Download Bright et al _GCB Invited Review_v4

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Heaven and Earth (book) wikipedia , lookup

German Climate Action Plan 2050 wikipedia , lookup

ExxonMobil climate change controversy wikipedia , lookup

Soon and Baliunas controversy wikipedia , lookup

Mitigation of global warming in Australia wikipedia , lookup

Global warming controversy wikipedia , lookup

Climate resilience wikipedia , lookup

Climate change denial wikipedia , lookup

Albedo wikipedia , lookup

Climatic Research Unit documents wikipedia , lookup

Climate change adaptation wikipedia , lookup

Economics of global warming wikipedia , lookup

Global warming hiatus wikipedia , lookup

Citizens' Climate Lobby wikipedia , lookup

Climate engineering wikipedia , lookup

Fred Singer wikipedia , lookup

Effects of global warming on human health wikipedia , lookup

Climate change in Tuvalu wikipedia , lookup

Climate change in Saskatchewan wikipedia , lookup

Climate governance wikipedia , lookup

Climate change and agriculture wikipedia , lookup

Effects of global warming wikipedia , lookup

Global warming wikipedia , lookup

Media coverage of global warming wikipedia , lookup

Politics of global warming wikipedia , lookup

Climate change feedback wikipedia , lookup

Scientific opinion on climate change wikipedia , lookup

Climate change in the United States wikipedia , lookup

General circulation model wikipedia , lookup

Effects of global warming on humans wikipedia , lookup

Public opinion on global warming wikipedia , lookup

Instrumental temperature record wikipedia , lookup

Climate sensitivity wikipedia , lookup

Attribution of recent climate change wikipedia , lookup

Solar radiation management wikipedia , lookup

Climate change and poverty wikipedia , lookup

Surveys of scientists' views on climate change wikipedia , lookup

Climate change, industry and society wikipedia , lookup

IPCC Fourth Assessment Report wikipedia , lookup

Transcript
1
Quantifying surface albedo and other direct biogeophysical climate forcings of
2
forestry activities: A review
3
Running Head: Biogeophysical climate forcings and forestry
4
Ryan M. Bright*12, Kaiguang Zhao3, Robert B. Jackson4, Francesco Cherubini2
5
6
1
Norwegian Forest and Landscape Institute, Ås, Norway
7
2
Industrial Ecology Program, Energy and Process Engineering, Norwegian University of
8
Science and Technology, Trondheim, Norway
9
3
School of Environment and Natural Resources, Ohio Agricultural Research and
10
Development Center, The Ohio State University, Wooster, Ohio, USA
11
4
12
Energy, Stanford University, Palo Alto, California, USA
School of Earth Sciences, Woods Institute for the Environment and Precourt Institute for
13
14
*
15
[email protected]
16
Keywords:
17
management, climate metric
18
Paper type: Invited Review
Corresponding author contact:
Ryan M. Bright, phone:
+47 649 49003; email:
forest management, land use change, biophysical, climate forcing, land
19
1
20
Abstract
21
The regulation by vegetation of heat, momentum, and moisture exchanges between the land
22
surface and atmosphere is a major component in Earth’s climate system. By altering surface
23
biogeophysics, forestry and other land use activities often perturb these exchanges and affect
24
climate. Although long recognized scientifically as being important, surface biogeophysical
25
climate forcings continue to evade inclusion in climate policies enveloping forestry and other
26
land management projects.
27
atmospheric science and terrestrial ecology in light of three main objectives: i) to elucidate
28
the challenges associated with quantifying biogeophysical climate forcings connected to land
29
use and land management, with a particular focus on the forestry sector; ii) to identify and
30
describe current scientific approaches and/or metrics that facilitate their quantification; and
31
iii) to identify and recommend research priorities that can help overcome challenges in the
32
quantification and interpretation of biogeophysical climate impacts, serving to bridge the
33
knowledge gap between the climate modeling, forest ecology, and resource management
34
communities.
Here, we review the scientific literature in the field of
35
2
36
Table 1. Nomenclature. Variable “type” denotes how they are used in surface energy and
37
moisture budget equations.
Variable
notation
LAI
LAI e
SAI
hc
Unit
Definition
Type of Variable
Function of:
m 2 m 2
m 2 m 2
Leaf Area Index
Effective LAI
Structural
Structural
LAI
m 2 m 2
Stem Area Index
Canopy height
Structural
Structural
m
f cs
%
Fraction of canopy
intercepted snow
Structural &
Environmental
fg
%
unitless
Ground fraction
Snow-free ground
albedo
Structural
Physiological &
Environmental
Fraction of ground
covered snow
Snow albedo
Structural &
Environmental
Environmental
LAI; Ta ; P
Physiological &
Environmental
Moisture budget
term
Moisture budget
term
Moisture budget
term
Environmental
 leaf ; branch
g0
f gs
%
 sn
c0
unitless
P
mm
Snow-free canopy
albedo
Precipitation
RO
mm
Run-off
I
mm
Infiltration
Ta
˚C
Air temperature
Ts
˚C
Surface temperature

˚C(Wm-2)-1

RLW
Wm 2

Wm-2˚C -4
s
kPa ˚C -1
VPD
kPa
Temperature
sensitivity
Outgoing longwave
radiation
Stefan-Boltzmann
constant
Slope of saturation
vapor pressuretemperature curve
Vapor pressure deficit
ew
kPa
e* w
kPa
AP
RH

unitless
SAI; LAI; Ta ;
P
LAI
 leaf ;  soil ;
 leaf
P; Ta
H; L(E&T)
Surface energy
budget term
Environmental
 ; RN
Surface energy
budget term
Constant
Ts ;  s ; 
Environmental
Ta ; RH
Environmental
Ta ; RH
Environmental
Ta ; AP
Environmental
Ta ; AP
kPa
%
Partial pressure of
water vapor
Equilibrium vapor
pressure
Air pressure
Relative humidity
Environmental
Environmental
ew ; e* w
kgm-3
Air density
Environmental
3

kPa ˚C -1
s
unitless
Psychrometric
constant
Surface emissivity
s
unitless
Surface albedo
Cs
Jm ˚C
Cp
Jm-3˚C -1
ks
Wm-1˚C -1

RLW
Wm 2
RG
Wm 2
L( E  T )
Wm 2
-3
L
MJkg-1
E
mm
-1
Environmental
Surface energy
budget term
Surface energy
budget term
Surface energy
budget term
RH;


RSW
/ RSW
,clear
Ground heat storage
from conduction
Latent heat flux from
evaporation &
transpiration
Surface energy
budget term
Surface energy
budget term
LAI; Cs ; ks
Latent heat of
vaporization
Evaporation
Environmental
Ta ;
Environmental
ra ; C p ;  ;
Environmental
Environmental
Transpiration
Physiological;
Environmental
H
Wm 2
Sensible heat flux
Surface energy
budget term
z0
m

RSW
Wm-2
RN
Wm-2

unitless
 sn ;  c 0
Environmental
mm
ms 1
m
f g ;  g 0 ; f gs ;
Soil volumetric heat
capacity (thermal
inertia)
Air volumetric heat
capacity (thermal
inertia)
Soil thermal
conductivity
Incoming longwave
radiation
T
u
d
C p ; AP
Wind speed
Zero plane
displacement height
Momentum roughness
length
Incoming shortwave
radiation incident at
surface (insolation)
Net radiation
Bowen Ratio
VPD ; s;  ;
( RN  RG )
VPD ; s;  ;
( RN  RG )
ra ; rc ; C p ;  ;
VPD ; s;  ;
( RN  RG )
ra ; C p ;  ; Ts
- Ta
Environmental
Structural
LAI; hc
Structural
LAI; hc
Surface energy
budget term
Surface energy
budget term
Surface energy
4
ra ; rc ; C p ;  ;


;  s ; RLW
RSW
;  ;  s ; Ts
H ; L( E  T )
2 1
2
ga (ra )
ms
gc (rc )
m2 s 1 ( sm2 )
gs (rs)
m2 s 1 ( sm2 )
zr
m
Wm-2
TOA
SW
RF

TSW
( sm )
Unitless
VCC
VMC
FCC
FMC
Bulk aerodynamic
conductance
(resistance)
Bulk canopy or
surface conductance
(resistance)
Leaf stomatal
conductance
(resistance)
Rooting depth
Shortwave radiative
forcing at top-ofatmosphere (TOA)
Share of reflected SW
radiation at surface
arriving at TOA
Vegetation Cover
Change
Vegetation
Management Change
Forest Cover Change
Forest Management
Change
budget term
Structural &
Environmental
LAI; hc; u
Structural &
Physiological
LAI e ; gl
Physiological
Physiological
FCC/FMC climate
metric
 s
Environmental
38
39
1. Introduction
40
The terrestrial biosphere and Earth’s climate are closely entwined. Climate strongly
41
influences terrestrial productivity and biome distributions. In turn, the vegetation, soils, and
42
other components of the terrestrial biosphere influence climate through the amount of energy,
43
water, carbon, and other chemical species that they store and exchange with the atmosphere.
44
Human interventions directly alter vegetation cover and structure through the conversion of
45
one land use type to another (i.e., forest to cropland), or through a change in management for
46
an existing land use type (i.e., conversion of hardwood to softwood forest; addition of
47
irrigation or fertilization; extension of rotation length, etc.). In turn, such changes alter not
48
just the carbon balance of a system but perturb surface solar and thermal infrared radiation
49
budgets and atmospheric turbulence, leading to alterations in the fluxes of heat, water vapor,
5
50
momentum, CO2/other trace gases, and organic and inorganic aerosols between the land
51
surface and the atmosphere (e.g., Pielke Sr. et al., (1998, 2011). A deeper quantitative
52
understanding of how human intervention on land will affect climate regulation -- and over
53
which temporal and spatial scales -- is essential for successful climate change mitigation
54
(Feddema et al., 2005, Jackson et al., 2008, Mahmood et al., 2010).
55
Historically, extensive vegetation cover or management changes (henceforth referred to as
56
VCC or VMC) have entailed forest conversions to crops or grasslands for agriculture
57
(Goldewijk, 2001, Pongratz et al., 2008). Although the pace of global deforestation has
58
slowed in recent decades (FAO &
59
assessment modeling (IAM) community (van Vuuren et al., 2011) and the 5th Assessment
60
Reports of the IPCC suggest that forests will play an increasingly large role in climate change
61
mitigation and adaptation (Scholes & Settele, 2014, Smith & Bustamante, 2013) – whether
62
land areas are re-/afforested or whether existing forests are managed more intensively.
63
Relative to our understanding of forests’ role in the global carbon cycle, our understanding of
64
their non-CO2 influences on atmospheric chemistry and climate is in its infancy. Apart from
65
providing carbon sequestration services, forest ecosystems emit biogenic volatile organic
66
compounds (BVOCs) that can rapidly oxidize in the atmosphere, generating O3 and secondary
67
organic aerosols (SOAs) (Arneth et al., 2010). This biogeochemical mechanism impacts
68
climate both directly and indirectly (Scott et al., 2014, Spracklen et al., 2008) and its global
69
magnitude has only recently been examined (Unger, 2014).
70
However, more established scientifically are forests’ direct biogeophysical contributions to
71
the climate system: that is, their regulation of the exchanges of energy, water, and momentum
72
between the earth’s surface and lower atmosphere. Within the climate science and global
73
change research communities, forest cover and management changes and the corresponding
JRC, 2012, Meyfroidt et al., 2010), the integrated
6
74
changes in surface biophysics are increasingly recognized as important forcings of local,
75
regional, and global climate (Abiodun et al., 2008, Avissar & Werth, 2005, Betts et al., 2007,
76
Bright et al., 2014a, Bright et al., 2014b, Chapin III et al., 2012, Chen et al., 2012, Durieux et
77
al., 2003, Juang et al., 2007, Klingaman et al., 2008, Lawrence et al., 2012, Lee et al., 2011,
78
Mohr et al., 2003, Peng et al., 2014, Ray et al., 2006, Rotenberg & Yakir, 2010, Swann et al.,
79
2011, Wang et al., 2014, Zhang et al., 2014, Zhao & Jackson, 2014, Zheng et al., 2002).
80
Outside these communities, however, biogeophysical climate impacts from VCC and VMC
81
are rarely quantified or even acknowledged, with forest sector policies based strictly on
82
carbon cycle dimensions. Most climate assessments overlook forest biogeophysical effects
83
altogether due to the many complexities and challenges involved in quantifying them (Pielke
84
Sr. et al., 2002).
85
86
Our aim here is therefore to shed light on some of the challenges involved in measuring and
87
quantifying biogeophysical climate change effects connected to VCC and VMC, with a focus
88
on forestry (henceforth FCC and FMC). To that end, we review the scientific literature,
89
relying as much as possible on observation-based studies while recognizing modeling studies
90
that have made important contributions. We strive to limit our review to literature not
91
covered in previous reviews; for example, those enveloping general land use-atmosphere
92
climate dynamics (Foley et al., 2003, Pielke, 2001, Pielke Sr. et al., 2007, Pielke Sr. et al.,
93
1998), biogeophysical impacts of land use/cover changes on climate (Mahmood et al., 2013,
94
Pielke Sr. et al., 2011), and those with a special emphasis on forested ecosystem (Anderson et
95
al., 2010, Bonan, 2008, Jackson et al., 2008).
96
Our review is structured as follows: In Section 1, we present surface energy and moisture
97
budgets and describe how they are modulated by both biological and environmental controls
98
(i.e., biogeophysical). We illustrate how the surface energy balance equation can be
7
99
manipulated for purpose of estimating the relative contribution by radiative vs. non-radiative
100
processes that shape the response by local surface temperature following FCC/FMC. In
101
Section 2, we elaborate in greater detail the mechanisms responsible for the non-radiative
102
biogeophysical climate forcings linked to FCC/FMC, while in Section 3, we elaborate on
103
shortwave radiative forcings connected to FMC/FMC through changes in surface albedo.
104
Section 4 reviews recently proposed metrics/indicators for biogeophysical climate forcings,
105
and in Section 5 we present two case studies and apply the concepts presented in Sections 1-3,
106
demonstrating the importance of biological vs. environmental and of radiative vs. non-
107
radiative factors responsible for shaping the local climate services provided by forests. We
108
conclude in Section 6 with a discussion on the relevancy of FMC/FMC metrics and identify
109
critical research needs.
110
1. Mechanisms
111
Surface Energy and Moisture Budgets
112
FCC/FMC affects climate by altering surface moisture and energy budgets, which can be
113
written as:
114



