ArticlePDF Available

Changes in sub-alpine tree distribution in western North America: A review of climatic and other causal factors

Authors:

Abstract

Changes in the distribution of sub-alpine tree species in western North America have been attributed to climatic change and other environmental stresses. These changes include tree-line fluctuations throughout the Holocene and recent invasion of sub-alpine meadows by forest. Most palaeoecological studies suggest that the tree-line was higher during a period of warmer climate approximately 9000 to 5000 BP and lower during the last 5000 years, with short periods of local tree-line advance. Recent advances in sub-alpine tree distribution can be compared with weather records, allowing an examination of relationships between tree advance and climate at a finer resolution. In general, recent sub-alpine forest advances in western North America, based on studies representing three climatic zones (maritime, Mediterranean and continental), have been associated with climatic periods favouring tree germination and growth, although factors such as fire and grazing by domestic livestock have had an impact in some areas. Limitations to tree establishment (e.g., winter snow accumulation, summer drought) vary in relative importance within each climate zone, as do predicted consequences of anthropogenic climatic change. Recent increases in establishment of sub-alpine trees may continue if climatic change alleviates the limitations to tree establishment important in each climatic zone. However, factors such as topography and disturbance may modify tree establishment on a local scale.
http://hol.sagepub.com
The Holocene
DOI: 10.1177/095968369400400112
1994; 4; 89 The Holocene
Regina M. Rochefort, Ronda L. Little, Andrea Woodward and David L. Peterson
causal factors
Changes in sub-alpine tree distribution in western North America: a review of climatic and other
http://hol.sagepub.com/cgi/content/abstract/4/1/89
The online version of this article can be found at:
Published by:
http://www.sagepublications.com
can be found at:The Holocene Additional services and information for
http://hol.sagepub.com/cgi/alerts Email Alerts:
http://hol.sagepub.com/subscriptions Subscriptions:
http://www.sagepub.com/journalsReprints.navReprints:
http://www.sagepub.co.uk/journalsPermissions.navPermissions:
http://hol.sagepub.com/cgi/content/refs/4/1/89
SAGE Journals Online and HighWire Press platforms):
(this article cites 100 articles hosted on the Citations
at UNIV ARIZONA LIBRARY on September 4, 2008 http://hol.sagepub.comDownloaded from
89-
Changes
in
sub-alpine
tree
distribution
in
western
North
America:
a
review
of
climatic
and
other
causal
factors
Regina
M.
Rochefort,
Ronda
L.
Little,
Andrea
Woodward
and
David
L.
Peterson*
(National
Park
Service,
Cooperative
Park
Studies
Unit,
College
of
Forest
Resources,
University
of
Washington,
AR-10,
Seattle,
Washington
98195,
USA)
Received
4
November
1993;
revised
manuscript
accepted
3
May
1993
Abstract :
Changes
in
the
distribution
of
sub-alpine
tree
species
in
western
North
America
have
been
attributed
to
climatic
change
and
other
environmental
stresses.
These
changes
include
tree-line
fluctuations
throughout
the
Holocene
and
recent
invasion
of
sub-alpine
meadows
by
forest.
Most
palaeoecological
studies
suggest
that
the
tree-line
was
higher
during
a
period
of
warmer
climate
approximately
9000
to
5000
BP
and
lower
during
the
last
5000
years,
with
short
periods
of
local
tree-line
advance.
Recent
advances
in
sub-alpine
tree
distribution
can
be
compared
with
weather
records,
allowing
an
examination
of
relationships
between
tree
advance
and
climate
at
a
finer
resolution.
In
general,
recent
sub-alpine
forest
advances
in
western
North
America,
based
on
studies
representing
three
climatic
zones
(maritime,
Mediterranean
and
continental),
have
been
associated
with
climatic
periods
favouring
tree
germination
and
growth,
although
factors
such
as
fire
and
grazing
by
domestic
livestock
have
had
an
impact
in
some
areas.
Limitations
to tree
establishment
(e.g.,
winter
snow
accumulation,
summer
drought)
vary
in
relative
importance
within
each
climate
zone,
as
do
predicted
consequences
of
anthropogenic
climatic
change.
Recent
increases
in
establishment
of
sub-alpine
trees
may
continue
if
climatic
change
alleviates
the
limitations
to
tree
establishment
important
in
each
climatic
zone.
However,
factors
such
as
topography
and
disturbance
may
modify
tree
establishment
on
a
local
scale.
Key
words:
climatic
change,
meadow
invasion,
sub-alpine
forest,
tree
establishment,
tree-line,
Holo-
cene,
North
America.
Introduction
Climate
change,
sub-alpine
forests
and
ecotones
Most
general
circulation
models
predict
an
increase
in
mean
annual
temperature
of
1-5°C
during
the
next
century
as
the
result
of
increased
levels
of
atmospheric
C02
and
other
greenhouse
gases
(Schneider,
1989).
Climatological
evidence
indicates
that
mean
annual
air
temperature
at
the
earth’s
surface
has
already
increased
approximately
0.6°C
since
AD
1880
(Jones et
al.,
1986;
Hansen
and
Lebedeff,
1988).
Even
small
changes
in
temperature,
such
as
the
1-1.5°C
cooler
period
of
the
’Little
Ice
Age’
(c.
AD
1650
to
1850;
Porter,
1981),
can
cause
significant
changes
in
vegetation
(Lamb,
1982;
Brubaker,
1988;
Sprugel,
1991).
Future
precip-
itation
patterns
are
also
predicted
to
change,
but
existing
models
predict
changes
in
only
magnitude
and/or
seasonal
pattern
(e.g.,
wetter
winters
with
drier
summers -
Kellogg
and
Zhao,
1988;
Silver
and
DeFries,
1990).
Even
small
changes
in
global
temperature
or
precipitation
patterns
are
likely
to
alter
conditions
affecting
earth’s
ecosystems
(Houghton
and
Woodwell,
1989),
resulting
in
changes
in
vegetation
patterns.
Linking
the
response
of
vegetation
to
future
changes
in
climate
is
a
major
challenge
facing
ecologists.
Detecting
initial
effects
of
environmental
change
on
vegetation
may
be
diffi-
cult
due
to
the
high
variability
within
most
biological
systems.
One
promising
approach
is
to
focus
on
ecotones -
transitions
between
adjacent
communities.
Ecotones
resulting
from
envi-
ronmental
gradients
(as
opposed
to
disturbances
such
as
fire)
may
be
particularly
responsive
to
climatic
changes
because
organisms
are
already
at
some
limit
to
their
existence.
Eco-
tones
at
the
sub-alpine/alpine
boundary
may
be
particularly
relevant
to
studies
of
climatic
change
because
the
physi-
ognomy
of the
dominant
vegetation
changes
from
trees
to
shrubs
and
herbs
(Peterson,
1991;
Rochefort
and
Peterson,
1991;
Woodward et
al.,
1991).
Tree
distribution
at
high
elevation
is
at
least
partially
limited
by
temperature
and
* Author
to
whom
correspondence
should
be
addressed.
at UNIV ARIZONA LIBRARY on September 4, 2008 http://hol.sagepub.comDownloaded from
90
snowpack
(Wardle,
1974;
Tranquillini,
1979;
Stevens
and
Fox,
1991),
factors
certain
to
change
if
temperatures
rise
in
the
future.
Palaeoecological
studies indicate that
altitudinal
tree-line
has
fluctuated
in
response
to
climate
throughout
the
Holo-
cene
(Pears,
1968;
Markgraf
and
Scott,
1981;
Luckman
and
Kearney,
1986;
COHMAP,
1988),
and
that
expanding
ranges
of
tree
species
are
often
associated
with
periods
of
warmer
climate
(LaMarche,
1973;
Kearney
and
Luckman,
1983;
Grace,
1989;
Kullman,
1991).
Recent
studies
provide
the
opportunity
to
discover
local
factors,
in
addition
to
climate,
responsible
for
enhancing
or
preventing
tree
establishment.
Studies
in
western
North
America
suggest
recent
expansions
in
the
spatial
distribution
of
sub-alpine
tree
species
at
many
locations.
Terminology
and
scope
Many
different
terms
are
used
to
describe
ecotonal
relation-
ships
of
high-altitude
tree
species
(Love, 1970;
Douglas,
1972;
Wardle,
1974;
Tranquillini,
1979).
We
use
the
following
terminology
in
this
discussion.
The
sub-alpine
park
land
is
the
highest-elevation
belt
that
contains
tree
species,
consisting
of
a
mosaic
of
individual
trees,
tree
clumps,
and
meadows
that
extend
from
closed
forest
to
alpine
(treeless)
vegetation
(Figure
1 -
Henderson,
1974;
Franklin
and
Dyrness,
1987).
The
lower
boundary
of
the
sub-alpine
park
land
is
indicated
by
the
upper
limit
of closed
contiguous
forest
or
forest
line.
Tree-line
indicates
the
highest
elevation
at
which
erect
trees
are
found
within
the
sub-alpine
park
land.
Tree-limit
indicates
the
highest
elevation
at
which
any
trees
are
found,
including
prostrate
or
krummholz
growth
forms.
This
paper
focuses
on
altitudinal
species
distribution,
although
’tree-line’
and
’tree-
limit’
can
also
be
used
to
describe
latitudinal,
edaphic
and
maritime
tree
species
limits
(Hustich,
1983;
Payette,
1983).
Information
is
available
on
worldwide
tree-line
dynamics
Figure
1
Diagrammatic
representation
of
sub-alpine
conifer
distribu-
tion.
A
mosaic
of
trees
and
meadows
dominates
the
ecotone
between
the
continuous
forest
below
and
the
treeless
alpine
above.
and
on
factors
limiting
sub-alpine
forest
distribution
(Dau-
benmire,
1954;
Zimina,
1973;
Wardle,
1977;
Gorchakovsky
and
Shiyatov,
1978;
Payette
and
Gagnon,
1979;
Tranquillini,
1979;
Kullman,
1986).
We
focus
on
western
North
America
because:
(1)
many
data
exist
on
recent
changes
in
sub-alpine
forest
distribution;
and
(2)
many
sub-alpine
mountain
loca-
tions
in
this
area
lie
within
protected
areas
relatively
free
of
human
influence
(Billings,
1969;
Holtmeier,
1989).
The
im-
pacts
of
woodcutting,
grazing
and
other
activities
more
com-
mon
in
eastern
North
America
and
Europe
make
it
difficult
to
interpret
the
influence
of
climate
and
other
’natural’
environmental
forces
on
sub-alpine
vegetation
(Holtmeier,
1973;
Ives
and
Hansen-Bristow,
1983;
Helms,
1987).
