.g., factors
related to structural characteristics of host stands) might also facilitate beetle infestation (Black et al.
2010, Bentz et al. 2009). Another example involves the area of the western U.S. burned by forest fires
annually. The extent of the western U.S. burned by fires each year is strongly linked to changing climate
conditions (Littell et al. 2009, 2008, Higuera et al. 2009, Morgan et al. 2008) and changes in surface
6
280
281
282
283
284
285
286
287
288
289
290
291
292
293
294
295
296
297
298
299
300
301
302
303
304
305
306
307
308
309
310
311
312
313
314
315
316
317
318
319
320
321
322
323
324
325
326
hydrology associated with reduced snowpack, earlier spring runoff and peak flows, and diminished
summer flows have been linked to increased frequency and duration of large fires and a lengthened fire
season in the western U.S., impacts where are most evident at mid-elevation forests in the northern U.S.
Rockies (Westerling et al. 2006). These are only a few examples of ecological impacts linked to recent
changes in climate, and all of these impacts are expected to become more pronounced in many parts of
the study area with future changes.
Northern U.S. Rockies
Temperature
Over the course of the 20th century, the instrumental record in the northern Rockies (NR) shows a
significant increase in average seasonal, annual, minimum and maximum temperatures (Figs. 24-25,
Loehman and Anderson 2009, Pederson et al. 2010, 2009). Regional average annual temperatures in
the Northern Rockies increased between 1-2C (Pederson et al. 2009, Pederson et al. submitted,
Loehman and Anderson 2009). Within this framework of increasing regional temperatures, seasonal
and annual minimum temperatures are generally increasing at a rate greater than that of the maximum
temperatures (Pederson et al. 2009, Pederson et al. submitted). In particular summer and winter
seasonal average minimum temperatures are increasing at a rate significantly greater than that of the
respective season’s average maximum temperatures, causing a pronounced reduction in the seasonal
daily temperature range (DTR) index (Pederson et al. 2009). The magnitude of minimum temperature
increases also appears seasonally variable: in area SNOTEL sites, Pederson et al. (submitted) estimated
minimum temperature increases since 1983 of +3.8 ± 1.72˚ C in winter (DJF), of +2.5 ± 1.23˚ C in spring
(MAM), and of +3.5 ± 0.73˚ C annually (Fig. 25). The magnitude of changes varies locally but there are
few exceptions to this general trend in warming.
Fig. 24 Comparison of variability and trends in western Montana (blue-green) and Northern Hemisphere (dark blue
line) annual average temperatures. A 5-year moving average (red line) highlights the similarity in trends and
decadal variability between records. Source: Pederson et al. 2010 Climatic Change
Temperature trends within the NR generally track Northern Hemisphere trends across temporal scales
(Fig. 24). This similarity between regional and continental scale trends suggests that large-scale climate
forcings such as greenhouse gases, patterns in sea surface temperatures (SSTs), volcanic activity, and
solar variability that drive decadal scale temperature global and continentally also drive regional
temperatures in the NR (Pederson et al. 2009).
7
352
353
354
355
356
357
358
359
360
361
362
363
364
365
366
367
368
369
370
371
372
373
374
327
328
329
330
331
332
333
334
335
336
337
338
339
340
341
342
343
344
345
346
347
348
349
350
351
Figure 25. SNOTEL station Tmin records have
been fit using a non-linear quadratic equation
due to characteristics of these time series. All
trends shown are significant (p ≤ 0.05) and
note the y-axis temperature scale changes for
each panel. Source: Pederson et al. submitted
Precipitation
When compared to the distinctive,
statistically significant trends present in
the 20th century temperature records for
the NR, no long-term (centennial scale)
trends are evident in the precipitation
time series data. Throughout the west,
high inter-annual, annual, and decadal
precipitation variability hinders trend
detection within the time series to derive
a consistent, centennial-scale trends in
precipitation that are statistically
significant (Ray et al. 2008). General
patterns throughout the latter part of the
20th century indicate that areas within the
NR experienced noticeable, albeit modest,
decreases in annual precipitation (Mote et al. 2005, Knowles et al. 2006). Although few statistically
significant long-term trends can be derived from regional 20th century precipitation time series data,
rising temperatures throughout the west have led to an increasing proportion of precipitation falling as
rain rather than snow (Knowles et al. 2006). However, due to regional winter temperatures averaging
significantly less than 0C, areas in the NR are generally less sensitive to shifts in precipitation type
spurred by rising temperature when compared to other regions in West (Knowles et al. 2006).
Surface Hydrology
In the NR, like most of the western United States, snow water equivalent (SWE) within a winter
snowpack largely controls surface runoff and hydrology for the water year and consequently has
significant impacts on water resources throughout the year (e.g. Pierce et al. 2009, Barnett et al. 2008,
Stewart et al. 2005, Pederson et al. submitted). Over the second half of the twentieth century, studies
have demonstrated a statistically significant decrease in winter snowpack SWE across the region
(Barnett et al. 2008, Pederson et al. submitted). Additionally, earlier onsets of springtime snowmelt and
streamflow have been documented (Stewart et al. 2005).
Ocean-atmosphere interactions
In the NR during the 20th century, the warm-phase (positive) of PDO is associated with reduced
streamflow and snowpack (Fagre et al. 2003) and the cool-phase of PDO is associated with increased
streamflow and snowpack (Fig. 26-27). Ecological responses are also evident in changes in the
distribution of mountain hemlock (Tsuga mertensiana). At high elevations growth of mountain hemlock
is limited by snowpack free days where a warm-phase PDO often results in decreased snowpack and
increased mountain hemlock growth (Fagre et al. 2003). The response is opposite at low elevation sites
8
375
376
377
378
397
398
399
400
401
402
403
404
405
406
407
408
409
410
411
412
413
414
415
416
417
418
419
420
421
422
423
424
where moisture is limiting and a warm-phase PDO commonly leads to less moisture which constrains
mountain hemlock growth and establishment (Fagre et al. 2003).
379
380
381
382
383
384
385
386
387
388
389
390
391
392
393
394
395
396
Figure 26. First EOF amplitudes representing spatially
broad patterns of (top) variability in winter wet-day
minimum temperature and (bottom) fraction of winter
precipitation falling as snow, with PDO phases indicated.
The thick, wavy lines are low-passfiltered versions of
each amplitude. The portion of the low-pass curve that
best describes a given period is used for that period.
Source:
Figure 27. Comparison of GNP summer drought, winter snowpack (May 1
SWE) and the Pacific Decadal Oscillation. All time series have been smoothed
using a 5 yr moving average. (a) Relationship between the average annual
instrumental PDO index (blue line, inverted for ease of comparison) [Mantua
and Hare, 2002] and May 1 SWE anomalies (red line) for GNP. (inset)
Correlations between winter (October–March) precipitation and the PDO index
for all U.S. climate divisions spanning 1949–2003
(http://www.cdc.noaa.gov/USclimate/Correlation/). (b) Reconstructed PDO
[D’Arrigo et al., 2001] used as a proxy for snowpack anomalies back to 1700.
Positive (negative) values of the reconstructed PDO correspond with low (high)
winter snowpack as indicated by the strong relationship with instrumental May
1 SWE anomalies (r = _0.688) for the common period of overlap (1922–1979).
(c) Summer drought (MSD) reconstruction for GNP. Source: Pederson et al.
2004 Geophysical Research Letters
Pederson et al. (submitted) summarize how variation in Pacific SSTs,
atmospheric circulation and surface feedbacks influence climate
conditions, snowpack and streamflow for the northern Rocky
mountains. Winters with high snowpack in the NRMs tend to be associated with negative PDO
conditions, a weakened Aleutian Low, and low pressure centered poleward of 45°N across western
North America (Fig. 28). During years of high snowpack, for example, the tendency is for mid-latitude
cyclones to track from the Gulf of Alaska southeast through the Pacific Northwest and into the NRMs.
The relatively persistent low-pressure anomaly centered over western North America is also conducive
to more frequent Arctic-air outbreaks resulting in colder winter temperatures. ENSO is also an
important driver of snowpack and streamflow at inter-annual time-scales, and the influence of related
tropical Pacific atmospheric circulation anomalies persists well into the spring.
9
425
426
427
428
429
430
431
432
433
434
435
436
437
438
461
462
463
464
465
466
467
468
469
470
Changes in spring (MAM) temperatures and precipitation are associated with changes in regional
atmospheric circulation, and are also shown to strongly influence the timing of NRM streamflow (Fig.
28). Springtime geopotential heights over western North America influence the amount, but more
importantly the timing, of snowmelt and streamflow across the northern U.S. Rockies. Specifically, high
pressure anomalies centered over western North America correspond with increased spring
temperatures and consequently the increasing number of snow-free days, early arrival of snow meltout, and streamflow CT. Atmospheric circulation changes occurring in March and April can, in turn,
initiate surface feedbacks that contribute to surface temperature and hydrograph anomalies (Fig. 28).
Hence, in the northern U.S. Rockies warming temperatures influence earlier snowmelt and runoff and
associated decreasing snowpack and streamflow but can also be partially attributed to seasonallydependant ocean-atmosphere teleconnections and atmospheric circulation patterns as well as surfacealbedo feedbacks that interact with broad-scale controls on snowpack and runoff (Pederson et al.
submitted).