RSW
(1   s )  RLW
 RLW
 RN  RG  H  L( E  T )
(1)
115
P  RO  I  E  T
(2)
116



where RN is the sum of the net short- ( RSW
) radiative
(1   s ) ) and longwave ( RLW
 RLW
117


fluxes; RSW
is insolation;  s is surface albedo; RLW
is the downwelling longwave radiation
118


flux; RLW
is the upwelling longwave radiation flux (equal to: RLW
(1   s )   s Ts4 , where  s
119
is the surface emissivity,  is the Stefan-Boltzmann constant, and Ts is the surface
120
radiometric temperature); E is evaporation (from soils and the physical environment); T is
8
121
transpiration (biologically-controlled); RG is heat stored in the ground and vegetation; H is
122
the turbulent sensible heat flux; L( E  T ) is the turbulent latent heat flux, with L as the latent
123
heat of vaporization; P is precipitation; RO is runoff; and I is infiltration.
124
Eq. (1) represents the surface energy budget, while Eq. (2) represents the surface moisture
125
budget. They are presented together because they are intimately linked (Pielke, 2001). For
126
instance, the latent heat flux L( E  T ) is directly related to the amount of moisture exchanged
127
from the surface to the atmosphere (E + T) which is governed largely by moisture availability
128
(P – RO - I) (Wang & Dickinson, 2012). For example, owed to their deeper rooting depths
129
and enhanced ability to access water stored in soils, L( E  T ) fluxes in temperate forests can
130
remain relatively high compared to grasslands during times of drought, when L( E  T ) fluxes
131
would otherwise be similar to forests under wet conditions (Stoy et al., 2006). Thus a change
132
to any term in Eq.’s (1) or (2) will affect the heat and moisture fluxes within the planetary
133
boundary layer and potentially act on atmospheric water vapor, cloud formation, precipitation,
134
and atmospheric circulation patterns. These boundary layer processes are dynamic, variable,
135
and difficult to predict, which generally limits the ability to predict the impact of land-use
136
change and landscape dynamics on climate patterns (Cotton & Pielke, 1995, Pielke, 2001).
137
Quantifying the full climate change effect of forest cover or management changes at
138
(inter)annual time scales is thus unrealistic for site-level observations. Such an approach
139
would require coupled surface-atmosphere models to account for boundary layer dynamics,
140
atmospheric albedo from clouds, and frontal and convective precipitation.
141
142
9
143
144
145
146
Figure 1. Differences in the annual surface energy and moisture budgets between a temperate
147
forest and an open grassland during dry (A) and wet (B) conditions. Horizontal fluxes of heat
148
and moisture are excluded, and “RG” includes heat stored by both the ground and vegetation.
149
Although annual RN is partitioned differently under arid and wet conditions, annual sensible
150
heat fluxes in temperate forests are typically higher relative to the grassland.
151
temperature; “ra” = aerodynamic resistance.
152
10
“Ta” = air
153
However, an analysis of changes in the major components of the surface energy and moisture
154
budgets due to FCC/FMC can contribute to an understanding of first order biogeophysical
155
effects. Changes in Eq. (1) will result in a change in the land surface temperature, since the
156
radiation that impinges on the surface must be balanced by the reflected and emitted radiation
157
and by energy lost or gained through sensible heat, latent heat, and conduction.
158
temperatures can also be affected by a changes in H and L(E+T), with a magnitude that
159
depends on the depth of the atmospheric boundary layer (Baldocchi & Ma, 2013). FCC/FMC
160
that results in a long-term perturbation in air temperature can potentially affect ecosystem
161
structure and functioning (Chapin III et al., 2012). For example, a switch in dominant tree
162
species (FMC) could create warmer surface conditions and enhance rates of soil respiration,
163
thus decreasing ecosystem productivity in the short-term (Rustad et al., 2001). In the longer-
164
term, the warmer surface could shift the distribution of understory plant species in ways that
165
enhance primary productivity (Chapin III et al., 1995, McGuire et al., 2006). It is therefore
166
important to understand feedbacks from FCC/FMC on local climate, and in particular,
167
attribute them quantitatively to biogeophysical factors. This necessitates a deeper
168
understanding of the relative roles of structural, physiological, and environmental controls on
169
surface energy and moisture budgets.
Air
170
171
Vegetation Structure
172
173
Structural parameters like leaf area index (LAI) and vegetation height play an important role
174
in determining resistances (or conductivities) to heat, moisture, and momentum transfer.
175
When a parcel of turbulent air meets a vegetated stand, wind speed is reduced, transferring
176
momentum from the atmosphere to the surface, creating turbulence that mixes the air and
177
transports heat and water from the surface into the lower atmosphere (Bonan, 2002, Monteith
11
178
& Unsworth, 2008, Oke, 2002). The transport of momentum, heat, and moisture is more
179
efficient with height above the surface and with densely vegetated canopies. LAI and forest
180
canopy heights thus play an essential role in determining roughness lengths and aerodynamic
181
resistances to heat, moisture, and momentum between the canopy and the atmospheric surface
182
layer. Relative to shorter-statured vegetation like crop- and grasslands, forested surfaces have
183
higher roughness lengths and lower aerodynamic resistances that facilitate more sensible heat
184
and water vapor dissipation away from the surface during the daytime (Hoffmann & Jackson,
185
2000).
186
187
Together with stem area index (SAI), LAI is also an important structural variable determining
188
the surface albedo and hence net radiation R N . SAI and LAI control the amount of solar
189
radiation incident at the ground level which is often covered in snow during winter in many
190
temperate and boreal regions. The albedo of snow is much higher than the albedo of foliage
191
or branches, thus SAI/LAI play a central role in regulating radiation budgets in regions with
192
long snow cover seasons due to the “masking” effects of forested canopies on the underlying
193
snow surface and hence the total albedo. Across North America between 45-60°N, the
194
zonally averaged white-sky albedo in January was at least twice as high for croplands and
195
grasslands (0.57 and 0.50, respectively) compared to locations with deciduous broadleaf or
196
evergreen needleleaf forests (0.26 and 0.20, respectively (Zhao & Jackson, 2014).
197
198
LAI is also an important variable determining bulk canopy conductance to heat and moisture
199
transfer, thus acting as controls on T as well as intercepted precipitation and canopy E. In
200
non-arid regions in summer, E & T is often highly correlated with LAI (see Wang &
201
Dickinson (2012) and cited studies therein).
202
12
203
Vegetation Physiology
204
205
Tree physiology plays an important role in governing T and I.
206
conductance (inverse of resistance) directly control rates of T at the individual leaf level,
207
while root structure and depth affect T through access to soil water. Root structure and depth
208
also affect I and the water storage capacity of soils (and thereby indirectly that which is
209
ultimately available for T and soil E). Forest management decisions that lead to a change in
210
tree species directly affects physiological controls of surface energy and moisture budgets.
211
For example, under non-drought conditions and given equal LAIs, a shift towards more
212
loblolly pine plantation area (Pinus taeda L.) at the expense of oak-hickory hardwood area
213
(Quercus – Carya) would increase regional T (and L(T)) in the SE USA due the loblolly
214
pine’s higher leaf stomatal conductance (Stoy et al., 2006).
For example, stomatal
215
216
Environmental Controls and Feedbacks with the Energy Balance
217
218
Local meteorological conditions can play an equally large role in determining surface energy
219

and moisture budgets. For instance, R N is determined by RSW
which is affected by cloud
220
cover and by surface albedo  s , which can be affected by temperature and P (snow has a
221
higher albedo than dry soil which has a higher albedo than wet soil). The partitioning of the
222
turbulent heat fluxes ( R N  RG ) into H and L( E  T ) is also partially controlled by differences
223
between the air and surface temperatures and by differences in saturated vs. actual vapor
224
pressures (i.e., the vapor pressure deficit), with vapor pressure having an exponential
225
relationship with air temperature (Bonan, 2002, Monteith & Unsworth, 2008). Wind speed
226
also determines the aerodynamic resistance of the surface-atmospheric boundary layer, with
227
resistance decreasing as wind speeds increase.
13
228
229
Table 2 describes the individual variables in Eq. (1) in terms of their important controls. Refer
230
to Table 1 for variable descriptions.
231
232
Table 2. Surface energy budget variables (Eq. (1)) and their controls.
Variable Unit
notation
Definition
s
unitless

RSW
Wm 2
Surface albedo,

 1
RSW
RSW
Insolation at
surface

LW
2
R
Wm
RG
Wm 2
H
Wm 2
L( E  T )
Wm 2
Longwave
irradiance at
surface
Heat storage flux
from conduction
Turbulent sensible
heat flux
Turbulent latent
heat flux
s
unitless
Surface emissivity