Climatic
regimes
in
western
North
America
are
highly
varied,
and
ecosystems
range
from
temperate
rainforest
to
desert.
However,
climatic
regimes
in
mountainous
areas
can
be
classified
by
their
dominant
weather
patterns
and
geo-
graphic
location
into:
maritime,
continental
and
Mediterra-
nean
zones
(Schroeder
and
Buck,
1970;
Trewartha,
1980;
Critchfield,
1983).
Figure
2
indicates
the
location
of the
climatic
zones
and
principal
mountain
ranges
discussed
in
this
paper.
Alaska
and
Mexico
are
not
included
because
we
have
found
no
published
reports
of
recent
altitudinal
changes
in
the
distribution
of
sub-alpine
forests
in
these
areas
(Griggs,
1937;
Beaman,
1962;
Lauer,
1978).
The
climatic
zones
are
not
intended
to
correspond
exactly
with
macroscale
or
synoptic
climatic
classifications.
Characteristics
of
the
three
climatic
regimes
differ
sub-
stantially.
Maritime
climate -
predominant
in
the
northwest-
ern
portion
of the
continental
United
States
and
southwestern
Canada -
is
characterized
by
high
coastal
rainfall,
a
long
winter
period
of
rain
and
snow,
and
extensive
cloud
cover.
Winter
temperatures
are
cool,
summer
tem-
peratures
are
mild,
and
summer
precipitation
is
low.
Con-
tinental
climate
prevails
in
the
western
interiors
of
the
continental
United
States
and
Canada.
Much
of
the
winter
precipitation
is
snow,
and
summer
precipitation
occurs
as
thunderstorms
in
the
central
and
southern
part
of the
range.
Winter
temperatures
are
cold;
summer
temperatures
vary
from
cool
in
the
north
to
warm
in
the
south.
Mediterranean
climate
is
predominant
in
the
far
southwestern
continental
United
States.
Annual
precipitation
is
low,
and
most
of
it
falls
during
winter.
Summers
are
hot
and
dry,
with
persistent
droughts
in
many
years,
and
winter
temperatures
are
mild.
Several
authors
have
reviewed
the
causes
of
tree-line
and
tree
species
limit
(Griggs,
1946;
Daubenmire,
1954;
Wardle,
,
1971;
Tranquillini,
1979;
Arno
and
Hammerly,
1984;
Innes,
1991;
Stevens
and
Fox,
1991);
however,
these
reviews
empha-
size
environmental
impacts
on
growth
and
survival
of
mature
trees
rather
than
on
seedling
establishment.
In
this
paper,
we
focus
on
the
effects
of
environmental
factors
on
sub-alpine
tree
establishment
and
initial
survival,
because
this
early
stage
of
forest
development
determines
tree-line
advance
and
meadow
invasion.
We
discuss
changes
in
tree
establishment
with
respect
to
fluctuations
in
altitudinal
tree-line
and
tree-
limit,
and
forest
advance
into
sub-alpine
meadows.
Although
most
of
the
literature
cited
in
this
paper
is
from
North
America,
general
concepts
may
be
applicable
to
sub-alpine
forests
in
other
areas
of
the
world,
especially
the
temperate
zone
of
Europe
(Troll,
1973).
In
order
to
evaluate the
effects
of
climate
on
the
distribution
of
sub-alpine
species,
we
discuss
how
weather
components
(e.g.,
snow,
wind
and
temperature),
as
well
as
natural
disturbance,
limit
the
distribution
of
sub-
alpine
trees.
We
then
review
palaeoecological
and
modern
evidence
for
sub-alpine
vegetation
changes
during
the
Holo-
cene,
and
we
infer
the
relative
importance
of
environmental
variables
associated
with
these
changes.
Finally,
we
project
at UNIV ARIZONA LIBRARY on September 4, 2008 http://hol.sagepub.comDownloaded from
91
Figure
2
Location
of
climatic
zones
(maritime,
continental,
Mediterranean)
and
major
mountain
ranges
of
western
North
America.
how
potential
climatic
change
could
affect
sub-alpine
forests
in
the
future.
Environmental
influences
on
seedling
establishment
and
survival
Patterns
of
snow
accumulation
and
snowmelt
either
promote
or
hinder
seedling
establishment
and
survival.
Snowpack
shields
conifer
needles
from
the
desiccating
winter
winds
common
at
upper
elevations
(Wardle,
1968;
Lindsay,
1971;
Hadley
and
Smith,
1983;
Minnich,
1984)
and
insulates
seed-
ling
roots
and
foliage
from
lethal
cold
temperatures
(Kur-
amoto
and
Bliss,
1970).
Snowmelt
delays
summer
drought
and
provides
increased
soil
moisture
early
in
the
growing
season
when
tree
seeds
germinate
(Evans
and
Fonda,
1990).
Snowpack
may
be
especially
beneficial
to
seedling
establish-
ment
in
the
windswept
continental
regions
and
dry
Medi-
terranean
mountains.
On
the
other
hand,
heavy
snow
accumulation
that
remains
on
the
ground
until
midsummer
results
in
a
shorter
growing
season
and
reduced
seedling
establishment
(Brink,
1959;
Fonda
and
Bliss,
1969;
Kuramoto
and
Bliss,
1970;
Douglas,
1972;
Heikkinen,
1984;
Butler,
1986;
Armstrong et
al.,
1988;
Hansen-Bristow et
al.,
1988).
Pro-
longed
snow
cover
also
promotes
infection
by
brown
felt
fungi
(Herpotrichia
spp.),
which
can
damage
and
kill
foliage
(Donaubauer,
1963;
Simms,
1967;
Holtmeier,
1987).
Al-
though
some
tree
seeds
can
germinate
directly
on
snow,
they
generally
have
a
low
survival
rate
(Franklin
and
Krueger,
1968;
Brooke
et
al.,
1970),
and
even
established
seedlings
can
be
physically
crushed
or
uprooted
by
the
weight
and
move-
ment
of
snow
(Ives
and
Hansen-Bristow,
1983).
The
majority
of
these
negative
effects
of
snow
on
establishment
are
re-
ported
for
maritime
regions,
where
deep
snowpacks
develop.
The
effect
of
both
snowdrift
and
wind
patterns
on
tree
survival
in
the
Rocky
Mountains
is
demonstrated
by
the
creation
of
gradually
moving
ribbon
forests
aligned
perpen-
dicular
to
westerly
winds
(Billings,
1969;
Benedict,
1984;
Holtmeier, 1987).
Availability
of
soil
moisture
is
critical
for
seedling
germina-
tion
and
survival,
particularly
during
the
first
year
(Cui
and
Smith, 1991).
Soil
moisture
depends
on
rainfall
and
snowmelt,
as
modified
by
soil
properties,
aspect,
microtopography
and
evapotranspiration
(Fonda,
1976;
Sawyer
and
Kinraide,
1980).
Periods
of
above
average
summer
precipitation
can
enhance
tree
establishment
(Agee
and
Smith,
1984;
Allen,
1984;
Taylor,
1990),
and
increased
water
availability
from
a
melting
snowpack
can
be
important
for
seedling
survival
at UNIV ARIZONA LIBRARY on September 4, 2008 http://hol.sagepub.comDownloaded from
92
during
summer
droughts
(Billings, 1969;
Lindsay,
1971;
Cana-
day
and
Fonda,
1974).
Seedling
establishment
can
also
occur
when
a
wet
period
provides
adequate
soil
moisture
for
tree
seedlings
following
a
series
of
dry
years
that
reduces
the
competitive
ability
of
meadow
vegetation
(Vale,
1981;
Butler,
1986;
Taylor,
1990).
Foliar
desiccation
can
occur
during
winter
when
the
transpirational
demand
of
foliage
exceeds
the
ability
of the
roots
to
provide
water.
This
problem
is
exacerbated
by
frozen
or
cold
soils,
where
water
absorption
by
roots
is
reduced.
Winter
drying
of
tissues
controls
tree
morphology
and
distribution
at
many
high-elevation
locations
(Marr,
1977;
Tranquillini,
1979;
Ives
and
Hansen-Bristow,
1983;
Bella
and
Navratil,
1987).
Adequate
soil
moisture
is
important
for
seedlings
in
all
climatic
zones
because
they
all
have
dry
periods;
however,
winter
desiccation
may
be
more
prevalent
in
the
continental
zone,
because
high
winds
can
remove
snow
and
expose
soil
to
freezing
temperatures.
Summer
moisture
stress
is
one
of
the
primary
limits
to
tree-
line
advance
in
Mediterranean
areas,
especially
when
com-
bined
with
warm
temperatures
(Klikoff,
1965;
LaMarche,
1973).
Extreme
soil
temperatures
have
a
negative
impact
on
the
survival
of
tree
seedlings
(Wardle,
1968;
Munn et
al.,
1978;
DeLucia
and
Smith,
1987).
High
soil
temperatures
during
early
summer
can
be
lethal
to
seedlings
(Baig,
1972;
Ballard,
1972;
Douglas,
1972),
although
warm
summers
generally
benefit
seedling
survival
(Kearney,
1982).
Years
with
warm
air
temperatures
can
promote
tree
establishment
in
meadows
(Brink,
1959;
Franklin,
Moir et
al.,
1971;
Lowery,
1972;
Kearney
and
Luckman,
1983;
Heikkinen,
1984;
Clague
and
Mathewes,
1989;
MacDonald,
1989),
while
periods
with
cold
air
temperatures
can
inhibit
tree
establishment
(Brink,
1964;
Daly
and
Shankman,
1985).
Seedling
establishment
can
also
be
limited
in
the
immediate
path
of
cold
air
drainage
and
melt
water
originating
from
snow
banks
(Wardle,
1968;
Fonda
and
Bliss,
1969;
Brooke et
al.,
1970;
Moore,
1991).
Freezing
soil
and
air
temperature
can
damage
tissues
directly
through
intercellular
freezing
and
indirectly
by
dehydration
resulting
from
extracellular
freezing
(Kramer
and
Kozlowski,
1979;
Tranquillini,
1979).
Early
autumn
freezes
and
late
spring
freezes
may
be
especially
deleterious
to
seedling
tissues
that
are
not
cold-hardened
(Lindsay,
1971;
Ives
and
Hansen-
Bristow,
1983).