439
440
441
442
443
444
445
446
447
448
449
450
451
452
453
454
455
456
457
458
459
460
Figure 28. Idealized relationship
between NRM snowpack and
streamflow anomalies with associated
Pacific SSTs, atmospheric circulation,
and surface feedbacks. Source:
Pederson et al. submitted
Central U.S. Rockies and GYA
Temperature
Temperatures for the central U.S. Rockies and the GYA (CR-GYA) have increased 1-2C during the past
century, with the greatest increases occurring in the latter half of the 20th century. Trends in
temperatures during the past century are slightly higher than the southwestern U.S. and slightly lower
than the northern U.S. Rockies, following a pattern of more pronounced temperature increases for
higher latitudes in the latter half of the 20th century (Cayan 2001, Ray et al. 2008). Increasing winter and
spring temperatures have resulted in reduced snowpack, earlier spring snowmelt and peak flows, and, in
10
471
472
473
474
475
476
477
478
479
480
481
482
483
484
485
486
487
488
489
490
491
492
493
494
495
496
497
498
499
500
501
502
503
some cases, lower summer flows for major basins in the climate region (Watson et al. 2009, Gray et al.
2007, 2004, 2003, Graumlich et al. 2003)
Precipitation
Records of precipitation for the CR-GYA show highly variable patterns across gradients in elevation,
latitude and longitude. No long-term trends over the past century are evident although patterns in
inter-annual, decadal and multi-decadal variation are evident in reconstructions of past hydrology from
tree-ring records (Fig. 29, Watson et al. 2009, Gray et al. 2007, 2004, Graumlich et al. 2003). A greater
proportion of precipitation is falling as rain versus snow in this climatic region but the impacts are less
pronounced than other parts of the western U.S. (Knowles et al. 2006). In many parts of the CR-GYA,
the 1930s and 1950s were significantly drier and the 1940s were wetter than average although subregional variation is high, likely associated with the location of this climate region in a transition
between Pacific Northwest and southwestern U.S. atmospheric circulation patterns discussed further
below (Watson et al. 2009, Gray et al. 2007, 2004, 2003, Graumlich 2003).
Figure 29. (a) Observed annual (previous July through current year June) precipitation for Wyoming Climate
Division 1 (gray line) vs. estimates of precipitation based on the stepwise regression (black line) model. (b) The full
stepwise version of reconstructed annual precipitation (black line) for AD 1173 to 1998. The horizontal line (solid
gray) near 400 mm represents the series mean, and the vertical line (dotted gray) at AD 1258 divides the wellreplicated portion of the record from reconstructed values in the earlier, less-replicated years (see Methods).
Source: Gray et al. 2007 Quaternary Research
Ocean-atmosphere interactions
The influence of ocean-atmospheric interactions on decadal, multi-decadal and inter-annual variation in
climatic conditions in the CR-GYA is more spatially variable than the other climate regions because the
region falls in a transition area between northwestern U.S. and southwestern U.S. circulation patterns
that strongly influence climatic conditions differently for these two areas of the western U.S. (Gray et al.
2007, 2004, Graumlich et al. 2003). Because the CR-GYA encompasses this transition between major
circulation types, a complex interaction between natural variations in ocean-atmospheric interactions
such as El Niño/Southern Oscillation (ENSO) and Pacific Decadal Oscillation (PDO) and topography,
11
504
505
506
507
508
509
510
511
512
513
514
515
516
517
518
519
520
521
522
523
524
525
526
527
528
529
530
531
532
533
534
535
536
537
538
539
540
541
latitude and longitude often result in opposite trends in climatic conditions at sites within the same
region (Gray et al. 2007).
Variation in precipitation and water availability in the CR-GYA is complex, largely due to the fact that CRGYA is influenced by both Southwest and Pacific Northwest modes of ocean-atmospheric variability and
is split between the winter-wet and summer-wet divisions outlined by Whitlock and Bartlein (1993).
High elevation snow basins within the central Rockies and western portions of the GYA typically
responds similar to the Pacific Northwest where the cool-phase (negative) PDO results in cool, wetter
than average winters and warmer and drier than average winter precipitation coincides with the warmphase (positive) PDO (Gray et al. 2007, 2004, Graumlich et al. 2003, Dettinger et al. 1998). Similar to the
Pacific Northwest, these portions of the climate region experience increased precipitation during La
Nina ENSO events and decreased precipitation during El Niño events, and ENSO seems to be linked to
the magnitude of PDO anomalies, especially so during winter months (Gray et al. 2007). Alternatively,
lower elevation sites and eastern portions of the GYA respond more like areas of the southwestern U.S.
or show little response to ENSO events (Gray et al. 2004). Here, years with strong El Nino SST’s result in
increased winter precipitation and drier conditions during La Nina events. This decoupling of high and
low and low elevation precipitation regimes is common throughout the central U.S. Rockies,
complicating predictions of future precipitation for the region (Woodhouse 2001).
Southern U.S. Rockies
Temperature
In the last 30 years temperatures have increased between 1-2 ˚F throughout the Southern U.S. Rockies.
The north central mountains of Colorado warmed the most, ~2.5 ˚F and high elevations have warmed
more quickly than lower elevations (Diaz and Eischeid 2007). Warming is evident at almost all locations
and temperatures have increased the most in the north central mountains and the least in the San Juan
Mountains of southwestern Colorado (Ray et al. 2008). Only the Arkansas River Valley in southeastern
Colorado show slight cooling trend during the 20th century and no trend is evident in this area for the
latter half of the century (Ray et al. 2008).
Precipitation
Records of precipitation for the Southern U.S. Rockies (SR) for the past century indicate highly variable
annual amounts and no long term trends are evident for the region (Ray et al. 2008, Dettinger et al.
2005). Like elsewhere in the interior west, a greater proportion of precipitation is falling as rain versus
snow than in the past but these changes are less pronounced than in the Northern U.S. Rockies
(Knowles et al. 2006). Decadal variability is evident in records of precipitation and surface flows and is
linked to variability in ocean-atmosphere and land-surface interactions (Fig. 30, Dettinger et al. 2005).
12
556
557
558
559
560
561
562
563
564
565
566
567
568
569
570
571
572
573
574
575
576
577
578
579
580
581
582
583
584
585
586
542
543
544
545
546
547
548
549
550
551
552
553
554
555
Fig 30. Observed time series (18952007) of annually averaged
precipitation departures areaaveraged over the Upper Colorado
drainage basin (top panel) and annual
Colorado River natural flow
departures at Lees Ferry in million
acre-feet (bottom panel). The
precipitation data are based on 4kmgridded PRISM data. Colorado River
natural flow data from the Bureau of
Reclamation. Source: Ray et al. 2008
Surface Hydrology
Paralleling trends evident throughout the interior west, more precipitation is falling as rain than snow,
spring snowpack is decreasing, especially at lower and mid-elevations below 2500 m and peak
streamflows are occurring earlier because of warmer spring temperatures (Knowles et al. 2006, Bales et
al. 2006, Stewart et al. 2005, Hamlet et al. 2005, Clow 2007, Mote 2006, 2003). Additionally, summer
flows are typically lower although annual flows show high variability but no significant trends in most
locations (Ray et al. 2008).
Ocean-atmosphere interactions
The Colorado River Basin spans an important transition area where the influence of Pacific Northwest
and southwestern U.S. circulation patterns show opposite trends (Gray et al. 2007, Clark et al. 2001).
Averages in SWE and annual runoff during El Nino years reflect this transition as northern parts of this
climate region experience drier-than-average conditions, whereas the southern portions experience
wetter-than-average conditions and the opposite conditions occur during La Nina years (i.e., wetter than
average in the north and drier than average in the south and west) and anomalies tend to be more
pronounced in spring in southern portions of the climate region. Long-term droughts are linked more
closely to low-frequency oscillations in PDO and AMO and are most commonly associated with the
interaction between a cool-phase (negative) PDO and warm-phase (positive) AMO (McCabe et al. 2004).
Upper Columbia Basin
Temperature
For most of the Upper Columbia Basin (UCB), average annual 20th century temperatures (data from
1920-2003) have increased by 0.7-0.8 °C and the warmest decade was the 1990s (Fig. 31, CIG 2010).
Average temperatures have increased as much as 2° C in parts of the climate region and increases have
been more pronounced at higher elevations (cits.). Winter and spring average temperatures and daily
minimum temperatures have increased more than other seasons and more than maximum
temperatures during the mid-20th century. During the latter half of the 20th century minimum and
maximum temperatures have increased at similar rates (Watson et al. 2009, Karl et al. 1993, Dettinger
et al. 1995, Lettenmaier et al. 1994, CIG 2010).
13
605
606
607
608
609
610
611
612
613
614
615
635
636
637
587
588
589
590
591
592
593
594
595
596
597
598
599
600
601
602
603
604
Figure 31. 20th century trends in (a) average annual PNW
temperature (1920-2000). This figure shows widespread
increases in average annual temperature for the period
1920 to 2000. The size of the dot corresponds to the
magnitude of the change. Pluses and minuses indicate
increases or decreases, respectively, that are less than the
given scale. Source: Climate Impacts Group, University of
Washington.
Precipitation
Trends in precipitation for the Upper Columbia Basin are less clear than trends in temperatures and
observations indicate high decadal variability. Precipitation has generally increased in the northwestern
U.S., by 14% for the entire region (1930-1995) and range between 13%-38% for different parts of the
region (Fig. 32, Mote et al. 2003) although trends are often not statistically significant depending on the
area and time interval measured. Paralleling trends observed for much of the interior western U.S.,
variability in winter precipitation has increased since 1973 (Hamlet and Lettenmaier 2007).