RLW
Wm 2
Longwave
emittance at
surface
Controls
Environmental
 sn ; P; Ta
Structural
LAI; SAI
Physiological
 leaf ;  leaf
 atm ;  atm ;
latitude
 atm ;  atm ;  atm ;
Ta ; RH
ks ; Cs
LAI; SAI
C p ;  ; u;
hc ; LAI (in
Ts  Ta
VPD, C p ;  ;
z0 , in ra )
LAI e (in rc )
rs , zr
Ta ; u ; s; 
Ts ;  s
233
234
Energy Budget Decomposition Analysis
235
236
A useful approach allowing for an assessment of the contribution by radiative, aerodynamic,
237
and physiological (i.e., the biogeophysical) factors to local climate forcings connected to
238
FCC/FMC is to rearrange the various terms of the surface energy balance equation (Eq. (1)).
239
The local climate forcing may be measured in terms of a surface (radiative) temperature
240
change, ∆Ts, or by an air temperature change, ∆Ta. Some have formalized approaches to
241
estimate a discrete change in Ts by rearranging terms of the surface energy balance and taking
14
242
first order derivatives to allow for an assessment of both radiative and non-radiative terms in
243
isolation (Juang et al., 2007, Lee et al., 2011, Luyssaert et al., 2014). Based on the findings
244
of Juang et al. (2007) that heat storage ( RG ) and emissivity (  s ) terms of Eq. (1) are
245
negligible on annual time scales, Lee et al. (2011) formulated an alternative model that
246
recognizes that a radiative forcing at the surface must be compensated by atmospheric
247
feedbacks governing the energy re-distribution at the surface, which is brought about by the
248
concomitant changes to important aerodynamic and physiological attributes:
249
250
Ts 
0
1 f
RFSFC

s
0
(1  f )2
RN f
(3)
251
252
where RFSFC
is the radiative forcing at the surface1 due to changes in albedo, and f is an
s
253
energy “redistribution efficiency” parameter that is determined by the intrinsic aerodynamic
254
and physiological attributes of the vegetation:
255
f 
256
 C p  L( E  T ) 
1 

4ra Ts3 
H

(4)
257
258
where ra is the bulk aerodynamic resistance (defined in Section 2, Eq. (8)),  is air density
259
(kg m-3), and C p is the thermal inertia of air (J m-3 ˚C-1).
260
temperature sensitivity resulting from the longwave radiation feedback ( 1 / (4 s Ts3 ) ; a weak
261
function of Ts with units in ˚C(W m-2)-1) (Lee et al., 2011, National Research Council, 2003).
262
The 0 term essentially sets an upper limit on surface radiative forcing contribution to Ts ; in
1
In Eq. (3), 0 is the local
We use the term ”radiative forcing” to denote an energy balance perturbation that is ”radiative” in nature as
opposed to non-radiative, regardless of atmospheric level. Note the IPCC definition of “radiative forcing” refers
strictly to planetary energy balance perturbations (i.e., at TOA).
15
263
other words, the first right-hand term in Eq. (3) would be the change in surface skin
264
temperature if radiative processes were the only energy transfer mechanisms at play.
265
However, the actual surface temperature response following a radiative forcing depends on
266
internal energy re-distribution through convection and evapotranspiration, which in turn
267
depends on the structural and physiological properties of the vegetation and on ambient
268
environmental conditions (i.e., air temperature, humidity, wind speed). These non-radiative
269
factors are largely responsible for f and can be equally important in determining the overall
270
local Ts connected to FCC/FMC.
271
272
2. Non-radiative Forcing
273
Convective heat transfer of sensible heat H is directly proportional to the difference in air
274
temperature at some reference height and at the surface level, and is inversely related to an
275
overall resistance:
276
H   C p (Ta  Ts ) / ra
277
where  is the air density (kg m-3), C p is the heat capacity of air (J kg-1 ˚C-1), and ra is the
278
overall aerodynamic resistance to heat transfer from the surface to the atmospheric boundary
279
layer.
280
Transfer of latent heat ( L( E  T ) ) is directly proportional to the difference in the vapor
281
pressure of air at some reference height and at the surface level, and is also inversely related
282
to an overall resistance. Latent heat exchanges can be parameterized in a variety of ways,
283
although the Penman-Monteith equation (Monteith, 1965) is widely considered an accurate
284
expression to estimate E+T (Allen et al., 1989, Allen et al., 1998), developed to use surface
285
radiation, temperature, and humidity data (Wang & Dickinson, 2012). For forests, the
(5)
16
286
surface can be regarded as a “big leaf” (Deardorf 1978) in which a separate resistance to
287
water vapor transfer from the canopy is introduced:
288
L( E  T ) 
289
where s is the slop of the saturation vapor pressure-temperature curve (kPa ˚C-1) (Murry,
290
1967, Tetens, 1930), VPD is the vapor pressure deficit (kPa; a function of Ta and relative
291
humidity),  is the psychrometric constant (kPa ˚C-1), and rc is the canopy resistance to water
292
vapor transfer. Eq.’s (5) and (6) demonstrate that both the canopy and aerodynamic resistance
293
terms are critical parameters controlling turbulent heat (and moisture) exchanges with the
294
atmosphere, with the former being aerodynamically controlled and the latter being both
295
aerodynamically and physiologically controlled.
296
Canopy resistance
297
At the scale of an individual leaf, stomatal control of transpiration is known as the leaf
298
stomatal resistance rl . At the scale of a canopy of leaves, canopy resistance rc is used to
299
describe the aggregate resistance. It is often calculated by scaling up the leaf stomatal
300
resistance ( rs ) of the leaves acting in parallel while treating the canopy as one “big leaf”:
301
rc 
302
where LAI e is the effective LAI which is empirically equal to the actual LAI for LAI ≤ 2,
303
LAI/2 for LAI ≥ 4, and 2 for others (Ding et al., 2014). “Dual-leaf” canopy resistance models
304
that take into account the share of sunlit vs. shaded leaves in the canopy often give more
305
accurate results than the “big leaf” model but require additional computations of the sunlit
s( RN  RG )   C pVPD / ra
(6)
s  (1  rc / ra )
rl
LAI e
(7)
17
306
fraction and separate values of mean leaf stomatal resistances for shaded and unshaded leaves
307
(Ding et al., 2014, Irmak et al., 2008, Zhang et al., 2011).
308
Table 3. Typical minimum canopy ( rc ) and leaf stomatal resistances ( rs ) for various
309
vegetation types (adapted from Kelliher et al., (1995)). Effective LAI (“ LAI e ”) is deduced
310
with Eq. (7).
Vegetation Type
rs (s m-1)
Temperate grassland
125
Coniferous forest
175
Temperature deciduous 215
forest
Tropical rainforest
165
Cereal crops
90
Broadleaved herbaceous 80
crops
rc (s m-1)
60
50
50
LAI e
2.1
3.5
4.3
80
30
35
2.1
3
2.3
311
312
Canopy resistance is sometimes referred to as surface resistance when it describes the
313
aggregate resistance of all transpiration and evaporation processes occurring on the ground
314
and in the canopy including the evaporation of water intercepted by the canopy. The most
315
advanced models for predicting E & T typically estimate separate surface resistances for T, E
316
occurring at the ground level, and E occurring in the vegetation canopy (Wang & Dickinson,
317
2012).
18
318
319
Figure 2. Differences in the partitioning of daily E & T between adjacent young and mature
320
spruce-dominant stands (Picea abies (L.) H. Karst.) in eastern Norway during 2006. Adapted
321
from Bright et al. (2014a).
322
Figure 2 illustrates the contribution to the daily latent heat flux from T and E occurring at the
323
surface and in the canopy (from intercepted moisture) at neighboring sites in eastern Norway
324
estimated with the Penman-Monteith scheme of Mu et al. (2011). L( E  T ) between the
325
two sites (sharing identical ambient environmental conditions) stems mostly from the
326
additional contribution by T and canopy E at the mature forest site. The region is not
327
moisture-limited, thus the contribution from soil E is large at both sites and dominates total
328
E+T throughout most of the year (Fig. 2, green).
329
Aerodynamic resistance
19
330
Sources of heat and water vapor will generally be found lower in the canopy than the apparent
331
sink of momentum, thus the overall aerodynamic resistances to heat and mass transfer may
332
therefore be described in terms of ram -- the aerodynamic resistance to momentum transfer,
333
and rb -- an additional resistance term assumed to be identical for heat and water vapor
334
(Monteith & Unsworth, 2008):
335
ra  ram  rb   log(( z  d ) / z0 )  / k 2u ( z )  2(ku* ) 1
336
where k is von Karman’s constant (0.41), z is the reference height (m), d is the zero plane
337
displacement height (m), u(z) the wind speed at reference height (m s-1), z0 the roughness
338
length of momentum (m), u* is the friction velocity (around 0.05-0.1u) (m s-1), Sc is the
339
Schmidt, and Pr is the Prandtl number (the ratio of the two being the Lewis number – or the
340
ratio of thermal to mass diffusivity). The second right-hand expression in Eq. (8) is the
341
additional resistance term ( rb ) for rough or fibrous vegetation surfaces and is based on the
342
empirical works of Thom (1972) and Wesely & Hicks (1977).
343
Key terms in Eq. (8) influencing the value of ra are momentum roughness length z0 and the
344
zero plane displacement height d – both of which are often parameterized as a function of
345
vegetation structure (Pereira et al., 1999, Perrier, 1982):
346
2


d  hc 1 
1  e  LAI /2  

 LAI

(9)
347
z0  hc e LAI / 2 (1  e  LAI / 2 )
(10)
348
where hc is canopy height (m). Eq.’s (9) & (10) are valid for LAI ≥ 0.5 (Colaizzi et al., 2004).
349
Other empirical formulations can involve additional forest structural attributes like stand
2
20
Sc
Pr
0.67
(8)
350
density (number of trees per hectare) (Nakai et al., 2008), but many modelers simply scale d
351
with canopy height hc (2/3hc for forests and 1/8hc for uniform crops (Allen et al., 1998)).
352
What should be apparent when looking at Eq.’s (5) - (10) is that both canopy (or surface) and
353
aerodynamic resistances are key terms controlling L(E+T) and H, and these terms are in turn
354
both largely determined by vegetation structure (i.e., LAI and hc on d, z0 , and rc ) and
355
physiology (i.e., rl ). While it is often assumed that decreases in L( E  T ) result in an
356
increase in surface temperature, this is not always necessarily the case. Ts responds to
357
changes in both aerodynamic roughness and in the Bowen ratio – or the ratio of sensible to
358
latent heat -- which can be either positive or negative in response to FCC (Lee et al., 2011).
359
360
Air Temperature
361
362
Thus far we have dealt with the principles by which the energy absorbed when vegetation
363
exposed to radiative fluxes is transferred to the atmosphere by convection and evaporation of
364
water (latent heat transfer) and the resulting effects on the surface radiative temperature, Ts .
365
These energy transfers take place through a boundary layer with properties dependent on the
366
viscosity of air and the transport of momentum from moving air to the vegetation surface
367
(Monteith & Unsworth, 2008). As such, the amount of warming in the air ( Ta ) depends on
368
the extent of turbulent mixing in the atmosphere, which is described by the depth of the
369
atmospheric boundary layer (Oke, 2002).
370
properties, forests are more efficient at dissipating sensible heat away from the surface and
371
into the boundary layer relative to open areas with shorter vegetation, particularly during the
372
daytime (Hoffmann & Jackson, 2000, Lee et al., 2011, Zhang et al., 2014). At nighttime,
373
however, their higher roughness properties can also serve to bring more heat from the stably
Due to their larger aerodynamic roughness
21
374
stratified nocturnal boundary layer down towards the surface layer relative to open areas. The
375
net result for forests in many extra-tropical regions is a net warming effect over the diel (24-
376
hr.) cycle (see, for example, Fig. 2 in Lee et al. (2011) and Fig. 4 in Zhang et al. (2014)).
377
378
Due to atmospheric turbulence, changes in the radiometric temperature of the surface ( Ts )
379
following FCC/FMC may not always provide the best indication of the actual air temperature
380
change, Ta . When comparing seasonal Ts and Ta observations (24-hr.) between a mature
381
coniferous and a recently clear-cut stand in boreal Canada (Figure 3), for example, we find
382
that the sign of the Ta does not equal that of Ts for all seasons except winter (DJF, Figure
383
3). This is more noticeable in summer (JJA), where Ta is negative and Ts is positive at
384
the Clear-cut site relative to the Old Jack Pine site.
385
22
386
387
388
Figure 3. Observed 2004-2010 mean seasonal differences in diel (24-hr.) surface ( Ts ) and air
389
( Ta ) temperatures (shown with 1 SD of inter-annual variations) in a Canadian boreal forest
390
cluster of different species and age compositions. The “Clear-cut” site was established in
391