Seedlings
are
more
sensitive
than
mature
trees
to
frost
heaving
caused
by
differential
freezing
of
soil
layers
because
their
shallow
roots
break
or are
exposed
to
desiccation
(Brink,
1964;
Kramer
and
Kozlowski,
1979).
In
general,
seedlings
in
continental
and
Mediterranean
regions
are
exposed
to
extreme
cold
temperatures,
while
seedlings
in
maritime
regions
are
protected
by
deep
snow.
In
continental
regions,
consistent
cold
temperatures
during
summer
in
com-
bination
with
desiccating
winds
may
result
in
inadequate
maturation
of
new
tissue,
which
is
subsequently
killed
during
winter
(Hadley
and
Smith,
1986).
Disturbances
such
as
avalanches
and
fire
strongly
affect
sub-alpine
vegetation
patterns.
Species
composition
after
a
physical
disturbance
often
depends
on
the
type,
extent
and
intensity
of
disturbance
(Oliver et
al.,
1985).
For
example,
a
snow
avalanche
can
produce
a
gradient
of
disturbance
condi-
tions
within
its
path,
allowing
various
regeneration
mecha-
nisms
to
occur,
such
as
advance
regeneration,
resprouting,
or
establishment
of
species
with
light
seeds
(Cushman,
1981).
Frequent
avalanches
(15-20
years
apart)
in
the
Alberta
Rocky
Mountains
cause
trees
to
be
replaced
by
shrubs
in
avalanche
paths
(Johnson,
1987).
The
role
of
natural
fires
in
maintaining
sub-alpine
meadow
communities
has
been
shown
in
many
studies
(Kuramoto
and
Bliss, 1970;
De-Benedetti
and
Parson,
1979;
Vale,
1981;
Butler,
1986;
Helms,
1987;
Shank-
man
and
Daly,
1988).
Both
natural
fires
and
those
caused
by
people,
especially
by
early
European
settlers
and
Native
Americans,
have
favoured
open
meadows
(Vankat
and
Ma-
jor,
1978).
Fires
near
the
tree-line
are
usually
small
but
can
result
in
new
snowdrift
patterns
and
long-lasting
replacement
by
meadow
vegetation
(Billings,
1969).
Establishment
of sub-
alpine
trees
following
fire
is
often
slow
because
of
snow
creep,
substrate
instability,
herbivory
or
unfavourable
climate
(Agee
and
Smith,
1984;
Little
and
Peterson,
1991).
Fire
is
the
most
important
disturbance
in
the
Rocky
Mountains,
and
periods
of
frequent
and
intense
fires
have
created
open
forests
in
the
sub-alpine
park
land
(MacDonald,
1989;
Moore,
1991).
The
tree-line in
Colorado
can
be
depressed
below
its
climatic
upper
limit
in
areas
frequented
by
fire
(Peet,
1981;
Shankman
and
Daly,
1988).
Fire
is
also
common
in
regions
with
Medi-
terranean
climate
but
less
frequent
in
maritime
climatic
zones.
Interpretations
of
other
factors
controlling
tree-line
and
tree-limit
must
carefully
examine
impacts
of
disturbance
on
sub-alpine
vegetation
patterns.
Changes
in
sub-alpine
forest
distribution
during
the
Holocene
Fluctuating
tree-lines
and
changing
climatic
regimes
during
the
Holocene
are
recorded
by
pollen
records,
wood
fragments
and
macrofossils.
The
majority
of
palaeoecological
studies
are
from
the
continental
zone,
while
studies
of
recent
tree
establishment
are
concentrated
in
the
maritime
zone.
The
causes
of
these
recent
changes
are
often
difficult
to
ascertain,
because
the
changes
coincide
with
the
end
of
the
’Little
Ice
Age’
and
the
expansion
of
European
settlement
in
western
North
America.
Palaeoecological
evidence
of
tree-line
dynamics
Palaeoecological
studies
generally
indicate
a
warmer
climate
during
the
early
to
mid-Holocene
(9000
to
5000
BP)
with
tree-line
advances
(Table
1 -
Andrews et
al.,
1975;
Petersen
and
Mehringer,
1976;
Carrara et
al.,
1984;
Luckman
and
Kearney,
1986;
Clague
and
Mathewes,
1989;
Clague et
al.,
1992).
Cooler
climates
since
5000
BP
have
generally
resulted
in
stable
or
lower-elevation
tree-lines.
Detailed
reconstruc-
tions
of
past
precipitation
patterns
and
temperature
regimes
are
based
on
pollen
analysis
(Maher,
1972;
Luckman
and
Kearney,1986;
Clague
and
Mathewes,
1989),
isotope
analysis
(Carrara et
al.,
1984;
Friedman et
al.,
1988;
Carrara et
al.,
1991)
and
tree-ring
growth
patterns
(LaMarche
and
Mooney,
1972;
Payette et
al.,
1989).
These
reconstructions
describe
a
warm/dry
early
Holocene
in
the
maritime
and
Mediterranean
areas
(LaMarche
and
Mooney, 1972;
LaMarche, 1973;
Clague
and
Mathewes,
1989),
but
a
warm/moist
early
period
in
the
continental
zone
(Markgraf
and
Scott,
1981;
Friedman
et
al.,
1988;
Carrara
et al. ,
1991).
Maritime
Tree-lines
in
southwestern
British
Columbia
9100-8200
BP
were
approximately
60-130
m
higher
than
today
(Clague
and
Mathewes,
1989;
Clague et
al.,
1992).
Using
adiabatic
lapse
rates
(6.5°C
per
1000 m),
the
authors
estimated
that
tem-
perature
was
0.4-0.8°C
warmer
at
this
time.
Wood
fragments
from
colluvial
and
alluvial
sediments
and
pollen
collected
from
a
bog
above
the
present
tree-line
were
identified
as
Pinus
albicaulis
and
Abies
lasiocarpa,
species
still
present
in
the
area
today.
These
fragments
had
wider
annual
rings
than
living
krummholz
from
the
same
site,
indicating
that
condi-
at UNIV ARIZONA LIBRARY on September 4, 2008 http://hol.sagepub.comDownloaded from
93
at UNIV ARIZONA LIBRARY on September 4, 2008 http://hol.sagepub.comDownloaded from
94
tions
favourable
for
tree
growth
were
present
at
high-eleva-
tion
sites
during
the
early
Holocene.
Wood
fragments
from
additional
sites
in
southern
British
Columbia
and
Alberta
represent
two
periods
of
higher
tree-line:
9100-7600
BP,
and
6600-5100
BP
(Clague
and
Mathewes,
1989).
Wood
frag-
ments
from
the
earlier
period
indicate
the
early
Holocene
may
have
been
warmer
than
the
mid-Holocene.
Pollen
de-
posited
in
the
upper
part
of
the
core
contained
high
percen-
tages
of
Cyperaceae,
Salix
and
other
wet-habitat
species,
indicating
a
recent
change
to
wetter
conditions.
Continental
Studies
in
the
continental
zone
reveal
a
similar
but
more
detailed
record
of
tree-line
movement.
Analysis
of
pollen,
macrofossils
and
oxygen-isotope
data
from
tree-rings
from
alpine
bogs
in
the
Canadian
Rocky
Mountains
indicate
a
fluctuating
tree-line
over
the
last
8800
years
(Kearney
and
Luckman,
1983;
Luckman
and
Kearney,
1986).
Pollen
ratios
(Abies/Pinus)
and
macrofossils
indicate
two
periods
of
ele-
vated
tree-line:
8800-7500
BP
and
7200-5200
BP.
Wood
frag-
ments
from
a
site
100 km
south
of the
bogs
also
provide
evidence
of
a
higher
tree-line
8300-8000
BP
(Luckman,
1988).
The
climate
was
warmer
and
drier
during
these
peri-
ods,
and
the
tree-line
was
at
least
100 m
higher.
Oxygen-
isotope
analysis
of
buried
logs
suggests
temperatures
were
at
least
0.5°C
warmer
8770-8060
BP
and
1.2-1.6°C
warmer
6000-5300
BP
(Luckman
and
Kearney, 1986).
Tree-lines
have
been
similar
to
or
lower
than
present
levels
since
4500
BP,
with
minimum
levels
during
the
last
500
years.
An
additional
study
in
Yoho
National
Park,
British
Columbia,
used
pollen
and
macrofossil
data
to
show
that
the
tree-line
was
at
least
90 m
above
the
modern
tree-line
during
the
period
8500-3000
BP,
in
response
to
warmer
climatic
conditions
(Reasoner
and
Hickman,
1989).
Tree-lines
declined
to
cur-
rent
levels
during
cooler
climates
after
3000
BP.
Studies
in
southwestern
Colorado
indicate
tree-line
and
tree-limit
were
depressed
prior
to
9800
BP,
advanced
upward
at
least
twice
between
9800
and
5400
BP,
and
have
generally
been
lower
than
or
equal
to
present
elevations
since
3500
BP
(Petersen
and
Mehringer,
1976;
Carrara et
al.,
1984;
1991).
Petersen
and
Mehringer
(1976)
used
Pinus/Picea
pollen
ratios
and
Picea
macrofossils
to
identify
three
periods
of
tree-line
advance
in
the
La
Plata
Mountains:
8500,
6700
and
2500
BP.
Studies
in
the
San
Juan
Mountains
of
Colorado
show
patterns
similar
to
those
described
by
Petersen
and
Mehringer
(1976)
(Andrews
et al.,1975;
Carrara
et al.,1984;1991).
Pollen,
wood
fragments
and
macrofossils
indicate
three
periods
of
tree-line
and
tree-limit
advance
of
at
least
80 m:
9600-7800
BP,
6700-5400
BP,
and
3100 BP.
The
tree-line
may
have been
140
m
higher
than
at
present
c.
8000
BP.
Mid-Holocene
July
temperatures
were
estimated
at
0.5-0.9°C
higher
than
today.
The
tree-line
declined
to
present-day
elevations
5400-3500
BP
and
again
after
a
short
upward
advance around
3100
BP,
when
tree
species
may
have
been
70
m
higher
than
today
(Andrews et
al.,
1975;
Carrara et
al.,
1984;
1991).
Two
studies
from
the
Colorado
Rocky
Mountains
describe
contrasting
climates
in
the
early
to
mid-Holocene.
Maher
(1972)
used
PicealPinus
pollen
ratios
to
infer
that
climates
from
10000
to
7600
BP
and
from
6700
to
3000
BP
were
cooler
and/or
wetter
than
today
and
tree-lines
were
lower.
Warming
trends
from
7600
to
6700
BP,
and
since
300
BP,
resulted
in
higher
tree-lines.