616
617
618
619
620
621
622
623
624
625
626
627
628
629
630
631
632
633
634
Figure 32. 20th century trends in average annual
precipitation (1920-2000). Increases (decreases) are
indicated with blue (red) dots. The size of the dot
corresponds to the magnitude of change. Source: CIG,
University of Washington.
14
638
639
640
641
642
643
644
645
646
647
648
649
650
670
671
672
673
674
675
676
677
678
679
680
681
682
683
684
685
686
Surface hydrology
Spring snowpack and snow water equivalent (SWE) declined throughout the UCB in the latter half of the
20th century and was most pronounced at low and mid-elevations (Figure 33, Mote et al. 2003, Hamlet
et al. 2005, Mote 2006). Declines in snowpack and SWE are associated with increased temperatures and
declines in precipitation over the same time period and declines of up to 40% or more are recorded for
some parts of the climate region (Mote et al, 2003, 2005). In addition to declines in snowpack, the
timing of peak runoff has shifted 2-3 weeks earlier for much for much of the region during the latter half
of the 20th century (Stewart et al. 2004) and the greatest shifts have occurred in the mountain plateaus
of Washington, Oregon and western Idaho (Hamlet et al. 2007). Because ecosystems in the Upper
Columbia Basin rely on the release of moisture from snowpack these shifts are significantly impacting
plant species which are blooming and leafing out earlier in the spring (Mote et al. 2005, Cayan et al.
2001, Schwartz and Reiter 2000).
651
652
653
654
655
656
657
658
659
660
661
662
663
664
665
666
667
668
669
Figure 33. 20th century trends in (b) April 1 snow
water equivalent (1950-2000). This figure shows
widespread increases in April 1 snow water equivalent
(an important indicator for forecasting summer water
supplies) for the period 1950 to 2000. The size of the
dot corresponds to the magnitude of the change.
Pluses and minuses indicate increases or decreases,
respectively, that are less than the given scale. Source:
CIG, University of Washington.
Ocean-atmosphere interactions
Variations in climatic conditions in the UCB are related to ocean-atmosphere and land-surface
interactions, namely El Niño/Southern Oscillation (ENSO) and Pacific Decadal Oscillation (PDO)
phenomena. In their warm phases (i.e., El Niño conditions for ENSO), both ENSO and PDO increase the
chance for a warmer winter and spring in the UCB and decrease the chance that winter precipitation will
meet historical averages. The opposite tendencies are true for cool phase ENSO (La Niña) and PDO: they
increase the odds that UCB winters will be cooler and wetter than average (Clark et al. 2001). While
typically warmer than average, SWE anomalies during strong El Niño are often less pronounced and
winter precipitations are commonly close to historical averages (Clark et al. 2001, CIG). Clark et al.
(2001) suggest that El Nino circulation anomalies are centered more in the interior west than for La Nina
circulation anomalies and are most evident in the middle of winter.
15
687
688
689
690
691
692
693
694
695
696
697
698
699
700
701
702
703
704
705
706
707
708
709
710
711
712
713
714
715
716
717
718
719
720
721
722
723
724
725
726
727
728
729
730
731
732
733
734
Lessons from the recent past: what can we learn from 20th century observations?
Small changes can have large impacts
Modest increases in temperature are already having dramatic impacts on surface hydrology through
earlier and diminished spring snowpack (Mote 2006, 2003, Stewart et al. 2005, 2004, Pederson et al.
submitted), increases in the proportion of winter precipitation as rain versus snow (Knowles et al. 2006,
Bales et al. 2006), decreased snow season length at most elevations (Bales et al. 2006), earlier blooming
dates for plants (Cayan et al. 2001, Schwartz and Reiter 2000) and lower summer flows (Meko 2007).
Changes in the distribution of minimum temperatures and frost-free days illustrate how small changes
in temperature may have large changes to surface hydrology (Barnett et al. 2004, Stewart et al. 2004).
Evidence from a number of studies suggests that even small increases in temperatures are expected to
have dramatic impacts on water availability for much of the western U.S. Along with changes in
snowpack and earlier spring runoff, modest temperature increases predicted for the future suggest that
drought conditions will become much more common (Hoerling and Eischeid 2007, Seager et al. 2007,
Barnett and Pierce 2009). Increased winter precipitation that is predicted for central and northern
areas of the study area will likely not be adequate to offset increased rates of evapotranspiration,
leading to increased number of dry days and increased frequency (and possibly intensity) of droughts for
much of the interior west (Gray et al. 2009, Hoerling and Eischeid 2007, Seager et al. 2006).
Shifting distributions and new norms
The western U.S. is vulnerable to small changes in temperature and drier summers because much of the
western U.S. consists of arid to semi-arid ecosystems that depend on limited water resources regulated
by dynamics of mountain snowpack (Gray et al. 2009). While many ecosystems are adapted to natural
variations in water availability, a shift in the distribution of drought like conditions where they become
the norm and the frequency and magnitude of dry conditions exceed historical patterns could result in a
tipping point where major redistribution of vegetative communities ensues (Fig. 34, Jackson et al. 2009).
Rapid changes in the storage and distribution of surface flows related to snowpack could have significant
impacts on ecosystems adapted to past snowpack dynamics that may change to the point that they
exceed recent natural ranges of variation and ecosystem thresholds where adaptation and/or
fundamental transitions occur (Fig. 34).
Figure 34. Idealised version of a
coping range showing the
relationship between climate
change and threshold
exceedance, and how adaptation
can establish a new critical
threshold, reducing vulnerability
to climate change (modified from
Jones and Mearns 2005). Source:
IPCC 2007.
Small increases in temperatures are already resulting in greater evaporative losses from lakes, streams,
wetlands and from terrestrial ecosystems and modest increases in winter precipitation that some
models predict for the northern parts of the study area are not expected to keep pace with and offset
increased evapotranspiration (Gray et al. 2009, Arnell 1999). For example, models from Gray and
16
735
736
737
738
739
740
741
742
743
744
745
McCabe (in review) estimate a 15-25% decrease in water in the Yellowstone River in the future and,
even with increased precipitation it would take the wettest years of last 800 years to offset increased
evapotranspiration associated with warmer temperatures (Gray and McCabe in review). Additionally,
the frequency of extreme episodes of precipitation and temperature has occurred in the past century
(Karl et al. 2009, IPCC 2007, Groisman et al. 2005, Kunkel et al. 2003, Madsen and Figdor 2007) and
shifts in the distribution of climate conditions (e.g., average summer temperatures) will likely result in
increased occurrence of extreme conditions when compared with 20th century conditions (Fig. 35).
Research indicates that delivery of precipitation in many parts of the western U.S. will occur as extreme
events and will result in more dry days as there will be longer intervals without precipitation (Groisman
et al. 2005, Kunkel et al. 2003, Madsen and Figdor 2007).
746
747
748
749
750
751
752
Figure 35. Schematic showing the effect
on extreme temperatures when the mean
temperature increases, for a normal
temperature distribution. Source: IPCC
2007
753
754
755
756
What can we expect in the future?
757
758
759
760
761
762
763
764
765
766
767
768
769
770
771
772
773
Many of the trends in climate evident in the past century are expected to continue in the future.
Temperatures are expected to increase across the landscape, although short-term variation is still
expected. An example of short-term variation that some point to as a reversal of warming temperatures
relates to a modest cooling that much of the Northern Hemisphere experienced in the last decade, a
phenomenon which has now been largely attributed to a temporary decrease in atmospheric
concentrations of water vapor in the lower stratosphere (Solomon et al. 2010). This short-term cooling
illustrates of how short-term variations will be superimposed on longer-term trends. Changes in the
amount and spatial distribution of precipitation are related to a complex interaction of global circulation
patterns, ocean- and land surface -atmosphere interactions that are still poorly understood.
Additionally, the influence of human activities on natural variations appears to be increasing (Barnett et
al. 2008, Bonfils et al. 2008). While models generally predict increased precipitation for parts of the
northern and central U.S. Rockies and the Upper Columbia Basin and decreased amounts for the
southwestern U.S., both uncertainty and variability is expected to be high and is reflected in model
predictions. Levels of uncertainty are a critical component of any prediction of future conditions and
careful consideration of how levels vary should inform how managers anticipate future conditions.
Scenario planning can be an effective approach for considering future conditions that area highly
uncertain, an approach discussed in more detail later.
774
775
Model projections of future conditions
776
777
Analysis of future projections for temperature, precipitation, snowpack, stream flow, drought, growing
season. Jeremy Littell will provide figures representing wall-to-wall downscaled models (VIC model
17
778
779
780
781
forecasts) for key variables (temperature, precipitation, soil moisture, snowpack, growing season and
possibly others e.g. extreme events?) for most of our study area. He will also provide some text to
accompany these figures and, if possible, text discussing key sources of uncertainty associated with
future projections.
782
783
Figures of future projections for a few key variables (temperature, precipitation, soil moisture, snowpack
or streamflow, growing season) for each of the climate regions and text on uncertainty.
784
1. Northern U.S. Rockies
785
2. Central U.S. Rockies and the GYA
786
3. Southern U.S. Rockies
787
4. Upper Columbia Basin
788
789
Ecological Response: Area burned example
790
791
792
Phil Higuera is interested in providing predictions for future area burned under different climate
scenarios. This is new modeling that he has been working on and would provide an example of
projected ecological (process) response to biophysical changes identified in the VIC and other models.