2002. The sites share approximately identical global radiation ( RSW
) and other background
392
meteorological forcings (i.e, P). “DJF” = Dec.-Jan.-Feb.; “MAM” = Mar.-Apr.-May; “JJA” =
393
Jun.-Jul.-Aug.; “SON” = Sep.-Oct.-Nov. Ta data are from Fluxnet Canada (Barr & Black,
394
2013a, Barr & Black, 2013b, Barr et al., 2013), while Ts data are from MODIS (ORNL
395
DAAC, 2014).
396
23
397
Figure 3 demonstrates that considering either Ts or Ta in isolation can sometimes lead to
398
different conclusions about the relative local climate benefits of different vegetation cover
399
types. Although simple approximations can be applied to estimate Ta directly from H by
400
assuming a fixed mixing layer height in the convective boundary layer (West et al., 2011),
401
more accurate predictions require dynamic mixing layer models that take into account its
402
growth over the diurnal cycle (McNaughton & Spriggs, 1986). Without the application of a
403
dynamic boundary layer model, for example, Baldocchi and Ma (2013) could not have
404
explained the observed differences in air temperatures ( Ta ) between an oak savannah and
405
adjacent grassland site despite having equal sensible heat fluxes ( H = 0) during some
406
periods of the year. In other words, the convective boundary layer can serve as a buffer to the
407
turbulent sensible heat being dissipated away from the surface, complicating predictions of
408
Ta .
409
410
3. Radiative Forcing
411
Surface Albedo
412
Surface albedo is one of the most important biogeophysical mechanisms acting on radiation
413
budgets at both surface and top-of-atmosphere levels hence it affects both local and global
414
climate (Cess, 1978, Otterman, 1977). Forests and taller vegetation are often darker than
415
those with sparse or shorter vegetation (Henderson-Sellers & Wilson, 1983), particularly
416
when the underlying surface is covered in snow or light-colored soil. In temperate and boreal
417
regions, the interactions between forested vegetation and snow significantly complicate the
418
relationship between FCC/FMC and surface albedo changes (Boisier et al., 2012, Bright et
419
al., 2014b, de Noblet-Ducoudré et al., 2012, Pitman et al., 2009).
24
420
Parameterizations of surface albedo  s for forested areas are diverse with respect to treatment
421
of ground masking by vegetation, which can be classified according to three prevailing
422
methods introduced in Qu & Hall (2007) (and later described in Essery (2013)). Briefly, the
423
first method estimates radiative transfer between the vegetation canopy and the ground
424
surface; the second method combines the vegetation and ground albedos with weights
425
determined by vegetation cover; and the third method combines the snow-free and snow
426
albedo with weights determined by snow cover.
427
necessarily superior to another (Bright et al., 2014b, Essery, 2013), we find it helpful to
428
describe the albedo dynamics using the second approach, where the surface albedo is a
429
weighted share of the albedo of the ground and of the forest canopy, including the share of
430
ground and canopy covered in snow (Roesch & Roeckner, 2006, Verseghy et al., 1993):
431
 s  f g (1  f gs ) g 0  f gs sn   (1  f g ) (1  f cs ) c 0  f cs sn 
432
where f g is the fraction of exposed ground (sometimes referred to as the “canopy gap
433
fraction”), 1- f g the fraction of the exposed canopy (sometimes referred to as the “canopy
434
radiative fraction”), f cs the fraction of canopy covered with snow, f gs the fraction of ground
435
covered in snow,  sn the albedo of snow,  g 0 the snow-free ground albedo, and  c 0 the
436
snow-free canopy albedo. f g (or 1- f g ) is determined by vegetation structure and is often a
437
function of LAI and SAI.
438
determined by both vegetation structure and local meteorology (Bartlett et al., 2006, Essery et
439
al., 2009, Hedstrom & Pomeroy, 1998, Niu & Yang, 2004).  c 0 is an intrinsic property of
440
the vegetation largely determined by leaf level albedo and canopy structure (Hollinger et al.,
441
2010, Ollinger et al., 2008, Sellers, 1985).  g 0 is largely determined by soil geology but
442
varies with environmental factors influencing soil moisture (Idso et al., 1975), whereas  sn is
Although no one particular approach is
(11)
f cs is often analogous to canopy intercepted snow which is
25
443
purely controlled by environmental factors like precipitation rates, wind, temperature, and
444
other factors influencing snow grain size and impurities (like soot deposition) (Pirazzini,
445
2009, Wiscombe & Warren, 1980).
446
FCC/FMC primarily affects  s through alterations in vegetation structure and to some extent
447
physiology. In addition to the surface energy budget, changes to  s (henceforth  s ) can
448
directly alter the top-of-the-atmosphere (TOA) radiation balance and hence global mean
449
temperature (Arora & Montenegro, 2011, Bala et al., 2007, Davin & de Noblet-Ducoudré,
450
2010). Shortwave  s RFs at the TOA can be approximated with information on local
451
insolation and atmospheric conditions:
452


RFTOA
 s  RSW  sTSW
(12)
453


Where RSW
is the local insolation,  s is the local albedo change, and TSW
is the fraction of
454

reflected RSW
arriving back at TOA. During multiple reflection and on the final trajectory of
455
the reflected shortwave radiation towards TOA, there are opportunities for additional
456
atmospheric absorption which reduces the impact of  s upon the TOA flux change relative
457
to its impact at the surface (Winton, 2005). This is accounted for in Eq. (12) through the use
458

of the upward atmospheric transmittance parameter TSW
.
459
measure/obtain, particularly if it is estimated using vertical profiles of optical properties,
460
leading some to apply a value corresponding to the global annual mean (Bright et al., 2014a,
461
Caiazzo et al., 2014, Cherubini et al., 2012, Muñoz et al., 2010). Bright and Kvalevåg (2013)
462
showed that this may be a reasonable assumption.
463
As an alternative to Eq. (12) one could apply “radiative kernels” (Shell et al., 2008, Soden et
464
al., 2008) that relate  s directly to TOA RFs (for example, as in Flanner et al., (2011) and
26

Locally, TSW
is difficult to
465
Ghimire et al., (2014)). A radiative kernel describes the change in TOA fluxes for a standard
466
change in a surface property like  s and depends on the radiative properties and base state of
467
the climate model from which they are derived (Shell et al., 2008, Soden et al., 2008). Global
468
annual mean radiative kernels for  s have been estimated to range between 1.29 and 1.61
469
(W m-2 (0.01  s )-1) depending on the radiative transfer scheme and climate model (Shell et
470
al., 2008, Soden et al., 2008).
471
Global Mean Temperature Change
472
The extent to which RFTOA
 s impacts global mean surface temperatures can be estimated with
473
application of a global climate sensitivity parameter  that may describe either the transient,
474
effective, or equilibrium response by global mean temperature to a unit RFTOA
(in
s
475
C (Wm2 )1 ). The equilibrium climate sensitivity refers to the net change in the annual mean
476
global near-surface Ta when the climate system (or model) attains a new equilibrium with the
477
forcing change (typically corresponding to a doubling of atmospheric CO2-eq. concentration)
478
(Planton, 2013).
479
The effective climate sensitivity is a related measure that circumvents the requirement of
480
equilibrium. It is evaluated from model output for evolving non-equilibrium conditions and is
481
a measure of the strengths of the climate feedbacks at a particular time which may vary with
482
forcing history and climate state.
483
The transient climate response defined by the IPCC is the change in the global surface
484
temperature, averaged over a 20-year period, centered at the time of atmospheric carbon
485
dioxide doubling (year 70 in a 1% yr–1 compound carbon dioxide increase experiment with a
486
global coupled climate model) (Planton, 2013). However, the term is often loosely applied in
487
the literature to denote the transient response to a forcing regardless of the type of forcing
27
488
(CO2) or perturbation rate (1%/yr). It is essentially a measure of the strength and speed of the
489
response by Ta to a radiative forcing, and is typically lower than the equilibrium sensitivity
490
due to the long timescales of heat uptake by oceans.
491
The global climate sensitivity to RFTOA
 s is highly uncertain and is discussed in greater detail in
492
Section 6.
493
4. Climate Metrics for Land Use-Climate Forcings
494
Climate Regulation Index
495
West et al. (2011) developed “climate regulation indices” that combined two dominant
496
process influencing regional variations in climate: i) the biogeophysical regulation of heat
497
and moisture fluxes from local land surface processes; ii) the advection (transport) of heat and
498
moisture from large-scale atmospheric circulation. With such metrics, the local surface
499
energy and moisture balance impacts of V/FCC are scaled relative to the influence of
500
advection, thus providing an indication of the relative importance of the intrinsic
501
biogeophysical properties of the vegetated land surface: as advection increases, the relative
502
importance of the intrinsic biophysical mechanisms on local climate ( Ta in ˚C and moisture in
503
the atmospheric water column in mm) decreases.
504
To the best of our knowledge, these so-called “heat and moisture regulation indices” represent
505
the only spatially-explicit (0.5˚ x 0.5˚) V/FCC-climate metrics with global coverage
506
developed without the use of coupled (land surface-atmosphere) models and that characterize
507
the V/FCC biogeophysical impact the furthest along the local cause-effect chain from changes
508
in
509
V / FCC  RN  Ts  Ta ). Further work is required, however, to improve the temporal
510
resolution of the prescribed state variables used in the climate forcing characterizations, which
net
radiation
to
changes
in
near-surface
28
air
temperatures
(i.e.,
511
the authors themselves acknowledge. For example, West et al. (2011) prescribed a mixing
512
layer height of the near-surface atmosphere that is fixed annually, although in reality it varies
513
with season in response to V/FCC (Baldocchi & Ma, 2013, Oke, 2002). Further, the metrics
514
are developed using the theoretical potential vegetation cover for each grid cell compared to
515
bare ground/no vegetation; efforts would be required to make similar metrics more amenable
516
to forest management contexts.
517
Climate Regulation Value
518
Anderson-Teixeira et al. (2012) quantified both biogeochemical and biogeophysical
519
ecosystem climate services of 18 eco-regions across the Americas and combined them into a
520
single indicator referred to as the “Climate Regulation Value” (CRV) that indexes the relative
521
importance of biogeophysical to biogeochemical eco-physiological processes on land. CRV
522
is essentially the time-integrated net change in global LW GHG forcing at TOA less the local
523
surface energy balance change ( RFGHG  RN  L( E  T ) ; in Watts per global m2) relative to
524
a pulse emission of CO2 occurring in the year of LULCC.
525
The metric essentially combines the local direct biogeophysical climate effect (normalized to
526
the area of the earth) with the global biogeochemical effect which is normalized to CO2 as the
527
common currency (to obtain units in “CO2-eq.”). “Climate effect” here is simply the energy
528
gained or removed from climate system at multiple levels of the atmosphere (surface and
529
TOA). The metric does not account for the change in surface skin ( Ts ) or near-surface air
530
temperature ( Ta ) due to, for example, changes in surface aerodynamic properties (governed
531
by roughness) influencing convective heat transfer (Lee et al., 2011, Mahmood et al., 2013,
532
Pielke Sr. et al., 2011) -- which is of greater relevance to humans and to the functioning of
533
local ecosystems (Betts, 2007, Pielke Sr. et al., 2002). Additionally, local non-radiative (i.e.,
534
from ∆H) and global radiative (i.e., from CO2 RF) effects are summed, making meaningful
29
535
interpretation difficult. Like those derived by West et al. (2011), the baseline land cover in
536
Anderson-Teixeira et al. (2012) used to derive the CRV metric is bare soil depleted of organic
537
matter, which is not a realistic V/FCC scenario.
538
Entropy Production
539
Entropy production is a thermodynamic measure of the strength of dissipative processes
540
which perform physical work. Kleidon (2006), and later Stoy et al. (2014), argued that
541
entropy production is meaningful because it characterizes climate sensitivity (local Ta ) not
542
by how much warming or cooling occurs but by how much the suitability to perform physical
543
work differs by physical dissipative processes such as the turbulent exchange of the surface
544
and the atmosphere.
545
changes reduced the ability of natural ecosystems to perform work and that entropy
546
production rates may be a more meaningful measure of climate sensitivity than the global
547
mean Ta . Entropy reduction scales positively with increasing land use intensity (i.e., human
548
appropriation of NPP). In northern latitudes, this reduction stems from reductions in RN due
549
to reductions in SWNet (increased in albedo); in southern latitudes, entropy reductions stem
550
from a reduction in L(E+T) fluxes.
551
5. Case Studies
552
Environmental vs. Biological Controls
553
To what extent can the environmental controls (i.e., global radiation, precipitation, wind, etc.)
554
relative to the intrinsic biological properties of the vegetation itself (i.e., stomatal
555
conductance, LAI, vegetation height, etc.) influence the vegetation feedbacks on climate? For
556
any two forests having similar structural and physiological properties (i.e., LAI, hc, gs, zr),
557
 s will mostly be driven by differences in air temperature and precipitation as it affect snow
Kleidon showed that climate impacts from large-scale land cover
30
558