Conversely,
Short
(1984)
used
pollen
ratios
to
show
that
the
tree-line
reached
a
maximum
altitude
6500-3000
BP
and
that
it
has
since
lowered
in
response
to
climatic
cooling.
Mediterranean
Studies
of
Pinus
longaeva
in
the
White
Mountains,
California
(LaMarche, 1973),
and
Snake
Mountains,
Nevada
(LaMarche
and
Mooney, 1972),
indicate
higher
tree-lines
during
the
early
to
mid-Holocene.
Wood
fragments
above
the
current
tree-
line
at
the
California
sites
dated
7400
BP
indicate
that
the
tree-line
was
100
to
200 m
above
modem
levels
until
4200-2000
BP.
This
was
a
relatively
warm
period,
with
esti-
mated
mean
temperatures
3.5°C
higher
than
today.
Tree-line
elevation
became
lower
870-470
BP.
Analysis
of
site
condi-
tions
and
tree
growth/climate
relationships
(in
tree-rings)
showed
that
the
tree-line
was
controlled
by
precipitation
as
well
as
temperature.
Current
P.
longaeva
distribution
in
the
Snake
Mountains,
Nevada,
shows
a
progressive
dwarfing
with
increasing
altitude
as
follows:
erect,
tall-dwarf,
short-dwarf
and
krummholz
(LaMarche
and
Mooney,
1972).
The
authors
suggest
that
the
past
forest
included
only
erect
and
tall-dwarf
trees.
The
upper
boundary
of
erect
trees
was
at
least
100
m
higher
than
today,
and
tall-dwarf
trees
extended
to
the
summit
(3559
m)
of
Mount
Washington.
Surveys
conducted
in
the
southern
Sierra
Nevada
re-
constructed
similar
Holocene
tree-line
fluctuations
for
Pinus
balfouriana
(Scuderi,
1987).
Relict
wood
samples
collected
68
m
above
the
present
tree-line
were
dated
to
6300
BP.
Wide
growth
rings
indicate
warm,
favourable
growing
conditions.
Samples
collected
at
65
m
above
the
present
tree-line
were
dated
to
3530
BP,
suggesting
that
the
tree-line
was
main-
tained
at
this
elevated
position
for
about
3000
years.
Tree
mortality
at
20
m
above
the
present
tree-line
and
decreased
growth
between
2500
and
2300
BP
indicate
a
colder
climate.
The
tree-line
declined
to
10 m
below
the
present
level
between
1400
and
1300
BP,
and
increased
to
the
present
level
between
950
anjd
850
BP;
a
few
trees
established
5
to
10
m
above
the
present
tree-line
about
100
years
ago.
Summary
of
tree-line
fluctuation
.
Palaeoecological
studies
provide
a
fairly
uniform
scenario
of
climatic
change
and
tree
response
to
a
changing
environment.
The
Holocene from
10 000
to
3500
BP
was
generally
a
warmer
period
with
several
periods
of
tree-line
advance,
and
it
was
followed
by
a
cooler
period
with
tree-lines
lower
than
or
equal
to
those
of
today.
The
climate
in
the
maritime
and
Mediterranean
zones
was
generally
warm
and
dry
during
the
early
Holocene
and
became
cooler
and
more
moist
c.
3500
BP
(LaMarche,
1973;
Clague
and
Mathewes,
1989).
Conversely,
precipitation
in
the
continental
zone
has
decreased
since
the
early
Holocene
(Andrews et
al.,
1975;
Petersen
and
Mehringer,
1976;
Markgraf
and
Scott,
1981;
Carrara et
al.,
1984).
In
addition,
precipitation
patterns
have
influenced
tree
growth
form,
as
well
as
the
magnitude
and
timing
of
tree-line
movement
at
some
locations
(LaMarche, 1973).
However,
the
generally
synchronous
reconstruction
of
past
tree-line
fluctu-
ations
indicates
that
temperature
is
the
predominant
force
determining
tree-line
location
on
a
broad
geographic
scale.
Recent
tree
establishment
in
sub-alpine
meadows
Increases
in
tree
establishment
in
sub-alpine
meadows
have
been
documented
in
mountainous
areas
throughout
western
North
America
(Table
2).
Most
areas
show
initial
forest-
margin
expansion
after
AD
1890
and
significant
establish-
ment
peaks
between
1920
and
1950.
Additional
peaks
of
establishment
have
also
been
identified
in
some
areas.
Most
studies
conclude
that
recent
increases
in
tree
establishment
are
the
result
of
a
warmer,
drier
climate
following
the
’Little
Ice
Age’
(Franklin
et
al.,
1971;
Kearney,
1982;
Heikkinen,
1984;
Butler,
1986).
Although
most
studies
conclude
that
at UNIV ARIZONA LIBRARY on September 4, 2008 http://hol.sagepub.comDownloaded from
95
Table
2
Summary
of
recent
sub-alpine
tree
invasion
attributed
to
climate.
Studies
are
from
the
United
States
except
as
indicated.
:
these
increases
were
caused
by
warming
trends,
it
is
uncertain
whether
this
is
a
long-term
directional
change
or
a
short-term
variation
in
relatively
stable
ecotones.
Maritime
In
one
of
the
first
studies
on
recent
increases
of
tree
establish-
ment
in
sub-alpine
meadows,
Brink
(1959)
documented
es-
tablishment
of
Abies
lasiocarpa
and
Tsuga
mertensiana
between
1919
and
1939
in
heather
communities
(Phyllodoce
spp.lCassiope
mertensiana)
of
the
Coast
Range,
British
Co-
lumbia.
Heather
communities
were
on
topographic
con-
vexities,
with
earlier
snowmelt
and
a
longer
growing
season.
Soils
were
also
more
xeric
than
surrounding
depressions,
but
not
as
dry
as
south-facing
forb
meadows
without
trees.
Climatic
warming
during
this
period
is
supported
by
evidence
of
decreased
glacial
volumes.
Amo
(1970)
noted
massive
invasions
of
heather
commu-
nities
by
Larix
lyallii,
T.
mertensiana
and
A.
lasiocarpa
in
the
northern
Cascade
Mountains
of
Washington.
Most
of
the
establishment
occurred
from
1919
to
1937,
although
about
half
of the
L.
lyallii
had
invaded
the
area
between
1885
and
1910.
Douglas
(1972)
also
documented
meadow
invasion
by
T.
mertensiana
and
A.
amabilis
in
heather
communities
in
the
northern
Cascades.
He
attributed
establishment
to
several
consecutive
years
of
low
snow
accumulation,
moderate
spring
and
early
summer
temperatures,
and
the
black-body
effect
of
tree
clumps
that
caused
faster
snowmelt
around
trees.
Tree
establishment
had
been
occurring
for
120
years
in
his
study
area,
with
peak
establishment
from
1920
to
1940.
Franklin et
al.
(1971)
surveyed
six
areas
in
the
Cascade
at UNIV ARIZONA LIBRARY on September 4, 2008 http://hol.sagepub.comDownloaded from
96
Mountains
of
Washington
and
Oregon
and
documented
meadow
invasion
by
A.
lasiocarpa,
T.
mertensiana
and
L.
lyal-
lii
(one
site
in
the
northern
Cascades).
Most
invasion
oc-
curred
between
1925
and
1950,
predominantly
by
A.
lasiocarpa.
Seedling
density
and
growth
rates
varied
with
vegetation
type,
with
highest
densities
in
heather
commu-
nities
and
lowest
densities
in
Festuca
communities.
Invasion
was
attributed
to
a
series
of
warm,
dry
years
with
low
snowpack
and
the
possible
influence
of
large
seed
crops.
Henderson
(1974)
suggested
that
periods
of
establishment
in
this
area
occurred
during
warm
periods
with
average
or
above
average
precipitation,
which
reduced
moisture
stress
on
seedlings.
Heikkinen
(1984)
found
that
tree
advance
in
the
northern
Cascades
followed
the
same
temporal
patterns
as
noted
by
Franklin
et al.
(1971)
and
Brink
(1959).
The
oldest
trees
dated
back
to
1886,
and
tree
establishment
continued
until
1944,
with
invasion
peaks
during
1925-1934
and
1940-1944.
Tsuga
mertensiana
predominated
at
the
highest
elevations,
A.
lasio-
carpa
at
mid-elevations,
and
A.
amabilis
at
low
elevations.
In
a
study
of
stand
dynamics
in
the
northern
Cascades,
Lowery
(1972)
found
that
invasion
throughout
the
meadows
started
about
1920,
coinciding
with
the
expansion
of
older
tree
clumps,
warmer
weather
and
glacial
retreat.
Abies
lasio-
carpa
was
the
most
abundant
species,
especially
at
xeric
sites
such
as
the
centre
of
heather-dominated
mounds.
Tsuga
mertensiana
grew
on
wet,
cool
sites,
while
A.
amabilis
and
Chamaecyparis
nootkatensis
established
on
moist,
lower-ele-
vation
sites
and
the
exterior
of
tree
clumps.
Tree
invasion
and
enlargement
of
tree
clumps
continued
until
the
1950s,
when
a
cooler,
wetter
climate
prevailed
and
glacial
recession
ceased.
Studies
in
the
Olympic
Mountains
of
Washington
docu-
mented
tree
invasions
in
meadow
basins
with
normally
heavy
winter
snowpack
(Fonda
and
Bliss, 1969;
Kuramoto
and
Bliss,
1970).
These
studies
inferred
that
tree
establishment
occurred
during
periods
of
warm,
dry
weather.
There
were
several
discrete
peaks
of
establishment:
1923-1933,
1943-1948
and
1953-1960.
These
studies
also
found
that
establishment
was
highest
in
heather
communities,
and
that
tree
height
in-
creased
with
distance
from
melting
snow.
Agee
and
Smith
(1984)
studied
tree
establishment
on
three
sub-alpine
fire
sites
in
the
Olympics.
Fire
removed
all
vegeta-
tion
and
therefore
reduced
protection
from
sun,
drought
and
browsing
ungulates,
resulting
in
a
time
lag
of
40-70
years
before
seedlings
could
germinate
and
survive.
Burned
sites
had
peak
establishment
during
wet
periods
(1950s),
while
unburned
heather
sites
had
establishment
peaks
during
both
dry
periods
(1920-1950)
and
wet
periods
(1950s).
The
authors
concluded
that
mesic
conditions
were
optimal
for
tree
estab-
lishment ;
therefore,
drier,
burned
sites
required
wet
periods
for
establishment,
and
wetter,
unburned
areas
required
warm,
dry
periods.