793
794
795
796
797
798
799
800
801
802
816
817
818
819
820
821
Trends that will likely continue to impact large parts of the study area
Climate conditions
 2-6 + C degrees in temperature and higher at higher latitudes
 Increased but highly variable precipitation for parts of the Upper Columbia Basin, northern and
central U.S. Rockies.
 Decreased but highly variable precipitation for parts of the central and southern U.S. Rockies.
 Increased evapotranspiration for most of the western U.S. which will likely not be offset by
increased precipitation (Fig. 36, Hoerling and Eischeid 2007, Seager et al. 2006).
803
804
805
806
807
808
809
810
811
812
813
814
815
Fig. 36. Modeled changes
in annual mean
precipitation minus
evaporation over the
American Southwest
(125°W to 95°W and 25°N
to 40°N, land areas only),
averaged over ensemble
members for each of the
19 models. The historical
period used known and
estimated climate
forcings, and the
projections used the SResA1B emissions scenario. The median (red line) and 25th and 75th percentiles (pink
shading) of the P − E distribution among the 19 models are shown, as are the ensemble medians of P (blue
line) and E (green line) for the period common to all models (1900–2098). Anomalies (Anom) for each model
are relative to that model's climatology from 1950–2000. Results have been 6-year low-pass Butterworthfiltered to emphasize low-frequency variability that is of most consequence for water resources. The model
ensemble mean P − E in this region is around 0.3 mm/day. Source: Seager et al. 2007 Science
18
822
823
824
825
826
827
828
829
830
831
832
833
834
Surface Hydrology
 Greater proportion of winter precipitation falling as rain than snow (Knowles et al. 2006, Bales
et al. 2006, Dettinger et al. 2004)
 Decreased snow season length at most elevations Bales et al. 2006
 Less spring snowpack (Pederson et al. submitted, Mote 2006, 2005, 2003)
 Earlier snowmelt runoff and peak streamflows (Stewart et al. 2005, 2004, Hamlet et al. 2005,
Clow 2007).
 Increased frequency of droughts and low summer flows (Gray et al. 2009, Meko et al. 2007).
Increased evapotranspiration that will likely not be offset even where precipitation increases
amplifying dry conditions (Hoerling and Eischeid 2007, Seager et al. 2006).
 Underground recharge may decline with declines in snowpack (Winnograd et al. 1998)
 With poles getting wetter and subtropics getting drier variability in mid-latitude precipitation is
expected to increase (Dettinger et al. 2004)
835
836
837
838
839
840
841
842
Extreme conditions: Drought, Floods, Heat Waves
 More episodes of extreme temperatures (US Global Change report, Karl et al. 2009)
 Increased frequency of extreme precipitation (storm) events, rain on snow and consequent
winter/spring floods in mountains (Madsen and Figdor 2007, Groisman et al. 2005, Kunkel et al.
2003).
 Droughts are expected to become more frequent as a result of increased temperatures,
evapotranspiration and changes to surface hydrology that impact warm season conditions (Gray
et al. 2009, Meko 2007).
843
844
845
846
847
848
849
850
851
852
Circulation patterns
 A long-term trend in the northward shift in the winter storm track and jet stream will likely
result in more winter/autumn precipitation for the northwestern U.S. and less for southern
Rockies and Southwest U.S. although variability is expected to by high (McAfee and Russell
2009).
 Natural variation in ocean-atmosphere interactions (e.g., PDO and ENSO) will continue to
influence climate in the western U.S. but human drivers of change (i.e., greenhouse gases and
aerosol concentrations) are now interacting with these natural variations and are dampening,
amplifying and, in some cases, overriding these natural drivers of change (Meehl et al. 2009,
Barnett et al. 2008, Bonfils et al. 2008).
853
854
855
Productivity – Phenology
 Earlier blooming dates for many plant species (Cayan et al. 2001, Schwartz and Reiter 2000)
 Longer growing season (Bates et al. 2006)
856
857
858
859
Disturbance
 Increased impact of disturbances linked to drought stress. Examples include, increased large
fires resulting from increased temperatures, changes in surface hydrology and snowpack and
conditions during the warm season (Westerling et al. 2006)
19
860
861
862
863
864
865
866
867
868
869
870
871
872
873
874
875
876
877
878
879
880
881
882
883
884
885
886
906
907


Higher frequency of large-fires, longer fire season and increased area of western U.S. burned by
fire (Westerling et al. 2006, Littell et al. 2009, 2008, Higuera et al. 2009, Morgan et al. 2008)
Greater drought stress will likely result in more insect infestations and disease affecting forests
(Black et al. 2010, Bentz et al. 2009, Hicke et al. 2006, Romme et al. 2006, Logan et al. 2003,
Carroll et al. 2004, Breshears et al. 2005).
Planning for the future
I have made an initial attempt to adapt a part of Jackson et al. 2009 for this section but it needs further modification (Perhaps
Steve could look more closely at this).
Planning for future conditions that are highly uncertain presents a significant challenge for land
managers. Scenario planning provides an approach for preparing for future climate conditions that are
highly uncertain by anticipating a range of future conditions. Scenario planning uses a combination of
scientific input, expert opinion and forecast data to develop alternative scenarios for the future
(Schwartz 1991, van der Heijden 1996), contrasting with more traditional attempts at developing
precise, quantitative assessments of future conditions, which are often useless because of compounded
uncertainties. In scenario planning, alternative scenarios can be used as a starting point for exploring
species or ecosystem vulnerabilities under a range of future conditions, and as a means for examining
how management strategies might play out in the face of multiple drivers of change.
Jackson et al. (2009) developed an example to illustrate this process. In this example, alternative futures
can be arrayed along two axes comprising integrators of potential climate-change (drought frequency)
and potential changes in disturbance regimes (fire size). In concert with monitoring and modeling,
studies of past climates can define the range of drought frequency we might reasonably expect, and
past studies of fire can place bounds on potential fire size. This exercise yields four quadrants, each
comprising a distinct combination of climatic and fire-regime change (Fig. 37). These quadrants each
provide a contrasting scenario that can be used to explore potential impacts on species or ecosystems
and for examining the relative costs and benefits of various mitigation and adaptation measures.
887
888
889
890
891
892
893
894
895
896
897
898
899
900
901
902
903
904
905
Figure 37. Example of a scenario planning matrix.
Each axis represents a critical driver of system
change or a significant trend in the environment.
In common practice, the variables chosen for
analysis are likely to have the strongest influence
on the system or they are associated with a high
degree of uncertainty (Shoemaker 1995). In the
case presented here, the axes represent a
continuum between conditions that are similar to
those observed in the historical record and
conditions that are significantly altered from those
seen today. Combining these two drivers produces
four alternative scenarios for the future conditions
(e.g., frequent drought and large fires in the upper
right) that can then be further developed into
“storylines” that provide details about how each
scenario might unfold. Depending on the
application and available data, axes and the
resulting storylines may be defined quantitatively,
or they may be based on qualitative assessments alone. Source: Jackson et al. 2009 Paleontological Society Papers
20
908
909
910
911
912
913
914
915
916
917
918
919
920
921
922
923
924
925
926
927
928
929
930
931
932
933
934
935
936
937
938
939
940
941
942
943
944
945
946
947
948
949
950
951
952
953
954
955
956
957
958
959
At one extreme, major climate change and altered disturbance regimes interact to drive emergence of
novel ecosystems. Given limited experience with ecosystem turnover in many of the climate regions,
consideration of long-term paleoenvironmental records serves as a primary means for adding texture
and substance to the scenario. It also points out the risk of finding ourselves in any one of the four
quadrants. For example, transition to “novel ecosystems” is analogous to the transition observed
11,000 yr BP at Yellowstone while the transitions to “inevitable surprises” and “patches and fragments”
are analogous to the late Holocene transitions observed in (Fig. 38).
Figure 38. Multiple time scales of vegetation change. At the 15 year scale, a decline in the area of living lodgepole
pine forest in Colorado and Wyoming has arisen from a mountain pine beetle infestation; based on USFS data. At
the 150 years scale, the percent of a 129,600-ha subalpine study area in central Yellowstone occupied by three
different types of forest stages, recently burned (dashed line), even-aged (gray line), and all-age mixed stands
(black line), show successional and disturbance influences on forest structure (Romme and Despain, 1989). The
same data are shown at the 1500-year scale (extending back >250 yrs before AD 1950), and are compared with the
percent pine (Pinus) pollen in pollen records from Yellowstone region, which show limited plant assemblages
changes at that scale. More dramatic changes in plant assemblages are evident from the percent of pine and
sagebrush (Artemisia) pollen over 15,000 years from the same sites. Gray arrows indicate different types of
ecosystem changes. Pollen records shown here include the data used for climate reconstruction in Fig. 4 from
Blacktail Pond, black line (Gennett and Baker, 1986); data from location of charcoal record in Fig. 4 from Slough
Creek Pond, thick gray line (Millspaugh, 1997); data from the Romme and Despain (1989) study area from
Buckbean Fen, short dashed line (Baker, 1976); Emerald Lake, long dashed line (Whitlock, 1993); and other
datasets from Cygnet Lake Fen, gray line (Whitlock, 1993); Hedrick Pond, gray line (Whitlock 1993). Source: Jackson
et al. 2009
21
960
961
962
963
964
965
966
The greatest value in scenario planning comes from uncovering vulnerabilities and potential responses,
particularly those common to multiple story lines…
Could spend more time discussing novel climates and novel communities (Jackson et al. 2009 PNAS)
Final summary…
22
References (incomplete)
Aber J., Neilson R.P., McNulty S., Lenihan J.M., Bachelet D. and Drapek R.J. 2009. Forest Processes and
Global Environmental Change: Predicting the Effects of Individual and Multiple Stressors. Bioscience
51: 735-751.