cover (when f g and  g 0 are assumed negligible), by differences in RSW
due to differences
559
in latitude and atmospheric conditions (i.e., cloud cover, aerosol optical depth), and by
560
differences in 
561
Understanding the role of environmental controls is relevant for regionally-optimized climate
562
motivated forest management strategies and policies. Peng et al. (2014) showed that the
563
benefits of afforestation in China (in terms of local Ts relative to grassland or cropland) are
564
likely to be enhanced in wetter regions (P >1200 mm yr-1) due to higher L(E+T) and can
565
even be counter-productive in dryer regions. For instance, the annual mean Ts between
566
plantation forests and open lands was found to be ~ -2ºC for regions experiencing P >1600
567
mm yr-1 while ~ +2.5ºC for regions experiencing P = 400-600 mm yr-1.
568
Another example is in Norway, where coastal regions in the west are being considered for
569
large-scale afforestation of fallow pasture and cropland with spruce plantations. The climate
570
of the region is characterized as having much higher P and relatively milder annual mean air
571
temperatures (P = >3000 vs. 500-100 mm yr-1 and T = 6.3 C vs. 2.7 C, respectively)
572
(Norwegian Meteorological Institute, 2013).
573
contribute to a lower annual mean  s and higher L(E+T) relative to forests of similar
574
structure (basal area, LAI, dominant species) located in the cooler and dryer eastern regions.
575
Thus despite experiencing slightly higher mean annual RN loads, afforestation in the coaster
576
regions would warm Ts and cool Ta locally relative to similar forests in eastern regions of
577
Norway, as illustrated in Table 4.
578
Table 4. Site pair comparison demonstrating the relative importance of environmental versus
579
structural and physiological (“Biological”) controls on annual mean air and surface
580
temperatures. Flux data for the USA are for 2005 and are adapted from Katul & Oren (2011a,
H
due to differences in humidity, precipitation, and wind speed.
L( E  T )
31
Differences in these environmental factors
581
2011b), and data from 2004-2009 for “eastern Norway” are adapted from Bright et al.
582
(2014a). L(E+T),  s , and Ts for “western Norway” are means over the same time period
583
acquired from MODIS retrievals (ORNL DAAC, 2014). Radiation budget variables for
584


“western Norway” ( RSW
; RLW
) are from NASA (2014).
Environmental Control Examples

H
RN
 s
RSW

∆ Definition
2
Evergreen Needleaf,
western Norway (59.4 N;
6.1 E) – Evergreen
Needleleaf, eastern
Norway (61.2 N, 12.4 E)
Evergreen Needleleaf, SE
USA – Broadleaf
Deciduous, SE USA
Evergreen Needleleaf,
easternNorway –
Broadleaf Deciduous,
eastern Norway
2
L( E  T )
,
[Wm ]
[Wm ]
-0.02
2 (= 113 –
4 (= 50 -
(Bowen)
-0.8 (= 0.4
(= 0.10
- 0.12)
111)
46 )
– 1.2)
Biological Control Example
-0.04 0 (= 194 – 12 (= 108 -0.1 (= 0.39
-0.09
Ts [ C ]
Ta [ C ] ,
,
Local
1.4
Local
-1.4b [H =
14 - H = 24 ]
0.4a
0.3b [H = 30
194)
– 96)
– 0.44)
- H = 28]
0 (=111 –
7 (=46 –
0.5 (= 1.2 – 0.1
0.4b [H = 24
111)
39)
0.8)
- H = 21]

and an emissivity of 0.95
RLW
585
a
Calculated with
586
b
Calculated with a 12-hr. heating cycle and boundary layer mixing height of 250 m (West et al. 2011)
587
588
The importance of “Biological Controls” – or vegetation structure and physiology – becomes
589
apparent when one compares sites sharing identical environmental forcings, as in Table 4.
590
The lower  s of the coniferous relative to deciduous sites in both geographic regions is
591
attributed to a higher LAI during fall-spring months (Bright et al., 2014a, Juang et al., 2007).
592
The lower  s translates to higher radiation loads ( RN ) and higher H, resulting in mean air and
593
surface temperatures that were higher at the coniferous site relative to the deciduous site in the
594
SE USA, despite having higher L(E+T) (lower Bowen ratio) during 2005 (Juang et al.,
595
2007). As for the Norwegian site comparison, despite having larger RN loads due to the
32
596
higher fall-spring LAI and lower  s at the coniferous site, Ts was much less pronounced
597
owed to a larger surface roughness and lower aerodynamic resistance.
598
Radiative vs. Non-Radiative Contributions
599
The contribution by aerodynamic and physiological (non-radiative) relative to radiative
600
effects on Ts can be quantified in terms of the energy redistribution efficiency parameter
601
(i.e., f in Eq. (4)) formalized by Lee et al. (2011). This term can either be estimated using
602
aerodynamic and surface resistances (the latter is needed to estimate L(E+T)) as in Eq. (4) –
603
or with Eq. (13) below:
604
f 
0
Ts  Ta
( RN  RG )  1
(13)
605
where RN  RG is the net turbulent heat flux, 0 is the longwave radiation feedback, and
606
Ts  Ta is the temperature gradient between the surface and the air. With this information, f
607
can be approximated for different forest biomes or management regimes (Table 5), and the
608
model of Lee et al. (2011) (Eq. 3) can be used to approximate Ts between any two sites
609
provided they share the same background climate conditions (i.e., the environmental controls
610
are identical). Table 5 shows estimates of f (unitless) for a site cluster in boreal Canada
611
together with important micrometeorological and biophysical observations. Knowing f , one
612
can approximate the contribution to Ts by non-radiative mechanisms with Eq. (3). A high f
613
denotes a greater role played by the non-radiative mechanisms at keeping Ts low.
614
Table 5 reveals some interesting insights about the relative contributions of f on the annual
615
mean Ts across the Canadian boreal site cluster. Although  s is lower and RN is higher at the
616
Old Jack Pine site relative to the Old Aspen site, Ts is lower owed to a higher f , which we
33
617
suspect is due to a larger aerodynamic roughness serving to dissipate heat (H) more efficiently
618
away from the surface during the daytime, despite its higher Bowen ratio relative to the Old
619
Aspen stand over the observed time period.
620
Table 5. Site cluster comparison of the impacts of surface intrinsic biogeophysical properties
621
on annual mean local air ( Ta ) and surface ( Ts ) temperatures (24-hr.). Ts and L( E  T ) data
622
are from MODIS (ORNL DAAC, 2014), f is calculated with Eq. (13), and all other variables
623
are from Fluxnet (Barr, 2013, Barr & Black, 2013a, Barr & Black, 2013b, Barr et al., 2013).
624
Ta is measured 1-2 m above the surface.
s
Old
Aspen
(SKOA19)
Old Jack
Pine (SKOJP14)
Clear-cut
Pine (SKHJP02)
0

RSW
(1   s )
RN
[Wm2 ]
[Wm 2 ]
0.15
110.0
60.7
1.55
0.68
0.220
0.62
14.3
0.12
114.8
62.8
1.06
0.51
0.220
1.22
24.2
0.29
94.2
44.4
0.61
0.25
0.221
0.91
26.3
Clear-cut Pine – Old
Pine
Ts [ C ]
Ta [ C ]
[ C (Wm 2 ) 1 ]
H
L( E  T )
(Bowen)
Clear-cut Pine –
Old Aspen
Old Aspen – Old Pine
Ts
[ C ] -0.47
-0.49
0.79
Ts ,
[ C ] -0.20
-0.29
0.49
f
Modeled
625
626
Relative to the Old Jack Pine and Old Aspen sites, the value of f and the non-radiative
627
contribution to the observed and modeled Ts at the Clear-cut Pine site is surprisingly high,
628
which we suspect could stem from high rates of soil evaporation. The higher Bowen ratio2 at
2
Although L(E+T) is estimated over a spatial footprint much larger than the actual site area (~1 km), the Bowen
ratios trends are consistent with those reported in Eugster et al. (Eugster W, Rouse WR, Pielke Sr. RA et al.
(2000) Land-atmosphere energy exchange in Arctic tundra and boreal forest: available data and feedbacks to
34
629
the Old Jack Pine site could be attributed to a number of factors like lower transpiration
630
(relative to Old Aspen) and lower soil evaporation (relative to Clear-cut Pine) due to a ground
631
surface that is masked by the vegetated canopy year round (i.e., more shaded ground area).
632
Inserting the values for f into Eq. (1), the non-radiative contribution to Ts from 2004-2010
633
at the Old Aspen site relative to the Old Jack Pine site is 0.56 ˚C/yr while that from RFSFC
 s is -
634
0.07 ˚C/yr. Relative to the Old Jack Pine and Old Aspen sites, non- RFSFC
 s contributions to
635
Ts at the Clear-cut site are 0.03 ˚C/yr and -0.16 ˚C/yr, respectively. Although the modeled
636
Ts is not entirely accurate since we do not subtract the storage term ( RG ) from RN (as shown
637
in Eq. (13), it gives us the right sign of the forcing and, perhaps more importantly, a way to
638
estimate the approximate relative contribution of radiative vs. non-radiative processes to Ts
639
that one may attribute to FCC/FMC.
640
6. Critical Research Needs and Future Directions
641
Climate Sensitivity and RF Efficacy
642
In recent years, radiative forcing contributions from  s have been increasingly included in
643
climate impact assessment studies. Yet the extent to which a RFTOA
 s from local  s affects
644
global surface temperature is complicated, particularly when compared to the global forcing
645
from well-mixed CO2.
646
temperature change in response to RF (in C(Wm2 )1 ) – depends on the spatial distribution of
647
the RF (Hansen & Nazarenko, 2004, Hansen et al., 2005, Joshi et al., 2003). RFs at high
648
latitudes can be over twice as effective as RFs at low latitudes (Forster et al., 2000, Hansen &
649
Nazarenko, 2004, Hansen et al., 1997, Hansen et al., 2005, Joshi et al., 2003). This result is
The equilibrium climate sensitivity  – or the equilibrium
climate. Global Change Biology, 6, 84-115.) for open vs. closed canopy and coniferous vs. deciduous forests
types in boreal regions.
35
650
due to the stimulation of positive snow/ice albedo feedbacks and to the relative stability of the
651
atmospheric temperature profile at high latitudes (Hansen & Nazarenko, 2004, Hansen et al.,
652
2005). This has given rise to RF adjustments with a factor sometimes referred to as climate
653
“efficacy” (Hansen et al., 2005), which is defined as the ratio of  for some forcing agent
654
relative to that for CO2. For instance, Zhao & Jackson (2014) applied an efficacy of 0.5 to
655
adjust albedo change RFs connected to mid-latitude FCC/FMC in the southeast USA, and
656
Cherubini et al. (2012) applied a higher efficacy of 1.94 for a change in snow albedo because
657
the forest management question mostly pertained to seasonal (winter-spring) albedo changes
658
connected to changes in snow masking by forest canopies.
659
Recall from Sections 1 & 2, however, that changes to the aerodynamic and physiological
660
properties of the surface also act on near-surface temperatures by governing the efficiency by
661
which RN is dissipated from (or accumulated by) the surface following a RF. These non-
662
radiative internal feedbacks dampen the externally-driven radiative temperature change at the
663
surface (i.e., Eq. (3)) (Davin & de Noblet-Ducoudré, 2010, Lee et al., 2011). For these
664
reasons, Davin et al. (2007) report a  connected to historical global land use changes (VCC)
665
of 0.52 C(Wm2 )1 which gives a climate efficacy of ~0.5 for the particular climate model.
666
However, Hansen et al. (2005) report an efficacy of 1.02 +/- 0.6 for their global historical
667
VCC simulations using the same vegetation maps (i.e., (Ramankutty & Foley, 1999)), which
668
demonstrates the dependency of  (and efficacy terms) on the particular climate model from
669
which they are derived.
670
Table 6. Reported global climate sensitivities (  ; C(Wm2 )1 ) and/or efficacies (  / CO
671
; unitless) for RFTOA
s .
s
Model
(VCC)
 , RFTOA