Continental
Kearney
(1982)
noted
invasions
of
Abies
lasiocarpa
and
Picea
engelmannii
at
three
locations
in
the
Rocky
Mountains
of
southern
Alberta.
Invasion
occurred
during
1940-1960
and
was
especially
prominent
from
1965
to
1973.
Both
invasion
periods
were
correlated
with
warm,
dry
summers.
In
addition,
growth
rates
of
trees
on
heather
hummocks
were
often
twice
those
of
trees
in
the
hollows
between
hummocks.
Tree
invasion
in
the
Lemhi
Mountains
of
Idaho
began
in
the
1890s
and
continued
until
the
1950s
(Butler,
1986).
Invasion
peaks
occurred
in
1905-1919
and
1925-1945,
corre-
sponding
with
warm,
dry
periods
with
decreased
snowpack.
Tree
establishment
patterns
varied
slightly
among
the
three
meadows
studied,
indicating
some
variation
in
the
causal
agent
(climate)
or
site
conditions.
Invasion
peaks
in
the
lowest-elevation
meadow
were
earlier
than
in
the
other
two
by
about
ten
years.
Charcoal
in
the lowest
meadow
suggests
that
frequent
fire
could
have
limited
tree
establishment.
Daly
and
Shankman
(1985)
found
high
densities
of
Picea
engelmannii
and
Pinus
flexilis
seedlings
(5-27
years
old)
above
the
tree-limit
in
the
Colorado
Rocky
Mountains.
The
high
ratio
of
seedlings
to
trees
was
interpreted
as
an
increase
in
seedling
establishment,
which
could
produce
an
advance
of
the
tree-limit.
Mediterranean
Establishment
above
the
tree-limit
was
also
documented
at
one
site
in
the
Sierra
Nevada
(Scuderi,
1987)
and
two
sites
in
the
White
Mountains
of
California
(LaMarche,
1973).
Small,
erect
seedlings
of
Pinus
balfouriana
have
established
5-10
m
above
the
present
tree-line
at
Cirque
Peak,
Sierra
Nevada,
over
the
last
100
years.
Large
numbers
of
P.
longaeva
have
established
above
the
tree
limit
in
the
White
Mountains
since
1850.
Few
trees
established
between
1450
and
1850
at
one
site,
and
no
trees
established
between
1250
and
1850
at
the
other
site.
In
summary,
the
literature
indicates
that
climatic
warming
since
1850
has
resulted
in
periods
of
increased
tree
establish-
ment
in
sub-alpine
park
lands.
Studies
of
recent
tree
estab-
lishment
are
concentrated
in
the
maritime
zone,
so
it
is
difficult
to
determine
the
geographic
extent
and
variation
of
tree
establishment
in
western
North
America.
Modern
studies
demonstrate
a
more
complex
climate/site
interaction
than
can
be
inferred
from
palaeoecological
studies
of
tree-line
fluctuations.
Effect
of
potential
global
climate
change
on
tree
establishment
Our
review
indicates
that
weather
components
frequently
limit
establishment
of
trees
in
sub-alpine
and
alpine
meadows
in
the
maritime,
Mediterranean
and
continental
climatic
zones
of
western
North
America.
A
comparison
between
climatic
variables
and
limits
to
tree
establishment
(Table
3)
shows
that
the
most
extreme
characteristic
of
each
zone
is
often
the
most
limiting
factor,
specifically:
(1)
winter
precip-
itation
(snowpack)
in
maritime
climates,
(2)
cold
soil
tem-
peratures
and
high
winds
in
continental
climates,
and
(3)
summer
drought
in
Mediterranean
climates.
Our
review
suggests
that
tree
establishment
occurs
in
any
of
these
climates
when
these
extreme
conditions
are
alleviated.
Pre-
dictions
of
tree
establishment
on
a
fine
scale
are
difficult,
however,
because
establishment
patterns
are
influenced
by
local
variations
such
as
microtopography
and
disturbance
frequency.
Predictions
of
future
climate
are
based
on
numerical
mod-
els
limited
by
incomplete
understanding
of
atmospheric
and
oceanic
processes,
interactions
and
feedbacks.
Furthermore,
current
computer
capacities
limit
spatial
resolution
and
re-
strict
the
coupling
of
climate
system
components
whose
timescales
differ
by
fourteen
orders
of
magnitude
(Schle-
singer
and
Mitchell,
1987;
Cubash
and
Cess,
1990).
Models
predict
temperature
changes
most
accurately
and
are
less
able
to
predict
changes
in
precipitation
and
soil
moisture,
which
are
products
of
complex
interactions
within the
hydrologic
cycle.
Models
also
have
difficulty
in
resolving
subcontinental
variation.
Despite
weaknesses,
climate
predictions
generated
by
general
circulation
models
(GCMs)
provide
the
best
available
description
of future
climate.
We
have
summarized
at UNIV ARIZONA LIBRARY on September 4, 2008 http://hol.sagepub.comDownloaded from
97
Table
3
Levels
of
climate
variables
characteristic
of
three
climate
zones,
with
the
extremes
for
each
climate
indicated
by
bold
type.
These
extremes
are
also
the
limiting
factors
for
tree
establishment
for
each
zone.
Table
4
Range
of
changes
in
climate
variables
predicted
by
three
general
circulation
models
at
equilibrium
with
doubled
atmospheric
C02.
Values
indicate
change
from
present.
Models
were
developed
by
the
Canadian
Climate
Centre,
the
Geophysical
Fluid
Dynamics
Laboratory
and
the
United
Kingdom
Meteorological
Office
(Mitchell et
al.,
1990).
Note:
~
Represents
the
available
water
in
the
upper
layers
of
soil
having
field
capacity
of 15
cm.
the
range
of
changes
predicted
for
several
climate
variables
for
the
three
climate
zones
based
on
predictions
of
response
to
doubled
C02
from
three of
the
highest-resolution
models
(Table
4 -
Mitchell et
al.,
1990).
The
most
consistent
conclusions
of the
models
are:
(1)
winter
soil
moisture
will
increase
in
all
three
climates;
(2)
winter
and
summer
temperatures
will
increase
in
all
three
climates,
but
especially
in
the
continental
climate;
(3)
summer
soil
moisture
will
decline
in
maritime
and
continental
cli-
mates ;
and
(4)
distribution
of
precipitation
in
the
maritime
climate
will
be
skewed
even
more
heavily
towards
winter.
Predictions
for
winter
and
summer
precipitation
and
summer
soil
moisture
in
the
Mediterranean
climate
vary,
although
two
of
the
three
models
predict
an
increase
in
each
case.
Climate
model
predictions
also
provide
a
basis
for
specula-
tion
about
future
distribution
of
vegetation,
particularly
trees.
The
predicted
changes
for
all
three
climates
should
alleviate
the
identified
constraint
to
tree
establishment
(Table
3).
Winter
precipitation
is
predicted
to
increase
in
the
maritime
climate,
but
warmer
temperatures
are
expected
to
cause
more
to
fall
as
rain
at
higher
elevations,
resulting
in
less
snowpack.
Summer
drought
is
predicted
to
lessen
in
Mediterranean
climates.
Models
predict
increased
winter
temperatures
and
precipitation
in
continental
areas,
although
temperatures
will
still
be low
enough
to
produce
snow.
This
may
provide
greater
wind
and
cold
protection
than
current
conditions.
However,
changes
that
favour
sub-alpine
establishment
may
be
negated
by
other
changes
such
as
summer
drought,
which
is
predicted
to
intensify
in
maritime
and
continental
zones,
possibly
preventing
tree
establishment.
There
is,
of
course,
a
high
degree
of
uncertainty
about
how
changes
in
different
combinations
of
climatic
variables
might
affect
future
pat-
terns
of
sub-alpine
tree
establishment.
Moreover,
changes
in
other
factors
influenced
by
climate,
such
as
fire
frequency,
may
also
affect
tree
establishment
(Franklin et
al.,
1991).
In
summary,
we
expect
an
increase
in
tree
establishment
(i.e.,
higher
tree-lines,
increased
meadow
invasion)
in
all
three
climate
zones.
Tree-lines
may
be
similar
to
those
of
the
early
Holocene,
when
the
climate
was
similar
to
the
predicted
future
climate.
A
similar
prediction
has
been
made
from
a
landscape
model
for
a
single
site
in
the
continental
zone
(Romme
and
Turner,
1991).
Moreover,
we
may
be
observing
the
beginning
of
a
trend
in
response
to
the
warmer
climates
of
this
century
(Brink,
1959;
Butler,
1986;
Franklin et
al.,
1971;
LaMarche,
1973;
Kearney,
1982;
Rochefort
and
Peterson,
1991;
Woodward et
al.,
1991).
This
very
general
conclusion
must
be
qualified
because
(1)
small-scale
variability,
disturb-
ance,
availability
of
suitable
substrate,
and
other
variables
not
directly
related
to
climate
make
it
difficult
to
specify
precise
situations
in
which
establishment
will
occur;
(2)
loss
of
potential
invasion
sites
due
to
the
imposition
of
new
con-
straints
by
changing
climate
has
not
been
described;
and
(3)
the
predictions
of
GCMs
are
uncertain.
Therefore,
we
predict
that
vegetation
zones
will
not
simply
migrate
up
in
elevation
and
latitude
(Dale
and
Franklin,
1989;
Peters,
1990),
but
that
sub-alpine
species
may
be
distributed
differently
in
the
future
’1!’
than
in
the
past
(Brubaker, 1988;
Davis, 1989).
Detailed
study
of
site
conditions
and
tree
regeneration
requirements
will
help
refine
predictions
of
the
effects
of
climate
on
forest
establishment.
Long-term
monitoring
at
several
sites
will
determine
whether
recently
observed
patterns
of
sub-alpine
tree
establishment
will
continue.
Acknowledgements
We
are
grateful
to
James
Benedict,
Greg
Ettl,
Tom
Hinckley,
Darci
Horner,
John
Innes,
Brian
Luckman,
Kathleen
Mar-
uoka,
David
W.
Peterson,
June
Rugh,
Ed
Schreiner,
David
Silsbee,
Doug
Sprugel
and
an
anonymous
reviewer
for
helpful
comments
on
the
manuscript.
Beth
Rochefort
assisted
with
graphics.
This
study
was
supported
by
the
US
National
Park
Service
Global
Change
Program
and
the
USDA
Forest
Service
Pacific
Northwest
Research
Station.
at UNIV ARIZONA LIBRARY on September 4, 2008 http://hol.sagepub.comDownloaded from
98
References
Agee,
J.K.
and
Smith,
L.