Allen C.D. and Breshears D.D. 1998. Drought-induced shift of a forest-woodland ecotone: Rapid
landscape response to climate variation. Proceedings of the National Academy of Sciences 95:
14839-14842.
Alley R.B. and Ágústsdóttir A.M. 2005. The 8k event: cause and consequences of a major Holocene
abrupt climate change. Quaternary Science Reviews 24: 1123-1149.
Ammann C.M., Joos F., Schimel D.S., Otto-Bliesner B.L. and Tomas R.A. 2007. Solar influence on climate
during the past millennium: Results from transient simulations with the NCAR Climate System
Model. Proceedings of the National Academy of Sciences 104: 3713-3718.
Anderson R.S., Allen C.D., Toney J.L., Jass R.B. and Bair A.N. 2008. Holocene vegetation and fire regimes
in subalpine and mixed conifer forests, southern Rocky Mountains, USA. International Journal of
Wildland Fire 17: 96-114.
Andreadis K.M. and Lettenmaier D.P. 2006. Trends in 20th century drought over the continental United
States. Geophys. Res. Lett. 33: L10403.
Arnell N.W. 1999. The effect of climate change on hydrological regimes in Europe: a continental
perspective. Global Environmental Change 9: 5-23.
Bailey R.G. 1995. Description of the ecoregions of the United States. Miscellaneous Publication 1391.
USDA Forest Service, Washington D.C.
Barnett T.P., Pierce D.W., Hidalgo H.G., Bonfils C., Santer B.D., Das T., Bala G., Wood A.W., Nozawa T.,
Mirin A.A., Cayan D.R. and Dettinger M.D. 2008. Human-Induced Changes in the Hydrology of the
Western United States. Science 319: 1080-1083.
Barnosky C.W. 1985. Late Quaternary vegetation in the Southwestern Columbia Basin, Washington.
Quaternary Research 23: 109-122.
Bartlein P.J., Anderson K.H., Anderson P.M., Edwards M.E., Mock C.J., Thompson R.S., Webb R.S., Webb
III T. and Whitlock C. 1998. Paleoclimate simulations for North America over the past 21,000 years:
features of the simulated climate and comparisons with paleoenvironmental data. Quarternary
Science Reviews 17: pp. 549-585.
Beiswenger J.M. 1991. Late Quaternary Vegetational History of Grays Lake, Idaho. Ecological
Monographs 61: 165-182.
Beniston M., Diaz H.F. and Bradley R.S. 1997. Climatic Change at High Elevation Sites; A Review. Climatic
Change 36: 233-251.
Beniston M. and Stephenson D.B. 2004. Extreme climatic events and their evolution under changing
climatic conditions. Global and Planetary Change 44: 1-9.
Bentz B., C.D. Allen, M. Ayres, E. Berg, A. Carroll, M. Hansen, J. Hicke, L. Joyce, J. Logan, W. MacFarlane,
J., MacMahon S.M., J. Negrón, T. Paine, J. Powell, K. Raffa, J. Régnière, M. Reid, W. Romme, S.
Seybold, and D. Six D.T., J. Vandygriff, T. Veblen, M. White, J. Witcosky and D. Wood 2009. Bark
Beetle Outbreaks in Western North America: Causes and Consequences. University of Utah Press.
23
Betancourt J.L. 1990. Late Quaternary biogeography of the Colorado Plateau. In Betancourt J. L., Van
Devender, T.R. and Martin, P.S (ed.), Packrat middens - the last 40,000 years of biotic change, pp.
259-292. University of Arizona Press, Tucson, Arizona.
Biondi F., Perkins D.L., Cayan D.R. and Hughes M.K. July Temperature During the Second Millennium
Reconstructed from Idaho Tree Rings. Geophys. Res. Lett. 26.
Bitz C.M. and Battisti D.S. 1999. Interannual to decadal variability in climate and glacier mass balance in
Washington, Western Canada, and Alaska. American Meteorological Society 12: 3181-3196.
Bond G., Kromer B., Beer J., Muscheler R., Evans M.N., Showers W., Hoffmann S., Lotti-Bond R., Hajdas I.
and Bonani G. 2001. Persistent Solar Influence on North Atlantic Climate During the Holocene.
Science 294: 2130-2136.
Bonfils C., Santer B.D., Pierce D.W., Hidalgo H.G., Bala G., Das T., Barnett T.P., Cayan D.R., Doutriaux C.,
Wood A.W., Mirin A. and Nozawa T. 2008. Detection and Attribution of Temperature Changes in the
Mountainous Western United States. Journal of Climate 21: 6404-6424.
Booth R., Kutzbach J., Hotchkiss S. and Bryson R. 2006. A reanalysis of the relationship between strong
westerlies and precipitation in the Great Plains and Midwest regions of North America. Climatic
Change 76: 427-441.
Bradley B.A., Oppenheimer M. and Wilcove D.S. 2009. Climate change and plant invasions: restoration
opportunities ahead? Global Change Biology 15: 1511-1521.
Breshears D.D., Cobb N.S., Rich P.M., Price K.P., Allen C.D., Balice R.G., Romme W.H., Kastens J.H., Floyd
M.L., Belnap J., Anderson J.J., Myers O.B. and Meyer C.W. 2005. Regional vegetation die-off in
response to global-change-type drought. Proceedings of the National Academy of Sciences of the
United States of America 102: 15144-15148.
Brockway D.G. 1998. Forest Plant Diversity at Local and Landscape Scales in the Cascade Mountains of
Southwestern Washington. Forest Ecology and Management 109: 323-341.
Broecker W.S. and Denton G.H. 1989. The role of ocean-atmosphere reorganizations in glacial cycles.
Geochimica et Cosmochimica Acta 53: 2465-2501.
Brohan P., Kennedy J.J., Harris I., Tett S.F.B. and Jones P.D. 2006. Uncertainty estimates in regional and
global observed temperature changes: A new data set from 1850. J. Geophys. Res. 111: D12106.
Brown D.P. and Comrie A.C. 2004. A winter precipitation 'dipole' in the western United States associated
with multidecadal ENSO variability. Geophysical Research Letters 31: -.
Brown T.B., Barry R.G. and Doesken N.J. 1992. An Exploratory Study of Temperature Trends for Colorado
Paired Mountain-High Plains Stations. In Amer. Met. Soc. Sixth Conference on Mountain
Meteorology, pp. 181-184, Portland, OR.
Carrara P.E. 1989. Late Quaternary glacial and vegetative history of the Glacier National Park region,
Montana. U.S. Geological Survey Bulletin 1902: 64.
Carroll A.L., S.W. Taylor, J. Regniere and L. Safranyik 2004. Effects of climate change on range expansion
by the mountain pine beetle in British Columbia. In Shore T., Brooks J. E. and Stone J. E. (eds.),
Mountain Pine Beetle Symposium: Challenges and Solutions, pp. 223-232. Natural Resources
Canada, Canadian Forest Service, Pacific Forestry Centre, Kelowna, B.C.
Cayan D.R., Dettinger M.D., Diaz H.E. and Graham N.E. 1998. Decadal Variability of Precipitation over
Western North America. Journal of Climate 11: 3148.
24
Cayan D.R., Kammerdiener S.A., Dettinger M.D., Caprio J.M. and Peterson D.H. 2001. Changes in the
onset of spring in the western United States. Bulletin of the American Meteorological Society 82:
399-415.
Chang P. and Battisti D.S. 1998. The Physics of El Nino. Physics World 8: 41-47.
Clark M.P., Serreze M.C. and McCabe G.J. 2001. Historical effects of El Nino and La Nina events on the
seasonal evolution of the montane snowpack in the Columbia and Colorado River Basins. Water
Resources Research 37: 741-758.
Cole J.E., Overpeck J.T. and Cook E.R. 2002. Multiyear La Nina events and persistent drought in the
contiguous United States. Geophysical Research Letters 29: -.
Committee on Surface Temperature Reconstructions for the Last 2 Y., National Research Council 2006.
Surface Temperature Reconstructions for the Last 2,000 Years. The National Academies Press,
Washington, D.C.
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.
Crowley and Lowery 2000. Causes of climate change over the last 1000 years. Science 289: 270-277.
Dale V.H., Joyce L.A., McNulty S., Neilson R.P., Ayres M.P., Flannigan M.D., Hanson P.J., Irland L.C., Lugo
A.E., Peterson C.J., Simberloff D., Swanson F.J., Stocks B.J. and Wotton B.M. 2001. Climate change
and forest disturbances. Bioscience 51: 723-734.
Davis O.K. and Shafer D.S. 1992. A Holocene climatic record for the Sonoran Desert from pollen analysis
of Montezuma Well, Arizona, USA. Palaeogeography, Palaeoclimatology, Palaeoecology 92: 107-119.
Dettinger M.D. and Cayan D.R. 1995. Large-scale atmospheric forcing of recent trends toward early
snowmelt runoff in California. Journal of Climate 8: 606-623.
Dettinger M.D. and Ghil M. 1998. Seasonal and interannual variations of atmospheric CO2 and climate.
Tellus 50b: 1-24.
Diaz H.F. and Eischeid J.K. 2007. Disappearing “alpine tundra” Köppen climatic type in the western
United States. Geophys. Res. Lett. 34: L18707.