 Type
s
Efficacy
s
36
Reference
2
IPSL-CM4
0.52
Effective
0.5
(Davin et al., 2007)
IPSL
0.93a
Effective
0.78
(Davin & de NobletDucoudré, 2010)
GISS E vIII
0.45
Effective
1.02
(Hansen et al., 2005)
IAP RAS
0.49
Effective
N/A
(Eliseev, 2011)
CCSM4 v8
0.62
Equilibrium
0.79
(Jones et al., 2013)
CCSM4 v4
0.36b
Transient
N/A
(Lawrence et al., 2012)
CM
FCC only; b Ta
/ RFSFC

672
a
673
However, in a more recent modeling study limited to global-scale deforestation (FCC), Davin
674
and de Noblet-Ducoudré (2010) report a  of 0.93 C(Wm2 )1 – a value resembling that of
675
CO2 and thus giving an efficacy of ~0.8 for the particular model and set-up. Thus to some
676
extent uncertainty also stems from the type of vegetation changes that are modeled.
677
Jones et al. (2013) showed that global radiative forcings from well-mixed GHGs like CO2 and
678
those from VCC (  s ) – even if adjusted with efficacies – do not produce the same global
679
mean temperature response when added together, in part because of spatial heterogeneity,
680
non-radiative effects, and other factors.
681
responses should be added directly.
682
This argument should not detract from efforts to move down the cause-effect chain from  s
683
TOA
to RFTOA
 s . The metric RF s provides information about the net energy gained or lost from
684
the climate system and can be quantified for stand level perturbations, whereas a global
685
temperature response cannot unless quantified using a fully coupled global climate model and
686
a relatively large perturbation signal. If climate modelers worked towards quantifying and
s
They argue instead that the individual climate
37
687
building consensus on regional responses from VCC/VMC-induced RFTOA
 s , then non-climate
688
modelers and resource managers could better characterize FCC/FMC impacts. Davin & de
689
Noblet-Decoudré (2010) present latitudinal averages of the climate response from their
690
deforestation modeling simulations; if the corresponding RFTOA
 s values were also known,
691
researchers could derive regional climate sensitivities (or RF efficacies) for site-level
692
application in climate impact assessment studies.
693
Metrics of Forest Cover and Management Changes and Their Policy Implications
694
Figure 4 illustrates the principle biogeophysical forcing mechanisms at play following
695
TOA
vegetation perturbation on land.  s can be causally and linearly linked to RFSFC
 s , RF s ,
696
Ta , global , and any C-normalized metric through C-eq.’s. These metrics can be converted from
697
each other, although the conversion factors and/or procedures differ from one analyst to
698
another. Differences stem from the particular radiative transfer code used to convert  s to
699
TOA
RFTOA
 s and in the choice of the climate sensitivity term (  ) required to convert from RF s to
700
Ta , global .
701
38
702
Figure 4. Conceptual illustration of the biogeophysical mechanisms modulating local and
703
global Ta and corresponding metrics. “E” = L(E+T).
704
Using  s as the only biogeophysical forcer in land use studies should be done with caution.
705
While the  s metrics shown in Fig. 4 can be derived and converted from one to another with
706
ease, local Ta and Ts cannot. This is owed to the non-linear role of heat dissipation by
707
surface roughness and evapotranspiration. Relative to open areas, for example, forests in
708
many extra-tropical regions have generally been shown to cool locally (Ta and Ts) during the
709
daytime despite have lower surface albedos (Lee et al., 2011, Peng et al., 2014, Zhang et al.,
710
2014, Zhao & Jackson, 2014). Although H and L( E  T ) are intimately linked to Ts (In
711
Eq. (3) and Fig. (4)), Ta is also affected by turbulent mixing and the dynamics of the
712
boundary layer (Baldocchi & Ma, 2013, Oke, 2002) and has no direct relationship with Ts
713
(recall Fig. 3, where Ts and Ta displayed different signs between the coniferous example
714
(“Old Jack Pine”) and the open area example (“Clear-cut”) in most seasons).
715
Summary and Recommendations
716
Biogeophysical factors associated with forestry activities -- including albedo and turbulent
717
heat exchange -- are rarely considered by policy makers, despite the fact that such factors can
718
affect local climate in ways counter to carbon sequestration. Researchers need to assist policy
719
makers if they are to move beyond a strictly carbon-centric accounting framework for forest
720
mitigation activities. Recent suggestions to include global mean impacts from albedo change
721
RFs are an improvement but fall short of a full biophysical accounting. Forestry impacts on
722
local Ta are often more important than effects on the mean global Ta , despite the challenges
723
associated with quantifying them or comparing direct observations between sites, since air as
724
a fluid is very dynamic and unpredictable. For analyses at site level, it is often easier to
39
725
justify taking the difference between surface biophysical variables as they impact vertical heat
726
exchanges (i.e., L(E+T), H) and Ts because they are directly determined by the canopy-
727
ground composites.
728
Metrics based on RF will have greater policy relevance if appropriate adjustments are made to
729
account for differences in the local/regional response by temperature due to the internal
730
feedbacks from the non-radiative forcings.
731
coupled
732
micrometeorologists to quantify the energy redistribution efficiency parameter f (see Eq. 13)
733
for a variety of forest types and other terrestrial ecosystems.
734
The differences between local and global effects are relevant for mitigation activities
735
involving forestry. The net radiative forcing to date from CO2 emissions accompanying
736
global deforestation is ~0.4 W m-2; the accompanying global effect of increased surface
737
albedo is about -0.2 W m-2, but the local albedo effect can be two orders or magnitude greater,
738
as much as ~20 W m-2 in boreal and arid temperate forests, for instance (Betts 2007,
739
Houspanossian et al. 2013)(Chapin III et al., 2012, Rotenberg & Yakir, 2010). Thus some
740
forestry activities will cool globally while warming the land surface locally. Similarly, the
741
increased evapotranspiration of forests compared to grasslands or croplands often cools the
742
land surface locally. Globally the direct effect of increased L(E+T) is less clear as the net
743
global energy balance will effectively be zero when the water condenses elsewhere.
744
However, if the extra water vapor increases cloud cover, then a cooling factor may be
745
introduced due to enhancements in atmospheric albedo (Ban-Weiss et al., 2011). A small
746
warming factor is also introduced because water is a potent greenhouse gas. Determining the
747
net effect of these interactions remains difficult and requires both meso- and global-scale
748
models. Understanding these indirect interactions should be a future research priority.
climate
modelers
to
This approach requires sustained efforts by
quantify
40
regional
climate
sensitivities
and
by
749
As for the direct biogeophysical climate forcings connected to land use and land management,
750
we have reviewed and identified different approaches and metrics to quantify them. We have
751
also recommended research priorities to help overcome some of the challenges associated
752
with measuring radiative and non-radiative forcings. Such knowledge should help build
753
bridges among the climate modeling, forest ecology, and resource management communities
754
and, ultimately, allow us to include all biogeophysical forcings in our estimates of the climate
755
benefits of different land use activities.
756
Acknowledgements
757
This work was performed under the project ‘Approaches for integrated assessment of forest
758
ecosystem services under large scale bioenergy utilization’ funded by the Norwegian
759
Research Council (grant number: 233641/E50). Additional funding has been provided by
760
USDA-AFRI (grant number: 2012-00857).
761
References
762
763
764
765
766
767
768
769
770
771
772
773
774
775
776
777
778
779
780
781
Abiodun BJ, Pal JS, Afiesimama EA, Gutowski WJ, Adedoyin A (2008) Simulation of West African
monsoon using RegCM3 Part II: impacts of deforestation and desertification. Theoretical and
Applied Climatology, 93, 245-261.
Allen RG, Jensen ME, Wright JL, Burman RD (1989) Operational Estimates of Reference
Evapotranspiration. Agron. J., 81, 650-662.
Allen RG, Pereira LS, Raes D, Smith M (1998) Crop evapotranspiration - Guidelines for computing crop
water requirements - FAO irrigation and drainage paper 56. pp 15, Rome, Italy, UN. FAO.
Anderson-Teixeira K, Snyder P, Twine T, Cuadra S, Costa M, Delucia E (2012) Climate-regulation
services of natural and agricultural ecoregions of the Americas. Nature Clim. Change, 2.
Anderson RG, Canadell JG, Randerson JT et al. (2010) Biophysical considerations in forestry for
climate protection. Frontiers in Ecology & Environment, 9, 174-182.
Arneth A, Sitch S, Bondeau A et al. (2010) From biota to chemistry and climate: towards a
comprehensive description of trace gas exchange between the biosphere and atmosphere.
Biogeosciences, 7, 121-149.
Arora VK, Montenegro A (2011) Small temperature benefits provided by realistic afforestation
efforts. Nature Geoscience, 4, 514-518.
Avissar R, Werth D (2005) Global Hydroclimatological Teleconnections Resulting from Tropical
Deforestation. Journal of Hydrometeorology, 6, 134-145.
Bala G, Caldeira K, Wickett M, Phillips TJ, Lobell DB, Delire C, Mirin A (2007) Combined climate and
carbon-cycle effects of large-scale deforestation. PNAS, 104, 6550-6555.
41
782
783
784
785
786
787
788
789
790
791
792
793
794
795
796
797
798
799
800
801
802
803
804
805
806
807
808
809
810
811
812
813
814
815
816
817
818
819
820
821
822
823
824
825
826
827
828
829
830
831
832
833
Baldocchi D, Ma S (2013) How will land use affect air temperature in the surface boundary layer?
Lessons learned from a comparative study on the energy balance of an oak savanna and
annual grassland in California, USA. Tellus B; Vol 65 (2013).
Ban-Weiss GA, Bala G, Cao L, Pongratz J, Caldeira K (2011) Climate forcing and response to idealized
changes in surface latent and sensible heat. Environmental Research Letters, 6, 034032.
Barr A (2013) Saskatchewan BERMS 1994 Harvested Jack Pine Site Gap-filled Meteorological Dataset.
Available online [http://fluxnet.ornl.gov] from Oak Ridge National Laboratory Distributed
Active Archive Center (ORNL DAAC), Oak Ridge, Tennessee, USA. Accessed July 2, 2013.
Barr A, Black A (2013a) Saskatchewan BERMS 2002 Harvested Jack Pine Site Gap-filled
Meteorological Dataset. Available online [http://fluxnet.ornl.gov] from Oak Ridge National
Laboratory Distributed Active Archive Center (ORNL DAAC), Oak Ridge, Tennessee, USA.
Accessed July 2, 2013.
Barr A, Black A (2013b) Saskatchewan BERMS Old Aspen Site Gap-filled Meterological Dataset
Available online [http://fluxnet.ornl.gov] from Oak Ridge National Laboratory Distributed