1984:
Subalpine
tree
establishment
after
fire
in
the
Olympic
Mountains,
Washington.
Ecology
65,
810-19.
Allen,
C.D.
1984:
Montane
grasslands
in
the
landscape
of the
Jemaz
Mountains
of
New
Mexico.
M.S.
thesis,
University
of
New
Mexico,
Albuquerque.
Andrews,
J.T.,
Carrara,
P.E.,
King,
F.B.
and
Stuckenrath,
R.
1975:
Holocene
environmental
changes
in
the
alpine
zone,
northern
San
Juan
Mountains,
Colorado: evidence
from
bog
stratigraphy
and
palynology.
Arctic
and
Alpine
Research
5,
173-97.
Armstrong,
J.K.,
Williams,
K.,
Huenneke,
L.F.
and
Mooney,
H.A.
1988:
Topographic
position
effects
on
growth
depression
of
Cal-
ifornia
Sierra
Nevada
pines
during
the
1982-83
El
Niño.
Arctic
and
Alpine
Research
20,
352-57.
Arno,
S.F.
1970:
Ecology
of
alpine
larch
(Larix
lyallii
Parl.)
in
the
Pacific
Northwest.
Ph.D.
thesis,
University
of
Montana,
Missoula.
Arno,
S.F.
and
Hammerly,
R.P.
1984:
Timberline,
mountain
and
Arctic
forest
frontiers.
Seattle:
The
Mountaineers.
Baig,
M.N.
1972:
Ecology
of
timberline
vegetation
in
the
Rocky
Mountains
of Alberta.
Ph.D.
thesis,
University
of
Calgary,
Calgary,
Alberta.
Ballard,
T.M.
1972:
Subalpine
soil
temperature
regimes
in
southwest-
ern
British
Columbia.
Arctic
and
Alpine
Research
4, 139-46.
Beaman,
J.H.
1962:
The
timberlines
of
Iztaccihuatl
and
Popocatepetl,
Mexico.
Ecology
43,
377-85.
Bella,
I.E.
and
Navratil,
S.
1987:
Growth
losses
from
winter
drying
(red
belt
damage)
in
lodgepole
pine
stands
on
the
east
slopes
of
the
Rockies
in
Alberta.
Canadian
Journal
of
Forest
Research
17,
1289-92.
Benedict,
J.B.
1984:
Rates
of
tree-island
migration,
Colorado
Rocky
Mountains,
USA.
Ecology
65,
820-23.
Billings,
W.D.
1969:
Vegetational
pattern
near
alpine
timberline
as
affected
by
fire-snowdrift
interactions.
Vegetatio
19,
192-207.
Brink,
V.C.
1959:
A
directional
change
in
the
subalpine
forest-heath
ecotone
in
Garibaldi
Park,
British
Columbia.
Ecology
40, 10-16.
—
1964:
Plant
establishment
in
the
high
snowfall
alpine
and
subalpine
regions
of
British
Columbia.
Ecology
45,
431-38.
Brooke,
R.C.,
Peterson,
E.B.
and
Krajina,
VJ.
1970:
The
subalpine
mountain
hemlock
zone.
Ecology
of
Western
North
America
2,
148-349.
Brubaker,
L.B.
1988:
Vegetation
history
and
anticipating
future
vegetation
change.
In
Agee,
J.K.
and
Johnson,
D.R.,
editors,
Ecosys-
tem
management
for
parks
and
wilderness,
Seattle:
University
of
Washington
Press,
41-61.
Butler,
D.R.
1986:
Conifer
invasion
of
subalpine
meadows,
Central
Lemhi
Mountains,
Idaho.
Northwest
Science
60, 166-73.
Canaday,
B.B.
and
Fonda,
R.W.
1974:
The
influence
of
subalpine
snowbanks
on
vegetation
pattern,
reproduction,
and
phenology.
Bulletin
of the
Torrey
Botanical
Club
101, 340-50.
Carrara,
P.E.,
Trimble,
D.A.
and
Rubin,
M.
1991:
Holocene
fluctu-
ations
in
the
northern
San
Juan
Mountains,
Colorado,
U.S.A.,
as
indicated
by
radio-carbon-dated
conifer
wood.
Arctic
and
Alpine
Research
23,
233-46.
Carrara,
P.E.,
Mode,
W.N.,
Rubin,
M.
and
Robinson,
S.W.
1984:
Deglaciation
and
postglacial
timberline
in
the
San
Juan
Mountains,
Colorado.
Quaternary
Research
21,
42-55.
Clague,
JJ.
and
Mathewes,
R.W.
1989:
Early
Holocene
thermal
maximum
in
western
North
America:
new
evidence
from
Castle
Peak,
British
Columbia.
Geoloizv
17, 277-80.
Clague,
JJ.,
Mathewes,
R.W.,
Buhay,
W.M.
and
Edwards,
T.R.
1992.
Early
Holocene
climate
at
Castle
Peak,
south
Coast
Mountains,
British
Columbia,
Canada.
Palaeogeography,
Palaeoclimatology,
Palaeoecology
95, 153-67.
COHMAP,
1988:
Climatic
changes
of the
last
18,000
years:
observa-
tions
and
model
simulations.
Science
241,
1043-52.
Critchfield,
HJ.
1983:
General
climatology.
Englewood
Cliffs,
N.J.:
Prentice-Hall.
Cubash,
U.
and
Cess,
R.D.
1990.
Processes
and
modelling.
In
Houghton,
J.T.,
Jenkins,
G.J.,
Ephraums,
J.J.,
editors,
Climate
change:
the
IPCC
scientific
assessment,
Cambridge:
Cambridge
Uni-
versity
Press,
69-91.
Cui,
M.
and
Smith,
W.K.
1991:
Photosynthesis,
water
relations
and
mortality
in
Abies
lasiocarpa
seedlings
during
natural
establishment.
Tree
Physiology
8,
37-46.
Cushman,
MJ.
1981:
The
influence
of
recurrent
snow
avalanches
on
vegetation
patterns
in
the
Washington
Cascades.
Ph.D.
thesis,
Uni-
versity
of
Washington,
Seattle.
Dale,
V.H.
and
Franklin,
J.F.
1989:
Potential
effects
of
climate
change
on
stand
development
in
the
Pacific
Northwest.
Canadian
Journal
of
Forest
Research
19,
1581-90.
Daly,
C.
and
Shankman,
D.
1985:
Seedling
establishment
by
conifers
above
tree
limit
on
Niwot
Ridge,
Front
Range,
Colorado
U.S.A.
Arctic
and
Alpine
Research
17,
389-400.
Daubenmire,
R.F.
1954 :
Alpine
timberlines
in
the
Americas
and
their
interpretation.
Butler
University
Botanical
Studies
11,
119-36.
Davis,
M.B.
1989:
Insights
from
paleoecology
on
global
change.
Bulletin
of
the
Ecological
Society
of
America
70,
222-28.
DeBenedetti,
S.H.
and
Parsons,
DJ.
1979:
Natural
fire
in
subalpine
meadows:
a case
description
from
the
Sierra
Nevada.
Journal
of
Forestry
77,
477-79.
De
Lucia,
E.H.
and
Smith,
W.K.
1989:
Air
and
soil
temperature
limitations
on
photosynthesis
in
Engelmann
spruce
during
summer.
Canadian
Journal
of
Forest
Research
17, 527-33.
Donaubauer,
E.
1963:
On
the
snow-blight
(Phacidium
infestans)
disease
of
Pinus
cembra and
some
associated
fungi.
Mitteilungen
der
Forstlichen
Bundes-Versuchsanstalt
Mariabrunn
60,
577-600.
Douglas,
G.W.
1972:
Subalpine
plant
communities
of the
Western
North
Cascades,
Washington.
Arctic
and
Alpine
Research
4, 147-66.
Evans,
R.D.
and
Fonda,
R.W.
1990:
The
influence
of
snow
on
subalpine
meadow
community
pattern,
North
Cascades,
Washington.
Canadian
Journal
of
Botany
68,
212-20.
Fonda,
R.W.
1976:
Ecology
of
alpine
timberline
in
Olympic
National
Park.
Proceedings,
conference
on
scientific
research
in
National
Parks
1,209-12.
Fonda,
R.W.
and
Bliss,
L.C.
1969:
Forest
vegetation
of
the
montane
and
subalpine
zones,
Olympic
Mountains,
Washington.
Ecological
Monographs
39,
271-96.
Franklin,
J.F.
and
Dyrness,
C.T.
1987:
Natural
vegetation
of
Oregon
and
Washington.
Corvallis,
Ore.:
Oregon
State
University
Press.
Franklin,
J.F.
and
Krueger,
K.W.
1968:
Germination
of
true
fir
and
mountain
hemlock
seed
on
snow.
Journal
of
Forestry
66,
416-17.
Franklin,
J.F.,
Moir,
W.H.,
Douglas,
G.W.
and
Wiberg,
C.
1971:
Invasion
of
subalpine
meadows
by
trees
in
the
Cascade
Range,
Washington
and
Oregon.
Arctic
and
Alpine
Research
3,
215-24.
Franklin,
J.F.,
Swanson,
FJ.,
Harmon,
M.E.,
Perry,
D.A.,
Spies,
T.A.,
Dale,
V.H.,
McKee,
A.,
Ferrell,
W.K.,
Means,
J.E.,
Gregory,
S.V.,
Lattin,
J.D.,
Schowalter,
T.D.
and
Larsen,
D.
1991:
Effects
of
global
climate
change
on
forests
in
northwestern
North
America.
The
Northwest
Environmental
Journal 7,
233-54.
Friedman,
I.,
Carrara,
P.
and
Gleason,
J.
1988:
Isotopic
evidence
of
Holocene
climatic
change
in
the
San
Juan
Mountains,
Colorado.
Quaternary
Research
30,
350-53.
Gorchakovsky,
P.L.
and
Shiyatov,
S.G.
1978:
The
upper
forest
limit
in
the
mountains
of the
boreal
zone
of
the
USSR.
Arctic
and
Alpine
Research
10,
349-63.
Grace,
J.
1989:
Tree
lines.
Philosophical
Transactions
of
the
Royal
Society
of
London
B
324,
233-45.
Griggs,
R.F.
1937:
Timberlines
as
indicators
of
climatic
trends.
Science
85,
251-55.
1938:
Timberlines
in
the
Northern
Rocky
Mountains,
Ecology
19,
548-64.