Enfield D.B., Mestas-Nunez A.M. and Trimble P.J. 2001. The Atlantic multidecadal oscillation and its
relation to rainfall and river flows in the continental US. Geophysical Research Letters 28: 20772080.
Esper J., Cook E.R. and Schweingruber F.H. 2002. Low-Frequency Signals in Long Tree-Ring Chronologies
for Reconstructing Past Temperature Variability. Science 295: 2250-2253.
Fagre D.B., Peterson D.L. and Hessl A.E. 2003. Taking the Pulse of Mountains: Ecosystem Responses to
Climatic Variability. Climatic Change 59: 263-282.
Fall P.L. 1997. Timberline fluctuations and late Quaternary paleoclimates in the Southern Rocky
Mountains, Colorado. Geological Society of America Bulletin 109: 1306-1320.
Floyd M.L., Clifford M., Cobb N.S., Hanna D., Delph R., Ford P. and Turner D. 2009. Relationship of stand
characteristics to drought-induced mortality in three Southwestern pinon-juniper woodlands.
Ecological Applications 19: 1223-1230.
Garfin G.M. and Hughes M.K. 1996. Eastern Oregon Divisional Precipitation and Palmer Drought Severity
Index from Tree-Rings. Cooperative Agreement PNW 90-174. USDA Forest Service, Tucson, AZ.
25
Giorgi F., Hurrell J.W., Marinucci M.R. and Beniston M. 1997. Elevation Dependency of the Surface
Climate Change Signal: A Model Study. Journal of Climate 10: 288-296.
Gosse J.C., Evenson, E.B., Klein, J., Lawn, B., and and Middleton R. 1995. Precise cosmogenic 10Be
measurements in western North America: Support for a global Younger Dryas cooling event.
Geology 23: 877-880.
Graumlich L.J., Pisaric M.F.J., Waggoner L.A., Littell J.S. and King J.C. 2003. Upper Yellowstone River Flow
and Teleconnections with Pacific Basin Climate Variability during the Past Three Centuries. Climatic
Change 59: 245-262.
Gray S. and Andersen C. 2009. Assessing the Future of Wyoming’s Water Resources: Adding Climate
Change to the Equation Environment and Natural Resources. University of Wyoming, Laramie, WY.
p. 28.
Gray S.T., J.L. Betancourt, S.T. Jackson and R.G. Eddy 2006. Roll of multidecadal climate variability in a
range extension of pinyon pine. Ecology 87: 1124-1130.
Gray S.T., Betancourt J.L., Fastie C.L. and Jackson S.T. 2003. Patterns and sources of multidecadal
oscillations in drought-sensitive tree-ring records from the central and southern Rocky Mountains.
Geophysical Research Letters 30: 1316, doi:10.1029/2002GL016154.
Gray S.T., Fastie C.L., Jackson S.T. and Betancourt J.L. 2004. Tree-Ring-Based Reconstruction of
Precipitation in the Bighorn Basin, Wyoming, since 1260 a.d. Journal of Climate 17: 3855-3865.
Gray S.T., Graumlich L.J. and Betancourt J.L. 2007. Annual precipitation in the Yellowstone National Park
region since AD 1173. Quaternary Research 68: 18-27.
Groisman P.Y., Knight R.W., Easterling D.R., Karl T.R., Hegerl G.C. and Razuvaev V.N. 2005. Trends in
Intense Precipitation in the Climate Record. Journal of Climate 18: 1326-1350.
Hall M.P. and Fagre D.B. 2003. Modeled climate-induced glacier change in Glacier National Park, 18502100. Bioscience 53: 131-140.
Hamlet A.F. and Lettenmaier D.P. 2007. Effects of 20th century warming and climate variability on flood
risk in the western U.S. Water Resources Research 43: W06427.
Hamlet A.F., Mote P.W., Clark M.P. and Lettenmaier D.P. 2005. Effects of temperature and precipitation
variability on snowpack trends in the western United States. Journal of Climate 18: 4545-4561.
Hamlet A.F., Mote P.W., Clark M.P. and Lettenmaier D.P. 2007. Twentieth-century trends in runoff,
evapotranspiration, and soil moisture in the western United States. Journal of Climate 20: 14681486.
Harrison S.P., Kutzbach J.E., Liu Z., Bartlein P.J., Otto-Bliesner B., Muhs D., Prentice I.C. and Thompson
R.S. 2003. Mid-Holocene climates of the Americas: a dynamical response to changed seasonality.
Climate Dynamics 20: 663-688.
Hasselmann K. 1976. Stochastic climate models, part I: Theory. Tellus 28: 473-485.
Hicke J.A. and Jenkins J.C. 2008. Mapping lodge pole pine stand structure susceptibility to mountain pine
beetle attack across the western United States. Forest Ecology and Management 225: 1536-1547.
Hidalgo H.G. 2004. Climate precursors of multidecadal drought variability in the western United States.
Water Resources Research 40: -.
26
Hidalgo H.G. and Dracup J.A. 2003. ENSO and PDO effects on hydroclimatic variations of the Upper
Colorado River basin. Journal of Hydrometeorology 4: 5-23.
Higgins R.W. and Shi W. 2000. Dominant factors responsible for interannual variability of the summer
monsoon in the southwestern United States. Journal of Climate 13: 759-776.
Higuera P.E., Brubaker L.B., Anderson P.M., Hu F.S. and Brown T.A. 2009. Vegetation mediated the
impacts of postglacial climate change on fire regimes in the south-central Brooks Range, Alaska.
Ecological Monographs 79: 201-219.
Hoerling M.P. and Eischeid J.K. 2007. Past peak water in the southwest. Southwest Hydrology 6:
January/February.
Hunter T., Tootle G. and Piechota T. 2006. Oceanic-atmospheric variability and western U.S. snowfall.
Geophys. Res. Lett. 33.
IPCC 2007. Climate Change 2007: Synthesis Report. Contribution of Working Groups I, II and III to the
Fourth Assessment
Report of the Intergovernmental Panel on Climate Change. A Report of the Intergovernmental Panel on
Climate Change. Intergovernmental Panel of Climate Change, Geneva, Switzerland. p. 104.
IPCC 2007. Climate Change 2007: The Physical Science Basis. Summary for Policymakers. IPCC
Intergovernmental Panel on Climate Change.
Jackson S.T., Betancourt J.L., Booth R.K. and Gray S.T. 2009. Ecology and the ratchet of events: Climate
variability, niche dimensions, and species distributions. Proceedings of the National Academy of
Sciences 106: 19685-19692.
Jackson S.T., Gray S.T. and Shuman B. 2009. Paleoecology and resource management in a dynamic
landscape: case studies from the Rocky Mountain headwaters. In Dietl G. P. and Flessa K. W. (eds.),
Conservation Paleobiology: Using the Past to Manage for the Future. Paleontological Society,
Lubbock, TX.
Jolly W.M., Nemani R. and Running S.W. 2005. A generalized, bioclimatic index to predict foliar
phenology in response to climate. Global Change Biology 11: 619-632.
Karl T.R., Groisman P.Y., Knight R.W. and Heim R.R. 1993. Recent variations of snow cover and snowfall
in North America and their relations to precipitation and temperature variations. Journal of Climate
6: 1327-1344.
Karl T.R., Melillo J.M. and Peterson T.C. 2009. Global Climate Change Impacts in the United States.
Cambridge University Press, New York, NY.
Kelly A.E. and Goulden M.L. 2008. Rapid shifts in plant distribution with recent climate change.
Proceedings of the National Academy of Sciences of the United States of America 105: 11823-11826.
Knowles N., Dettinger M.D. and Cayan D.R. 2006. Trends in snowfall versus rainfall in the Western
United States. Journal of Climate 19: 4545-4559.
Kunkel K.E., Easterling D.R., Redmond K. and Hubbard K. 2003. Temporal variations of extreme
precipitation events in the United States: 1895-2000. Geophys. Res. Lett. 30: 1900.
Kutzbach J., Gallimore R., Harrison S., Behling P., Selin R. and Laarif F. 1998. Climate and biome
simulations for the past 21,000 years. Quaternary Science Reviews 17: 473-506.
27
Lean J.L. and Rind D.H. 2009. How will Earth's surface temperature change in future decades? Geophys.
Res. Lett. 36.
LeGrande A.N., Schmidt G.A., Shindell D.T., Field C.V., Miller R.L., Koch D.M., Faluvegi G. and Hoffmann
G. 2006. Consistent simulations of multiple proxy responses to an abrupt climate change event.
Proceedings of the National Academy of Sciences of the United States of America 103: 837-842.
Lettenmaier D.P., Wood E.F. and Wallis J.R. 1994. Hydroclimatological trends in the continental United
States, 1948-88. Journal of Climate 7: 586-607.
Licciardi J.M., Clark P.U., Brook E.J., Elmore D. and Sharma P. 2004. Variable responses of western U.S.
glaciers during the last deglaciation. Geology 32: 81-84.
Littell J.S., McKenzie D., Peterson D.L. and Westerling A.L. 2009. Climate and wildfire area burned in
western U.S. ecoprovinces, 1916-2003. Ecological Applications 19: 1003-1021.
Littell J.S., Peterson D.L. and Tjoelker M. 2008. Douglas-fir growth in mountain ecosystems: water limits
tree growth from stand to region. Ecological Monographs 78: 349-368.
Lloyd A.H. 1997. Response of tree-line populations of foxtail pine (Pinus balfouriana) to climate variation
over the last 1000 years. Canadian Journal of Forest Research 27: 936-942.