Active Archive Center (ORNL DAAC), Oak Ridge, Tennessee, USA. Accessed July 2, 2013.
Barr A, Black A, Mccaughey H (2013) Saskatchewan BERMS Old Jack Pine Site Gap-filled Meterological
Dataset Available online [http://fluxnet.ornl.gov] from Oak Ridge National Laboratory
Distributed Active Archive Center (ORNL DAAC), Oak Ridge, Tennessee, USA. Accessed July 2,
2013.
Bartlett PA, Mackay MD, Verseghy DL (2006) Modified snow algorithms in the Canadian land surface
scheme: Model runs and sensitivity analysis at three boreal forest stands. AtmosphereOcean, 44, 207-222.
Betts AK, Desjardins RL, Worth D (2007) Impact of agriculture, forest and cloud feedback on the
surface energy budget in BOREAS. Agricultural and Forest Meteorology, 142, 156-169.
Betts R (2007) Implications of land ecosystem-atmosphere interactions for strategies for climate
change adaptation and mitigation. Tellus B, 59, 602-615.
Boisier JP, De Noblet-Ducoudré N, Pitman AJ et al. (2012) Attributing the impacts of land-cover
changes in temperate regions on surface temperature and heat fluxes to specific causes:
Results from the first LUCID set of simulations. Journal of Geophysical Research:
Atmospheres, 117, D12116.
Bonan GB (2002) Ecological Climatology: Concepts and Applications, Cambridge, U.K., Cambridge
University Press.
Bonan GB (2008) Forests and Climate Change: Forcings, Feedbacks, and the Climate Benefits of
Forests. Science, 320, 1444-1449.
Bright RM, Antón-Fernández C, Astrup R, Cherubini F, Kvalevåg MM, Strømman AH (2014a) Climate
change implications of shifting forest management strategy in a boreal forest ecosystem of
Norway. Global Change Biology, 20, 607-621.
Bright RM, Kvalevåg MM (2013) Technical note: Evaluating a simple parameterization of radiative
shortwave forcing from surface albedo change. Atmospheric Chemistry and Physics, 13,
11169-11174.
Bright RM, Myhre G, Astrup R, Antón-Fernández C, Strømman AH (2014b) Radiative forcing bias of
surface albedo modifications linked to simulated forest cover changes at northern latitudes
Biogeosciences Discussions, Manuscript in Review.
Caiazzo F, Malina R, Staples MD, Wolfe P, J. ,, Yim SHL, Barrett SRH (2014) Quantifying the climate
impacts of albedo changes due to biofuel production: a comparison with biogeochemical
effects. Environmental Research Letters, 9, 024015.
Cess RD (1978) Biosphere-Albedo Feedback and Climate Modeling. Journal of the Atmospheric
Sciences, 35, 1765-1768.
Chapin Iii FS, Matson PA, Vitousek P (2012) Principles of terrestrial ecosystem ecology, 2nd Ed.,
Springer.
Chapin Iii FS, Shaver GR, Giblin AE, Nadelhoffer KJ, Laundre JA (1995) Responses of Arctic Tundra to
Experimental and Observed Changes in Climate. Ecology, 76, 694-711.
42
834
835
836
837
838
839
840
841
842
843
844
845
846
847
848
849
850
851
852
853
854
855
856
857
858
859
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
Chen G-S, Notaro M, Liu Z, Liu Y (2012) Simulated Local and Remote Biophysical Effects of
Afforestation over the Southeast United States in Boreal Summer*. Journal of Climate, 25,
4511-4522.
Cherubini F, Bright RM, Strømman AH (2012) Site-specific global warming potentials of biogenic CO2
for bioenergy: contributions from carbon fluxes and albedo dynamics. Environmental
Research Letters, 7, 045902.
Colaizzi PD, Evett SR, Howell TA, Tolk JA (2004) Comparison of aerodynamic and radiometric surface
temperature using precision weighing lysimeters. pp 215-229.
Cotton WR, Pielke RA (1995) Human impacts on weather and climate, Cambridge, UK; New York, NY,
USA; Melbourne, Australia, Cambridge University Press.
Davin EL, De Noblet-Ducoudré N (2010) Climatic impact of global-scale deforestation: Radiative
versus nonradiative processes. Journal of Climate, 23, 97-112.
Davin EL, De Noblet-Ducoudré N, Friedlingstein P (2007) Impact of land cover change on surface
climate: Relevance of the radiative forcing concept. Geophysical Research Letters, 34,
L13702.
De Noblet-Ducoudré N, Boisier J-P, Pitman A et al. (2012) Determining Robust Impacts of Land-UseInduced Land Cover Changes on Surface Climate over North America and Eurasia: Results
from the First Set of LUCID Experiments. Journal of Climate, 25, 3261-3281.
Ding R, Kang S, Du T, Hao X, Zhang Y (2014) Scaling Up Stomatal Conductance from Leaf to Canopy
Using a Dual-Leaf Model for Estimating Crop Evapotranspiration. PLoS ONE, 9, e95584.
Durieux L, Machado LaT, Laurent H (2003) The impact of deforestation on cloud cover over the
Amazon arc of deforestation. Remote Sensing of Environment, 86, 132-140.
Eliseev AV (2011) Comparison of climatic efficiency of the mechanisms of land-surface albedo
changes caused by land use. Izvestiya, Atmospheric and Oceanic Physics, 47, 290-300.
Essery R (2013) Large-scale simulations of snow albedo masking by forests. Geophysical Research
Letters, 40, 5521-5525.
Essery R, Rutter N, Pomeroy JW et al. (2009) SnowMIP2: An evaluation of forest snow process
simulations. Bulletin of the American Meteorological Society, August, 1-16.
Eugster W, Rouse WR, Pielke Sr. RA et al. (2000) Land-atmosphere energy exchange in Arctic tundra
and boreal forest: available data and feedbacks to climate. Global Change Biology, 6, 84-115.
Fao, Jrc (2012) Global forest land-use change 1990-2005. (eds Lindquist EJ, D'annunzio R, Gerrand A,
Macdicken K, Achard F, Beuchle R, Brink A, Eva HD, Mayaux P, San-Miguel-Ayanz J, Stibig H-J),
Rome, Italy, FAO Forestry Paper N. 169, Food and Agriculture Organization of the United
Nations and European Commission Joint Research Centre.
Feddema JJ, Oleson KW, Bonan GB, Mearns LO, Buja LE, Meehl GA, Washington WM (2005) The
Importance of Land-Cover Change in Simulating Future Climates. Science, 310, 1674-1678.
Flanner MG, Shell KM, Barlage M, Perovich DK, Tschudi MA (2011) Radiative forcing and albedo
feedback from the Northern Hemisphere cryosphere between 1979 and 2008. Nature Geosci,
4, 151-155.
Foley JA, Costa MH, Delire C, Ramankutty N, Snyder P (2003) Green surprise? How terrestrial
ecosystems could affect earth’s climate. Frontiers in Ecology and the Environment, 1, 38-44.
Forster PM, Blackburn M, Glover R, Shine KP (2000) An examination of climate sensitivity for
idealised climate change experiments in an intermediate general circulation model. Climate
Dynamics, 16, 833-849.
Ghimire B, Williams CA, Masek J, Gao F, Wang Z, Schaaf C, He T (2014) Global albedo change and
radiative cooling from anthropogenic land-cover change, 1700 to 2005 based on MODIS,
land-use harmonization, radiative kernels and reanalysis. Geophysical Research Letters,
2014GL061671.
Goldewijk KK (2001) Estimating global land use change over the past 300 years: The HYDE Database.
Global Biogeochemical Cycles, 15, 417-433.
Hansen J, Nazarenko L (2004) Soot climate forcing via snow and ice albedos. Proceedings of the
National Academy of Sciences of the United States of America, 101, 423-428.
43
886
887
888
889
890
891
892
893
894
895
896
897
898
899
900
901
902
903
904
905
906
907
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
Hansen J, Sato M, Ruedy R (1997) Radiative forcing and climate response. Journal of Geophysical
Research: Atmospheres, 102, 6831-6864.
Hansen J, Sato M, Ruedy R et al. (2005) Efficacy of climate forcings. Journal of Geophysical Research:
Atmospheres, 110, D18104.
Hedstrom NR, Pomeroy JW (1998) Measurements and modelling of snow interception in the boreal
forest. Hydrological Processes, 12, 1611-1625.
Henderson-Sellers A, Wilson MF (1983) Albedo Observations of the Earth's Surface for Climate
Research. Phil. Trans. R. Soc. B, 309, 285-294.
Hoffmann WA, Jackson RB (2000) Vegetation–Climate Feedbacks in the Conversion of Tropical
Savanna to Grassland. Journal of Climate, 13, 1593-1602.
Hollinger DY, Ollinger SV, Richardson AD et al. (2010) Albedo estimates for land surface models and
support for a new paradigm based on foliage nitrogen concentration. Global Change Biology,
16, 696-710.
Idso SB, Jackson RD, Reginato RJ, Kimball BA, Nakayama FS (1975) The Dependence of Bare Soil
Albedo on Soil Water Content. Journal of Applied Meteorology, 14, 109-113.
Irmak S, Mutiibwa D, Irmak A, Arkebauer TJ, Weiss A, Martin DL, Eisenhauer DE (2008) On the scaling
up leaf stomatal resistance to canopy resistance using photosynthetic photon flux density.
Agricultural and Forest Meteorology, 148, 1034-1044.
Jackson RB, Randerson JT, Canadell JG et al. (2008) Protecting climate with forests. Environmental
Research Letters, 3, 044006 (044005pp).
Jones AD, Collins WD, Torn MS (2013) On the additivity of radiative forcing between land use change
and greenhouse gases. Geophysical Research Letters, 40, 4036-4041.
Joshi M, Shine K, Ponater M, Stuber N, Sausen R, Li L (2003) A comparison of climate response to
different radiative forcings in three general circulation models: towards an improved metric
of climate change. Climate Dynamics, 20, 843-854.
Juang J-Y, Katul G, Siqueira M, Stoy P, Novick K (2007) Separating the effects of albedo from ecophysiological changes on surface temperature along a successional chronosequence in the
southeastern United States. Geophysical Research Letters, 34, L21408.
Katul G, Oren R (2011a) Duke Forest Hardwoods Flux Data. AmeriFlux Site and Data Exploration
System. Accessed online Jan. 5, 2015 at: http://ameriflux.ornl.gov/ .
Katul G, Oren R (2011b) Duke Forest Loblolly Pine Ameri-Flux Data. AmeriFlux Site and Data
Exploration System. Accessed online Jan. 5, 2015 at: http://ameriflux.ornl.gov/.
Kelliher FM, Leuning R, Raupach MR, Schulze ED (1995) Maximum conductances for evaporation
from global vegetation types. Agricultural and Forest Meteorology, 73, 1-16.
Kleidon A (2006) The climate sensitivity to human appropriation of vegetation productivity and its
thermodynamic characterization. Global and Planetary Change, 54, 109-127.
Klingaman NP, Butke J, Leathers DJ, Brinson KR, Nickl E (2008) Mesoscale Simulations of the Land
Surface Effects of Historical Logging in a Moist Continental Climate Regime. Journal of
Applied Meteorology and Climatology, 47, 2166-2182.
Lawrence PJ, Feddema JJ, Bonan GB et al. (2012) Simulating the Biogeochemical and Biogeophysical
Impacts of Transient Land Cover Change and Wood Harvest in the Community Climate
System Model (CCSM4) from 1850 to 2100. Journal of Climate, 25, 3071-3095.
Lee X, Goulden ML, Hollinger DY et al. (2011) Observed increase in local cooling effect of
deforestation at higher latitudes. Nature, 479, 384-387.
Luyssaert S, Jammet M, Stoy P et al. (2014) Land management and land-cover change have impacts
of similar magnitude on surface temperature. Nature Climate Change, 4, 389-393.
Mahmood R, Pielke RA, Hubbard KG et al. (2013) Land cover changes and their biogeophysical effects
on climate. International Journal of Climatology, In Press, doi:10.1002/joc.3736.
Mahmood R, Quintanar AI, Conner G et al. (2010) Impacts of Land Use/Land Cover Change on
Climate and Future Research Priorities. Bulletin of the American Meteorological Society, 91,
37-46.
44
937
938
939
940
941
942
943
944
945
946
947
948
949
950
951
952
953
954
955
956
957
958
959
960
961
962
963