— 1946:
The
timberlines
of
Northern
America
and
their
inter-
pretation.
Ecology
27,
275-89.
Hadley,
J.L.
and
Smith,
W.K.
1983:
Influence
of
wind
exposure
on
needle
desiccation
and
mortality
for
timberline
conifers
in
Wyoming,
U.S.A.
Arctic
and
Alpine
Research
15,
127-35.
—
1986:
Wind
effects
on
needles
of
timberline
conifers:
seasonal
influence
on
mortality.
Ecology
67, 12-19.
Hansen,
J.
and
Lebedeff,
S.
1988:
Global
surface
air
temperatures:
update
through
1987.
Journal
of
Geophysical
Research
Letters
15,
323-26.
Hansen-Bristow,
KJ.,
Ives,
J.D.
and
Wilson,
J.P.
1988:
Climatic
variability
and
tree
response
within
the
forest-alpine
tundra
ecotone.
at UNIV ARIZONA LIBRARY on September 4, 2008 http://hol.sagepub.comDownloaded from
99
Annals
of
the
Association
of American
Geographers
78,
505-19.
Heikkinen,
O.
1984:
Forest
expansion
in
the
subalpine
zone
during
the
past
hundred
years,
Mount
Baker,
Washington,
U.S.A.
Erdkunde
38, 194-202.
Hehns,
J.A.
1987:
Invasion
of
Pinus
contorta
var.
murrayana
(Pina-
ceae)
into
mountain
meadows
at
Yosemite
National
Park,
California.
Madroño
34,
91-97.
Henderson,
J.A.
1974:
Composition,
distribution,
and
succession
of
subalpine
meadows
of
Mount
Rainier
National
Park.
Ph.D.
thesis,
Oregon
State
University,
Corvallis.
Holtmeier,
F.
1973:
Geoecological
aspects
of
timberlines
in
northern
and
central
Europe.
Arctic
and
Alpine
Research
5,
A45-A54.
-1987:
Beobachtungen
und
Untersuchengen
über
den
Ausaper-
ungsverlaf
und
einige
Folgewirkungen
in
’ribbon-forests’
an
der
oberen
Waldgrenze
in
der
Front
Range,
Colorado.
Phytocoenologia
15, 373-96.
- 1989:
The
’Parkland’
in
America’s
high-mountain
ranges.
Ger-
man
Research:
Reports
of
the
Deutsche
Forschungsgemeinshaft
1/89,
4-7.
Houghton,
R.A.
and
Woodwell,
G.M.
1989:
Global
climate
change.
Scientific
American
260,
36-44.
Hustich,
I.
1983:
Tree-line
and
tree
growth
studies
during
50
years:
some
subjective
observations.
In
Morriset,
P.
and
Payette,
S.,
editors,
Tree-line
ecology,
Nordicana
47,
University
of
Laval,
Quebec,
181-88.
Innes,
J.L.
1991.
High-altitude
and
high-latitude
tree
growth
in
relation
to
past,
present
and
future
global
climate
change.
The
Holocene
1,
168-74.
Ives,
J.D.
and
Hansen-Bristow,
KJ.
1983:
Stability
and
instability
of
natural
and
modified
upper
timberline
landscapes
in
the
Colorado
Rocky
Mountains.
Mountain
Research
and
Development
3, 149-55.
Johnson,
E.A.
1987:
The
relative
importance
of
snow
avalanche
disturbance
and
thinning
on
canopy
plant
populations.
Ecology
68,
43-53.
Jones,
P.D.,
Wigley,
T.M.L.
and
Wright,
P.B.
1986:
Global
tem-
perature
variation
between
1861-1984.
Nature
322,
430-34.
Kearney,
M.S.
1982:
Recent
seedling
establishment
at
timberline
in
Jasper
National
Park,
Alberta.
Canadian
Journal
of
Botany
60,
2283-87.
Kearney,
M.S.
and
Luckman,
B.H.
1983:
Holocene
timberline
fluctu-
ations
in
Jasper
National
Park,
Alberta.
Science
221,
261-63.
Kellogg,
W.
and
Zhao,
Z.C.
1988:
Sensitivity
of
soil
moisture
to
doubling
of
carbon
dioxide
in
climate
model
experiments.
Part
I:
North
America.
Journal
of
Climate
1,
348-66.
Klikoff,
L.G.
1965:
Microenvironmental
influence
on
vegetational
pattern
near
timberline
in
the
central
Sierra
Nevada.
Ecological
Monographs
35,
187-211.
Kramer,
PJ.
and
Kozlowski,
T.T.
1979:
Physiology
of
woody
plants.
Orlando,
Fl.:
Academic
Press.
Kullman,
L.
1986:
Late
Holocene
reproductional
patterns
of
Pinus
sylvestris
and
Picea
abies
at
the
forest
limit in
Central
Sweden.
Canadian
Journal
of
Botany
64,
1682-90.
1991:
Structural
change
in
a
subalpine
birch
woodland
in
North
Sweden
during
the
past
century.
Journal
of
Biogeography
18, 53-62.
Kuramoto,
R.T.
and
Bliss,
L.C.
1970:
Ecology
of
subalpine
meadows
in
the
Olympic
Mountains,
Washington.
Ecological
Monographs
40,
317-45.
LaMarche,
V.C.
1973:
Holocene
climatic
variations
inferred
from
treeline
fluctuations
in
the
White
Mountains,
California.
Quaternary
Research
3,
632-60.
LaMarche,
V.C.
and
Mooney,
H.A.1972:
Recent
climatic
change
and
development
of
the
bristlecone
pine
(P.
longaeva
Bailey)
krummholz
zone,
Mt.
Washington,
Nevada.
Arctic
and
Alpine
Research
4,
61-72.
Lamb,
H.H.
1982:
Climate,
history,
and
the
modern
world.
London:
Methuen.
Lauer,
W.
1978:
Timberline
studies
in
central
Mexico.
Arctic
and
Alpine
Research
10,
383-96.
Lindsay,
J.H.
1971:
Annual
cycle
of
leaf
water
potential
in
Picea
engelmannii
and
Abies
lasiocarpa
at
timberline
in
Wyoming.
Arctic
and
Alpine
Research
3, 131-38.
Little,
R.L.
and
Peterson,
D.L.
1991:
Effects
of
climate
on
regenera-
tion
of
subalpine
forests
after
wildfire.
The
Northwest
Environmental
Journal 7,
355-57.
Love,
D.
1970:
Subarctic
and
subalpine:
where
and what?
Arctic
and
Alpine
Research
2,
63-73.
Lowery,
R.F.
1972:
Ecology
of
subalpine
zone
tree
clumps
in
the
North
Cascade
Mountains
of
Washington.
Ph.D.
thesis,
University
of
Washington,
Seattle.
Luckman,
B.H.
1988:
8000
year
old
wood
from
the
Athabasca
Glacier,
Alberta.
Canadian
Journal
of
Earth
Science
25, 148-5 1.
Luckman,
B.H.
and
Kearney,
M.S.
1986:
Reconstruction
of
Holocene
changes
in
alpine
vegetation
and
climate
in
the
Maligne
range,
Jasper
National
Park,
Alberta.
Quaternary
Research
26,
244-61.
MacDonald,
G.M.
1989:
Postglacial
palaeoecology
of the
subalpine
forest-grassland
ecotone
of
Southwestern
Alberta:
new
insights
on
vegetation
and
climate
change
in
the
Canadian
Rocky
Mountains
and
adjacent
foothills.
Palaeogeography,
Palaeoclimatology,
Palaeoecol-
ogy
73, 155-73.
Maher,
LJ.
1972:
Absolute
pollen
diagram
of
Redrock
Lake,
Boul-
der
County,
Colorado.
Quaternary
Research
2,
531-53.
Markgraf,
V.
and
Scott,
L.
1981:
Lower
timberlines
in
central
Colorado
during
the
past
15,000
yr.
Geology
9,
231-34.
Marr,
J.W.
1977:
The
development
and
movement
of
tree
islands
near
the
upper
limit
of
tree
growth
in
the
southern
Rocky
Mountains.
Ecology
58, 1159-64.
Minnich,
R.A.
1984:
Snow
drifting
and
timberline
dynamics
on
Mount
San
Gorgonio,
California,
U.S.A.
Arctic
and
Alpine
Research
16,395-412.
Mitchell,
J.F.B.,
Manabe,
S.,
Meleshko,
V.
and
Tokioka,
T.
1990:
Equilibrium
climate
change -
and
its
implications
for
the
future.
In
Houghton,
J.T.,
Jenkins,
G.J.,
Ephraums,
J.J.,
editors,
Climate
change:
the
IPCC
scientific
assessment,
Cambridge:
Cambridge
Uni-
versity
Press,
131-64.
Moore,
M.M.
1991:
Tree
encroachment
on
meadows
of the
North
Rim
of
Grand
Canyon
National
Park:
literature
review.
Report
to
the
National
Park
Service,
CA
8000-8-0002,
School
of
Forestry,
Northern
Arizona
University,
Flagstaff,
Arizona.
Munn,
L.C.,
Buchanan,
B.A.
and
Nieson,
G.A.
1978:
Soil
tem-
peratures
in
adjacent
high
elevation
forests
and
meadows
of
Mon-
tana.
Soil
Science
Society
of
America
Journal
42,
982-83.
Oliver,
C.D.,
Adams,
A.B.
and
Zasoski,
RJ.
1985:
Disturbance
patterns
and
forest
development
in
a
recently
deglaciated
valley
in
the
northwestern
Cascade
Range
of
Washington,
U.S.A.
Canadian
Journal
of
Forest
Research
15, 221-32.
Payette,
S.
1983:
The
forest
tundra
and
present
tree-lines
of
the
northern
Quebec-Labrador
Peninsula.
In
Morisset,
P.
and
Payette,
S.,
editors,
Tree-line
ecology,
Nordicana
47,
University
of
Laval,
Quebec, 3-23.
Payette,
S.
and
Gagnon,
R.
1979:
Tree-line
dynamics
in
Ungava
peninsula,
northern
Quebec.
Holarctic
Ecology
2,
239-48.
Payette,
S.,
Filion,
L.,
Delwaide,
A.
and
Begin,
C.
1989:
Reconstruc-
tion
of
tree-line
vegetation
response
to
long-term
climate
change.
Nature
341,
429-32.
Pears,
N.V.
1968:
Post-glacial
tree-lines
of
the
Cairngorm
Mountains,
Scotland.
Transactions
of
the
Botanical
Society
of
Edinburgh
40,
361-94.