Logan J.A., Regniere J. and Powell J.A. 2003. Assessing the impacts of global warming on forest pest
dynamics. Frontiers in Ecology and the Environment 1: 130-137.
Long C.J., Whitlock C., Bartlein P.J. and Millspaugh S.H. 1998. A 9000-year fire history from the Oregon
Coast Range, based on a high-resolution charcoal study. Canadian Journal of Forest Research 28:
774-787.
Luce C.H. and Holden Z.A. 2009. Declining annual streamflow distributions in the Pacific Northwest
United States, 1948–2006. Geophys. Res. Lett. 36.
MacDonald G.M., Bennett K.D., Jackson S.T., Parducci L., Smith F.A., Smol J.P. and Willis K.J. 2008.
Impacts of climate change on species, populations and communities: palaeobiogeographical insights
and frontiers. Progress in Physical Geography 32: 139-172.
Mann M.E., Bradley R.S. and Hughes M.K. 1999. Northern Hemisphere Temperatures During the Past
Millennium: Inferences, Uncertainties, and Limitations. Geophys. Res. Lett. 26.
Mann M.E. and Jones P.D. 2003. Global surface temperatures over the past two millennia. Geophys. Res.
Lett. 30.
Mann M.E., Zhang Z., Hughes M.K., Bradley R.S., Miller S.K., Rutherford S. and Ni F. 2008. Proxy-based
reconstructions of hemispheric and global surface temperature variations over the past two
millennia. Proceedings of the National Academy of Sciences 105: 13252-13257.
Mantua N. and Hare S. 2002. The Pacific Decadal Oscillation. Journal of Oceanography 58: 35-44.
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:
1069-1079.
McAfee S.A. and Russell J.L. 2008. Northern Annular Mode impact on spring climate in the western
United States. Geophys. Res. Lett. 35: L17701.
McCabe G. and Wolock D. Long-term variability in Northern Hemisphere snow cover and associations
with warmer winters. Climatic Change.
28
McCabe G. and Wolock D. 2009. Long-term variability in Northern Hemisphere snow cover and
associations with warmer winters. Climatic Change.
McCabe G.J., Betancourt J.L., Gray S.T., Palecki M.A. and Hidalgo H.G. 2008. Associations of multidecadal sea-surface temperature variability with US drought. Quaternary International 188: 31-40.
McCabe G.J., Betancourt J.L. and Hidalgo H.G. 2007. Associations of Decadal to Multidecadal Sea-Surface
Temperature Variability with Upper Colorado River Flow<sup>1</sup>. JAWRA Journal of the
American Water Resources Association 43: 183-192.
McCabe G.J. and Clark M.P. 2006. Shifting covariability of North American summer monsoon
precipitation with antecedent winter precipitation. International Journal of Climatology 26: 991-999.
McCabe G.J., Clark M.P. and Serreze M.C. 2001. Trends in Northern Hemisphere surface cyclone
frequency and intensity. Journal of Climate 14: 2763-2768.
McCabe G.J., Palecki M.A. and Betancourt J.L. 2004. Pacific and Atlantic Ocean influences on
multidecadal drought frequency in the United States. Proceedings of the National Academy of
Sciences of the United States of America 101: 4136-4141.
McCarthy J.J. 2009. Reflections On: Our Planet and Its Life, Origins, and Futures. Science 326: 1646-1655.
McDowell N., Pockman W.T., Allen C.D., Breshears D.D., Cobb N., Kolb T., Plaut J., Sperry J., West A.,
Williams D.G. and Yepez E.A. 2008. Mechanisms of plant survival and mortality during drought: why
do some plants survive while others succumb to drought? New Phytologist 178: 719-739.
Meehl G.A., Hu A. and Santer B.D. 2009. The Mid-1970s Climate Shift in the Pacific and the Relative
Roles of Forced versus Inherent Decadal Variability. Journal of Climate 22: 780-792.
Meko D.M., Woodhouse C.A., Baisan C.A., Knight T., Lukas J.J., Hughes M.K. and Salzer M.W. 2007.
Medieval drought in the upper Colorado River Basin. Geophysical Research Letters 34.
Menounos B., and Reasoner, M.A. 1997. Evidence for cirque glaciation in the Colorado Front Range
during the Younger Dryas chronozone. Quaternary Research 48: 38-47.
Millspaugh S.H., Whitlock C. and Bartlein P.J. 2000. Variations in fire frequency and climate over the past
17 000 yr in central Yellowstone National Park. Geology 28: 211-214.
Milly P.C.D., Dunne K.A. and Vecchia A.V. 2005. Global pattern of trends in streamflow and water
availability in a changing climate. Nature 438: 347-350.
Miriti M.N., Rodriguez-Buritica S., Wright S.J. and Howe H.F. 2007. Episodic death across species of
desert shrubs. Ecology 88: 32-36.
Moberg A., Sonechkin D.M., Holmgren K., Datenko N.M. and Karlen W. 2005. Highly variable Northern
Hemisphere temperatures reconstructed from low- and high-resolution proxy data Nature 433: 613
- 617.
Mock C.J. 1996. Climatic Controls and Spatial Variations of Precipitation in the Western United States.
Journal of Climate 9: 1111-1125.
Mock C.J. and Brunelle-Daines A.R. 1999. A modern analogue of western United States summer
palaeoclimate at 6000 years before present. The Holocene 9: 541-545.
Moore J.N., Harper J.T. and Greenwood M.C. 2007. Significance of trends toward earlier snowmelt
runoff, Columbia and Missouri Basin headwaters, western United States. Geophys. Res. Lett. 34.
29
Morgan P., Heyerdahl E.K. and Gibson C.E. 2008. Multi-season climate synchronized forest fires
throughout the 20th century, northern Rockies, USA. Ecology 89: 717-728.
Morin X., Augspurger C. and Chuine I. 2007. Process-based modeling of species' distributions: What
limits temperate tree species' range boundaries? Ecology 88: 2280-2291.
Morin X., Viner D. and Chuine I. 2008. Tree species range shifts at a continental scale: new predictive
insights from a process-based model. Journal of Ecology 96: 784-794.
Mote P.W. 2003. Trends in temperature and precipitation in the Pacific Northwest during the twentieth
century. Northwest Science 77: 271-282.
Mote P.W. 2006. Climate-Driven Variability and Trends in Mountain Snowpack in Western North
America&#42. Journal of Climate 19: 6209-6220.
Mote P.W., Hamlet A.F., Clark M.P. and Lettenmaier D.P. 2005. Declining mountain snowpack in western
north America. Bulletin of the American Meteorological Society 86: 39-+.
Mueller R.C., Scudder C.M., Porter M.E., Trotter R.T., Gehring C.A. and Whitham T.G. 2005. Differential
tree mortality in response to severe drought: evidence for long-term vegetation shifts. Journal of
Ecology 93: 1085-1093.
Patterson 2009. European Science Foundation BOREAS conference on humans in the Arctic, Rovaniemi,
Finland.
Pederson G.T., D.B. Fagre, S.T. Gray, and L.J. Graumlich 2004. Decadal-scale climate drivers for glacial
dynamics in Glacier National Park, Montana, USA. Geophysical Research Letters 31: L12203,
doi:10.1029/2004GL019770.
Pederson G.T., C. Whitlock, E. Watson, B. H. Luckman, and L. J. Graumlich 2007. Paleo-Perspectives on
Climate and Ecosystem Change. In Prato T. and Fagre D. (eds.), Sustaining Rocky Mountain
Landscapes: Science, Policy, and Management of the Crown of the Continent Ecosystem, pp. 151170. RFF Press, Washington D.C. USA.
Pestiaux P., Mersch I., Berger A. and Duplessy J.C. 1988. Paleoclimatic variability at frequencies ranging
from 1 cycle per 10 000 years to 1 cycle per 1000 years: Evidence for nonlinear behaviour of the
climate system. Climatic Change 12: 9-37.
Pohl K.A., Hadley K.S. and Arabas K.B. 2006. Decoupling tree-ring signatures of climate variation, fire,
and insect outbreaks in central Oregon. Tree-Ring Research 62: 37-50.
Ray A.J., Barsugli J.J. and Averyt K.B. 2008. Climate change in Colorado - A synthesis to support water
resources management and adaptation. University of Colorado at Boulder, Boulder, Colorado.
Reasoner M.A., Davis P.T. and Osborn G. 2001. Evaluation of proposed early-Holocene advances of
alpine glaciers in the North Cascade Range, Washington State, USA: constraints provided by
palaeoenvironmental reconstructions. The Holocene 11: 607-611.
Reasoner M.A. and Hickman M. 1989. Late Quaternary environmental change in the Lake O'Hara region,
Yoho National Park, British Columbia. Palaeogeography, Palaeoclimatology, Palaeoecology 72: 291316.
Reasoner M.A. and Huber U.M. 1999. Postglacial palaeoenvironments of the upper Bow Valley, Banff
National Park, Alberta, Canada. Quaternary Science Reviews 18: 475-492.
Reasoner M.A. and Jodry M.A. 2000. Rapid response of alpine timberline vegetation to the Younger
Dryas climate oscillation in the Colorado Rocky Mountains, USA. Geology 28: 51-54.
30
Reasoner M.A., Osborn G. and Rutter N.W. 1994. Age of the Crowfoot advance in the Canadian Rocky
Mountains: A glacial event coeval with the Younger Dryas oscillation. Geology 22: 439-442.