964
965
966
967
968
Mcguire AD, Chapin FS, Walsh JE, Wirth C (2006) Integrated Regional Changes in Arctic Climate
Feedbacks: Implications for the Global Climate System. Annual Review of Environment and
Resources, 31, 61-91.
Mcnaughton KG, Spriggs TW (1986) A mixed-layer model for regional evaporation. Boundary-Layer
Meteorology, 34, 243-262.
Meyfroidt P, Rudel TK, Lambin EF (2010) Forest transitions, trade, and the global displacement of
land use. Proceedings of the National Academy of Sciences, 107, 20917-20922.
Mohr KI, David Baker R, Tao W-K, Famiglietti JS (2003) The Sensitivity of West African Convective Line
Water Budgets to Land Cover. Journal of Hydrometeorology, 4, 62-76.
Monteith JL (1965) Evaporation and environment. Symp. Soc. Exp. Biol., 19, 205-224.
Monteith JL, Unsworth MH (eds) (2008) Principles of environmental physics, London, Elsevier
Academic Press.
Mu Q, Zhao M, Running SW (2011) Improvements to a MODIS global terrestrial evapotranspiration
algorithm. Remote Sensing of Environment, 115, 1781-1800.
Muñoz I, Campra P, Fernández-Alba AR (2010) Including CO2-emission equivalence of changes in land
surface albedo in life cycle assessment. Methodology and case study on greenhouse
agriculture. International Journal of Life Cycle Assessment, 15, 672-681.
Murry FW (1967) On the computation of saturation vapor pressure. Journal of Applied Meteorology,
6, 203-204.
Nakai T, Sumida A, Daikoku K et al. (2008) Parameterisation of aerodynamic roughness over boreal,
cool- and warm-temperate forests. Agricultural and Forest Meteorology, 14, 1916-1925.
Nasa Asdc (2014) NASA/GEWEX Solar Radiation Budget (SRB) Data, Release v.3.0. Accessed
November 18, 2014 at: https://eosweb.larc.nasa.gov/project/srb/srb_table . NASA
Atmospheric Science and Data Center.
National Research Council (2003) Understanding Climate Change Feedbacks. pp 166, Panel on
Climate Change Feedbacks, Climate Research Committee, National Research Council.
Niu G-Y, Yang Z-L (2004) Effects of vegetation canopy processes on snow surface energy and mass
balances. Journal of Geophysical Research: Atmospheres, 109, D23111.
Norwegian Meteorological Institute (2013) eKlima - Monthly Historical Meteorology. Norwegian
Meteorological Institute. Accessed Jan 31, 2013 at:
http://sharki.oslo.dnmi.no/portal/page?_pageid=73,39035,73_39049&_dad=portal&_schem
a=PORTAL
969
970
971
972
973
974
975
976
977
978
979
980
981
982
983
984
985
986
987
Oke TR (2002) Boundary layer climates, Taylor & Francis e-Library.
Ollinger SV, Richardson AD, Martin ME et al. (2008) Canopy nitrogen, carbon assimilation, and albedo
in temperate and boreal forests: Functional relations and potential climate feedbacks.
Proceedings of the National Academy of Sciences, 105, 19336-19341.
Ornl Daac (2014) MODIS subsetted land products, Collection 5. Available online
[http://daac.ornl.gov/MODIS/modis.html] from ORNL DAAC, Oak Ridge, Tennessee, USA.
Accessed Nov. 28, 2014., Oak Ridge National Laboratory Distributed Active Archive Center
(ORNL DAAC).
Otterman J (1977) Anthropogenic impact on the albedo of the earth. Climatic Change, 1, 137-155.
Peng S-S, Piao S, Zeng Z et al. (2014) Afforestation in China cools local land surface temperature.
Proceedings of the National Academy of Sciences, 111, 2915-2919.
Pereira LS, Perrier A, Allen RG, Alves I (1999) Evapotranspiration: concepts and future trends. J. Irrig.
Drain. Engrg., 125, 45-51.
Perrier A (1982) Land surface processes: Vegetation. In: Land surface processes in atmospheric
general circulation models. (ed Eagleson PS) pp 395-448. Cambridge, U.K., Cambridge
University Press.
Pielke RA (2001) Influence of the spatial distribution of vegetation and soils on the prediction of
cumulus Convective rainfall. Reviews of Geophysics, 39, 151-177.
45
988
989
990
991
992
993
994
995
996
997
998
999
1000
1001
1002
1003
1004
1005
1006
1007
1008
1009
1010
1011
1012
1013
1014
1015
1016
1017
1018
1019
1020
1021
1022
1023
1024
1025
1026
1027
1028
1029
1030
1031
1032
1033
1034
1035
1036
1037
Pielke Sr. RA, Adegoke J, Beltrán-Przekurat A et al. (2007) An overview of regional land-use and landcover impacts on rainfall. Tellus B, 59, 587-601.
Pielke Sr. RA, Avissar R, Raupach MR, Dolman AJ, Zeng X, Denning AS (1998) Interactions between the
atmosphere and terrestrial ecosystems: influence on weather and climate. Global Change
Biology, 4, 461-475.
Pielke Sr. RA, Marland G, Betts RA et al. (2002) The influence of land-use change and landscape
dynamics on the climate system: relevance to climate-change policy beyond the radiative
effect of greenhouse gases. Phil. Trans. R. Soc. Lond. A, 360, 1705-1719.
Pielke Sr. RA, Pitman A, Niyogi D et al. (2011) Land use/land cover changes and climate: modeling
analysis and observational evidence. WIREs Climate Change, 2, 828-850.
Pirazzini R (2009) Challenges in snow and ice albedo parameterizations. Geophysica, 45, 41-62.
Pitman AJ, De Noblet-Ducoudré N, Cruz FT et al. (2009) Uncertainties in climate responses to past
land cover change: First results from the LUCID intercomparison study. Geophysical Research
Letters, 36, L14814.
Planton S (2013) Annex III: Glossary. In: Climate Change 2013: The Physical Science Basis.
Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental
Panel on Climate Change. (eds Stocker TF, Qin D, Plattner G-K, Tignor M, Allen SK, Boschung
J, Nauels A, Xia Y, Bex V, Midgley PM). Cambridge, UK and New York, NY, USA, Cambridge
University Press.
Pongratz J, Reick C, Raddatz T, Claussen M (2008) A reconstruction of global agricultural areas and
land cover for the last millennium. Global Biogeochemical Cycles, 22, GB3018.
Qu X, Hall A (2007) What Controls the Strength of Snow-Albedo Feedback? Journal of Climate, 20,
3971-3981.
Ramankutty N, Foley JA (1999) Estimating historical changes in global land cover: Croplands from
1700 to 1992. Global Biogeochemical Cycles, 13, 997-1027.
Ray DK, Welch RM, Lawton RO, Nair US (2006) Dry season clouds and rainfall in northern Central
America: Implications for the Mesoamerican Biological Corridor. Global and Planetary
Change, 54, 150-162.
Roesch A, Roeckner E (2006) Assessment of Snow Cover and Surface Albedo in the ECHAM5 General
Circulation Model. Journal of Climate, 19, 3828-3843.
Rotenberg E, Yakir D (2010) Contribution of semi-arid forests to the climate system. Science, 327,
451-454.
Rustad L, Campbell J, Marion G et al. (2001) A meta-analysis of the response of soil respiration, net
nitrogen mineralization, and aboveground plant growth to experimental ecosystem warming.
Oecologia, 126, 543-562.
Scholes R, Settele J (2014) Chapter 4: Terrestrial and Inland Water Systems - Final Draft. In: Climate
Change 2014: Impacts, Adaptation, and Vulnerability. Contribution by Working Group II to
the Fifth IPCC Assessment Report. (eds Fischlin A, Moreno JM, Root T). Geneva, Switzerland,
Intergovernmental Panel on Climate Change (IPCC).
Scott CE, Rap A, Spracklen DV et al. (2014) The direct and indirect radiative effects of biogenic
secondary organic aerosol. Atmos. Chem. Phys., 14, 447-470.
Sellers PJ (1985) Canopy reflectance, photosynthesis, and transpiration. International Journal of
Remote Sensing, 6, 1335-1372.
Shell KM, Kiehl JT, Shields CA (2008) Using the Radiative Kernel Technique to Calculate Climate
Feedbacks in NCAR’s Community Atmospheric Model. Journal of Climate, 21, 2269-2282.
Smith P, Bustamante M (2013) Chapter 11 - Final Draft: Agriculture, Forestry, and Other Land Use
(AFOLU). In: In: Climate Change Mitigation - Contribution by Working Group III to the Fifth
IPCC Assessment Report. Geneva.
Soden BJ, Held IM, Colman R, Shell KM, Kiehl JT, Shields CA (2008) Quantifying Climate Feedbacks
Using Radiative Kernels. Journal of Climate, 21, 3504-3520.
46
1038
1039
1040
1041
1042
1043
1044
1045
1046
1047
1048
1049
1050
1051
1052
1053
1054
1055
1056
1057
1058
1059
1060
1061
1062
1063
1064
1065
1066
1067
1068
1069
1070
1071
1072
1073
1074
1075
1076
1077
1078
1079
Spracklen DV, Bonn B, Carslaw KS (2008) Boreal forests, aerosols and the impacts on clouds and
climate. Philosophical Transactions of the Royal Society A: Mathematical, Physical and
Engineering Sciences, 366, 4613-4626.
Stoy P, Lin H, Novick K, Siqueira M, Juang J-Y (2014) The Role of Vegetation on the Ecosystem
Radiative Entropy Budget and Trends Along Ecological Succession. Entropy, 16, 3710-3731.
Stoy PC, Katul GG, Siqueira MBS et al. (2006) Separating the effects of climate and vegetation on
evapotranspiration along a successional chronosequence in the southeastern US. Global
Change Biology, 12, 2115-2135.
Swann ALS, Fung I, Chiang JCH (2011) Mid-latitude afforestation shifts general circulation and tropical
precipitation. PNAS, 109, 712-716.
Tetens O (1930) Uber einige meteorologische Begriffe. z. Geophys., 6, 297-309.
Thom AS (1972) Momentum, mass and heat exchange of vegetation. Quarterly Journal of the Royal
Meteorological Society, 98, 124-134.
Unger N (2014) Human land-use-driven reduction of forest volatiles cools global climate. Nature Clim.
Change, advance online publication.
Van Vuuren D, Edmonds J, Kainuma M et al. (2011) The representative concentration pathways: an
overview. Climatic Change, 109, 5-31.
Verseghy DL, Mcfarlane NA, Lazare M (1993) CLASS - A Canadian land surface scheme for GCMs. II.
Vegetation model and coupled runs. International Journal of Climatology, 13, 347-370.
Wang K, Dickinson RE (2012) A review of global terrestrial evapotranspiration: observation,
modeling, climatology, and climatic variability. Review of Geophysics, 50, RG2005.
Wang Y, Yan X, Wang Z (2014) The biogeophysical effects of extreme afforestation in modeling future
climate. Theoretical and Applied Climatology, DOI: 10.1007/s00704-013-1085-8.
Wesely ML, Hicks BB (1977) Some Factors that Affect the Deposition Rates of Sulfur Dioxide and
Similar Gases on Vegetation. Journal of the Air Pollution Control Association, 27, 1110-1116.
West PC, Narisma GT, Barford CC, Kucharik CJ, Foley JA (2011) An alternative approach for
quantifying climate regulations by ecosystems. Frontiers in Ecology & Environment, 9, 126133.
Winton M (2005) Simple optical models for diagnosing surface-atmosphere shortwave interactions.
Journal of Climate, 18, 3796-3806.
Wiscombe WJ, Warren SG (1980) A model for the spectral albedo of snow. I. Pure Snow. Journal of
Atmospheric Science, 37, 2712-2733.
Zhang B, Liu Y, Xu D, Cai J, Li F (2011) Evapotranspiration estimation based on scaling up from leaf
stomatal conductance to canopy conductance. Agricultural and Forest Meteorology, 151,
1086-1095.
Zhang M, Lee X, Yu G et al. (2014) Response of surface air temperature to small-scale land clearing
across latitudes. Environmental Research Letters, 9, 034002.
Zhao K, Jackson RB (2014) Biophysical forcings of land-use changes from potential forestry activities
in North America. Ecological Monographs, 84, 329-353.
Zheng Y, Yu G, Qian Y, Miao M, Zeng X, Liu H (2002) Simulations of regional climatic effects of
vegetation change in China. Quarterly Journal of the Royal Meteorological Society, 128, 20892114.
1080
47