Regonda S.K., Rajagopalan B., Clark M. and Pitlick J. 2005. Seasonal Cycle Shifts in Hydroclimatology over
the Western United States. Journal of Climate 18: 372-384.
Rind D. 1999. Complexity and Climate. Science 284: 105-107.
Rochefort R.M., Little R.L., Woodward A. and Peterson D.L. 1994. Changes in sub-alpine tree distribution
in western North America: a review of climatic and other causal factors. The Holocene 4: 89-100.
Romme W.H., J. Clement, J.A. Hicke, D. Kulakoswki, L.H. MacDonald, T. Schoennagel and T.T. Veblen
2006. Recent forest insect outbreaks and fire risk in Colorado forests: A brief synthesis of relevant
research. Colorado State University, Fort Collins.
Salathe Jr E.P., Steed R., Mass C.F. and Zahn P.H. 2008. A High-Resolution Climate Model for the U.S.
Pacific Northwest: Mesoscale Feedbacks and Local Responses to Climate Change. Journal of Climate
21: 5708-5726.
Sato M., Hansen J.E., McCormick M.P. and Pollack J.B. 1993. Stratospheric Aerosol Optical Depths, 18501990. J. Geophys. Res. 98.
Schwartz M.D. and Reiter B.E. 2000. Changes in North American Spring. International Journal of
Climatology 20: 929-932.
Selkowitz D.J., Fagre D.B. and Reardon B.A. 2002. Interannual variations in snowpack in the Crown of the
Continent Ecosystem. Hydrological Processes 16: 3651-3665.
Shaw J.D., Steed B.E. and DeBlander L.T. 2005. Forest Inventory and Analysis (FIA) annual inventory
answers the question: What is happening to pinyon-juniper woodlands? Journal of Forestry 103:
280-285.
Shinker J.J., Bartlein P.J. and Shuman B. 2006. Synoptic and dynamic climate controls of North American
mid-continental aridity. Quaternary Science Reviews 25: 1401-1417.
Shinneman D.J. and Baker W.L. 2009. Historical fire and multidecadal drought as context for pinonjuniper woodland restoration in western Colorado. Ecological Applications 19: 1231-1245.
Shuman B., Bartlein P., Logar N., Newby P. and Webb Iii T. 2002. Parallel climate and vegetation
responses to the early Holocene collapse of the Laurentide Ice Sheet. Quaternary Science Reviews
21: 1793-1805.
Shuman B., Henderson A.K., Colman S.M., Stone J.R., Fritz S.C., Stevens L.R., Power M.J. and Whitlock C.
2009. Holocene lake-level trends in the Rocky Mountains, U.S.A. Quaternary Science Reviews 28:
1861-1879.
Solomon S., Rosenlof K.H., Portmann R.W., Daniel J.S., Davis S.M., Sanford T.J. and Plattner G.-K. 2010.
Contributions of Stratospheric Water Vapor to Decadal Changes in the Rate of Global Warming.
Science 327: 1219-1223.
Spears M., Brekke L., Harrison A. and Lyons J. 2009. Literature synthesis on climate change implications
for reclamation's water resources. U.S. Department of the Interior, Bureau of Reclamation, Denver,
Colorado.
Stahle D.W., Fye F.K., Cook E.R. and Griffin R.D. 2007. Tree-ring reconstructed megadroughts over North
America since AD 1300. Climatic Change 83: 133-149.
31
Stevens L.R., Stone J.R., Campbell J. and Fritz S.C. 2006. A 2200-yr record of hydrologic variability from
Foy Lake, Montana, USA, inferred from diatom and geochemical data. Quaternary Research 65: 264274.
Stewart I.T., Cayan D.R. and Dettinger M.D. 2004. Changes in Snowmelt Runoff Timing in Western North
America under a `Business as Usual' Climate Change Scenario. Climatic Change 62: 217-232.
Stewart I.T., Cayan D.R. and Dettinger M.D. 2005. Changes toward earlier streamflow timing across
western North America. Journal of Climate 18: 1136-1155.
Stuiver M., Grootes P.M. and Braziunas T.F. 1995. The GISP2 d18O climate record of the past 16,500
years and the role of the sun, ocean, and volcanoes. Quaternary Research 44: 341-354.
Swetnam T.W. and Betancourt J.L. 1998. Mesoscale Disturbance and Ecological Response to Decadal
Climatic Variability in the American Southwest. Journal of Climate 11: 3128.
Thompson R.S., Whitlock C., Bartlein P.J., Harrison S.P. and Spaulding W.G. 1993. Climate changes in the
western United States since 18,000 years B.P. In Wright H. E., Kutzbach J. E., Webb T. I., Ruddiman
W. F., Street-Perrott F. A. and Bartlein P. J. (eds.), Global climates since the last glacial maximum.
University of Minnesota Press, Minneapolis.
USGCRP 2001. Potential Consequences of Climate Change Variability and Change for the Pacific
Northwest. In U.S. Global Change Research Program N. A. S. T. (ed.), Climate Change Impacts on the
United States: The Potential Consequences of Climate Variability and Change, p. 612. Cambridge
University Press, Cambridge, UK.
van Geel B., Raspopov O.M., Renssen H., van der Plicht J., Dergachev V.A. and Meijer H.A.J. 1999. The
role of solar forcing upon climate change. Quaternary Science Reviews 18: 331-338.
van Mantgem P.J. and Stephenson N.L. 2007. Apparent climatically induced increase of tree mortality
rates in a temperate forest. Ecology Letters 10: 909-916.
Viau A.E. and Gajewski K. 2009. Reconstructing Millennial, Regional; Paleoclimates of Boreal Canada
during the Holocene. Journal of Climate 22: 316-330.
Viau A.E., Gajewski K., Sawada M.C. and Fines P. 2006. Millennial-scale temperature variations in North
America during the Holocene. J. Geophys. Res. 111.
Walsh M.K., Whitlock C. and Bartlein P.J. 2008. A 14,300-year-long record of fire-vegetation-climate
linkages at Battle Ground Lake, southwestern Washington. Quaternary Research 70: 251-264.
Watson E. and Luckman B.H. 2006. Long hydroclimate records from tree-rings in western Canada:
potential, problems and prospects. Canadian Water Resources Journal 31: 205-228.
Watson T.A., Barnett F.A., Gray S.T. and Tootle G.A. 2009. Reconstructed Streamflows for the
Headwaters of the Wind River, Wyoming, United States. JAWRA Journal of the American Water
Resources Association 45: 224-236.
Weisberg P.J. and Baker W.L. 1995. Spatial Variation in Tree Seedling and Krummholz Growth in the
Forest-Tundra Ecotone of Rocky Mountain National Park, Colorado, U.S.A. Arctic and Alpine
Research 27: 116-129.
Westerling A.L., Hidalgo H.G., Cayan D.R. and Swetnam T.W. 2006. Warming and earlier spring increase
western US forest wildfire activity. Science 313: 940-943.
32
Whitlock C. 1992. Vegetational and Climatic History of the Pacific Northwest during the Last 20,000
Years: Implications for Understanding Present-day Biodiversity. Northwest Environmental Journal 8:
5-28.
Whitlock C. 1993. Postglacial Vegetation and Climate of Grand Teton and Southern Yellowstone National
Parks. Ecological Monographs 63: 173-198.
Whitlock C. and Bartlein P.J. 1993. Spatial Variations of Holocene Climatic-Change in the Yellowstone
Region. Quaternary Research 39: 231-238.
Whitlock C., Dean W., Rosenbaum J., Stevens L., Fritz S., Bracht B. and Power M. 2008. A 2650-year-long
record of environmental change from northern Yellowstone National Park based on a comparison of
multiple proxy data. pp. 126-138.
Whitlock C., Reasoner M.A. and Key C.H. 2002. Paleoenvironmental history of the Rocky Mountain
region during the past 20,000 years. In Baron J. S. (ed.), Rocky Mountain futures: an ecological
perspective, pp. 41-57. Island Press, Washington DC.
Winter M.H. 1984. Altitudinal fluctuatons of upper treeline at two sites in the Lemhi Range, Idaho,
Unpublished Thesis, University of Kansas: Lawrence.
Wolter K. and Timlin M.S. 1993. Monitoring ENSO in COADS with a seasonally adjusted principal
component index. In Proc. of the 17th Climate Diagnostics Workshop, pp. 52-57. NOAA/NMC/CAC,
NSSL, Oklahoma Clim. Survey, CIMMS and the School of Meteor., Univ. of Oklahoma, Norman, OK.
Wolter K. and Timlin M.S. 1998. Measuring the strength of ENSO events - how does 1997/98 rank?
Weather 53: 315-324.
Woodhouse C.A., Gray S.T. and Meko D.M. 2006. Updated streamflow reconstructions for the Upper
Colorado River Basin. Water Resour. Res. 42.
Woodhouse C.A. and Overpeck J.T. 1998. 2000 years of drought variability in the central United States.
Bulletin of the American Meteorological Society 79: 2693-2714.
Yin J.H. 2005. A consistent poleward shift of the storm tracks in simulations of 21st century climate.
Geophysical Research Letters 32: L18701.
Zhang R. and Delworth T.L. 2007. Impact of the Atlantic Multidecadal Oscillation on North Pacific climate
variability. Geophysical Research Letters 34: -.
Zielinski G.A., Mayewski P.A., Meeker L.D., Whitlow S. and Twickler M.S. 1996. A 110,000-yr record of
explosive volcanism from the GISP2 (Greenland) ice core. Quaternary Research 45: 109-118.
33
Download