Peet,
R.K.
1981.
Forest
vegetation
of
the
Colorado
Front
Range:
composition
and
dynamics.
Vegetatio
45,
3-75.
Peters,
R.L.
1990:
Effects
of
global
warming
on
forest
ecology
and
management.
Forest
Ecology
and
Management
35,13-33.
Peterson,
D.L.
1991:
Sensitivity
of
subalpine
forests
in
the
Pacific
Northwest
to
global
climate
change.
The
Northwest
Environmental
Journal
7,
349-50.
Petersen,
K.L.
and
Mehringer,
PJ.
1976:
Postglacial
timberline
fluctuations,
La
Plata
Mountains,
Colorado.
Arctic
and
Alpine
Re-
search
8,
275-88.
Porter,
S.C.
1981:
Glaciological
evidence
of
Holocene
climatic
change.
In
Wigley,
T.M.L.,
Ingram,
M.J.
and
Farmer,
G.,
editors,
Climate
and
History,
Cambridge:
Cambridge
University
Press,
82-110.
Reasoner,
M.A.
and
Hickman,
M.
1989:
Late
Quaternary
environ-
mental
change
in
the
Lake
O’Hara
region,
Yoho
National
Park,
British
Columbia.
Palaeogeography,
Palaeoclimatology,
Palaeoecol-
ogy
72, 291-316.
Rochefort,
R.M.
and
Peterson,
D.L.
1991:
Tree
establishment
in
subalpine
meadows
of
Mount
Rainier
National
Park.
The
Northwest
Environmental Journal
7,
354-55.
Romme,
W.H.
and
Turner,
M.G.
1991:
Implications
of
global
climate
at UNIV ARIZONA LIBRARY on September 4, 2008 http://hol.sagepub.comDownloaded from
100
change
for
biogeographic
patterns
in
the
Greater
Yellowstone
Eco-
system.
Conservation
Biology
5,
373-86.
Sawyer,
D.A.
and
Kinraide,
T.B.
1980:
The
forest
vegetation
at
higher
altitudes
in
the
Chiricahua
Mountains,
Arizona.
The
American
Midland
Naturalist
104,
224-41.
Schneider,
S.H.
1989:
Global
warming.
San
Francisco:
Sierra
Club
Books.
Schlesinger,
M.E.
and
Mitchell,
J.F.B.
1987.
Climate
model
simula-
tions
of the
equilibrium
climatic
response
to
increased
carbon
dioxide.
Reviews
of
Geophysics
25:
760-98.
Schroeder,
MJ.
and
Buck,
C.C.
1970:
Fire
weather:
a
guide
for
application
of
meteorological
information
to
forest
fire
control
opera-
tions.
United
States
Department
of
Agriculture,
Forest
Service,
Washington
DC.
Scuderi,
L.A.
1987.
Late-Holocene
upper
timberline
variation
in
the
southern
Sierra
Nevada.
Nature
325,
242-44.
Shankman,
D.
and
Daly,
C.
1988:
Forest
regeneration
above
tree
limit
depressed
by
fire
in
the
Colorado
Front
Range.
Bulletin
of
the
Torrey
Botanical
Club
115, 272-79.
Short,
S.K.
1984:
Holocene
vegetational
history
of the
Colorado
Front
Range
(absract).
Sixth
Annual
Palynology
Conference,
Cal-
gary,
Alberta,
Canada,
26
August-1
September.
Silver,
C.S.
and
DeFries,
R.S.
1990:
One
earth,
one
future:
our
changing
global
environment.
Washington,
D.C.:
National
Academy
Press.
Simms,
H.R.
1967.
On
the
ecology
of
Herpotrichia
nigra.
Mycologia
59, 902-09.
Sprugel,
D.G.
1991:
Disturbance,
equilibrium,
and
environmental
variability:
what
is
’natural’
vegetation
in
a
changing
environment?
Biological
Conservation
58, 1-18.
Stevens,
G.D.
and
Fox,
J.F.
1991:
The
causes
of
treeline.
Annual
Review
of
Ecology
and
Systematics
22, 177-91.
Taylor,
A.H.
1990:
Tree
invasion
in
meadows
of
Lassen
Volcanic
National
Park,
California.
Professional
Geographer
42, 457-70.
Tranquillini,
W.
1979:
Physiological
ecology
of
the
alpine
timberline.
New
York:
Springer-Verlag.
Tewartha,
G.T.
1980:
An
introduction
to
climate.
New
York:
McGraw-Hill.
Troll,
C.
1973:
The
upper
timberlines
in
different
climatic
zones.
Arctic
and
Alpine
Research
5,
A3-A18.
Vale,
T.R.
1981:
Tree
invasion
of
montane
meadows
in
Oregon.
The
American
Midland
Naturalist
105,
61-69.
Vankat,
J.L.
and
Major,
J.
1978:
Vegetation
changes
in
Sequoia
National
Park,
California.
Journal
of
Biogeography
5,
377-402.
Wardle,
P.
1968:
Engelmann
spruce
(Picea
engelmannii
Engel.)
at
its
upper
limits
on
the
Front
Range,
Colorado.
Ecology
49,
483-95.
—
1971:
An
explanation
for
alpine
timberline.
New
Zealand
Journal
of
Botany
9,
371-402.
— 1974:
Alpine
timberlines.
In
Ives,
J.D.
and
Barry,
R.G.,
editors,
Arctic
and
alpine
environments,
London:
Methuen,
371-402.
-
1977:
Japanese
timberlines
and
some
geographic
comparisons.
Arctic
and
Alpine
Research
9,
249-58.
Winter,
M.H.
1984:
Altitudinal
fluctuations
of
upper
treeline
at
two
sites
in
the
Lemhi
Range,
Idaho.
MA
thesis,
University
of
Kansas,
Lawrence,
Kansas,
USA.
Woodward,
A.,
Gracz,
M.B.
and
Schreiner,
E.G.
1991:
Climatic
effects
on
establishment
of
subalpine
fir
(Abies
lasiocarpa)
in
mea-
dows
of the
Olympic
Mountains.
The
Northwest
Environmental
Journal
7,
353-54.
Zimina,
R.P.
1973:
Upper
forest
boundary
and
the
subalpine
belt
in
the
mountains
of the
southern
USSR
and
adjacent
countries.
Arctic
and
Alpine
Research
5,
A29-A32.
at UNIV ARIZONA LIBRARY on September 4, 2008 http://hol.sagepub.comDownloaded from
... A logical definition would be the abovedescribed transition zone between the upper limit of the closed montane forest and the tree species line (i.e. the beginning of the treeless alpine zone), but not everybody might agree on that. Subalpine parkland is another plausible term for this transition zone (Rochefort et al. 1994). In the Swiss Alps, closed forests, several hundreds of meters below the treeline, have been termed subalpine (as rated by the presence of Pinus cembra), which would correspond to what is called uppermost montane forest in other parts of the world. ...
... Troll 1973;Wardle 1974Wardle , 1993Arno 1984;Holtmeier 2003). Historical trends were reviewed by Rochefort et al. (1994). For tropical and subtropical mountains, reviews were published by Ohsawa (1990), Miehe and Miehe (1994), and Leuschner (1996). ...
... Evidence for this comes from fossil tree stumps, pollen records, and reconstructions of temperatures from sediments in montane lakes (Chironomid-thermometer), suggesting that temperate zone treelines varied in elevation by no more than 200 m while temperatures varied by several K (e.g. LaMarche and Mooney 1972;Rochefort et al. 1994;Graumlich and Brubaker 1995;Heiri et al. 2003;Lotter et al. 2006; see the summary in Körner 2012a, b). ...
Chapter
The distribution of trees reaches a natural limit somewhere along thermal, drought, or disturbance gradients. The limit set by low temperature at high elevation or high polar latitudes is termed ‘treeline’ (alpine or arctic treeline). Treelines represent an abrupt change in land cover by a dominant life form, a change from tall woody to small, mostly herbaceous or graminoid forms, and the treeline is defined irrespective of the species of trees that reach it. Beyond the treeline, tall, single-stemmed, woody plants with crowns protruding into the free atmosphere either can not establish or be sustained (Fig. 7.1). Why do trees disappear above a certain elevation? What causes the alpine life zone to be treeless? The answer to this question would also indicate which functional attributes alpine plants must have to thrive where trees are unable to exist. Thus, there is a reciprocal interest in this question, upslope for forest ecology, and downslope (because of its lower boundary) for alpine ecology.
... Por ejemplo, variables de forma de los árboles (altura, tamaño y forma de la copa, número de pies por cepa, etc.) pueden formar fronteras distintas de las delineadas por datos de densidad de árboles o de formas arbustivas o por las creadas por datos de edad. Sin embargo, todos estos aspectos muestran perpespectivas parciales y relacionadas de la dinámica del límite forestal (Kullman, 1979;Payette, Filion, 1985;Payette, Lavoie, 1994;Rochefort et al., 1994;Lloyd, 1996). ...
Article
We describe the spatial pattern of a subalpine forest-alpine pasture ecotone in the Central Pyrenees, that includes altitudinal timberline and treeline, and it is dominated by Pinus uncinata Ram. A rectangular (30 x 140 m) plot was located crossing the ecotone with its longest side parallel to the slope. We measured for each P. uncinata individual inside the plot: location (coordinates x, y), and structure (e. g. height) and growth form variables (number and type —living or dead, vertical or shrubby— of stems per individual). P. uncinata individuals were classified according to their size (adults, poles, saplings and seedlings) and growth form (krummholz —shrubby and multistemmed individuals— and krummholz with vertical stems). We described quantitatively the type of substrate (bare soil, organic matter, gravel and rock) and cover of herbs, shrubs and P. uncinata using transects parallel to the slope. The ecotone structure was described through: (1) point pattern (Ripley’s K) and (2) surface pattern analyses (spatial correlograms of height); (3) the detection and description of boundaries using density, size or growth form variables; (4) the synthesis of variations of presence and diversity of substrates and herbs and shrubs; and (5) the ordination of quadrats (the plot was previously subdivided into 115 6 x 6 m quadrats), according to their spatial position in the ecotone, the type of substrate, the cover of herbs and shrubs and the number, size and growth form of P. uncinata individuals. Most P. uncinata living individuals were krummholz, located above the timberline. Krummholz individuals showed significant and positive spatial interaction with seedlings. Bigger, vertical and unistemmed individuals predominated in the lower area of the ecotone, in the forest. The change of height with increasing elevation was abrupt and masked an underlying pattern of patches of trees with similar height in the forest. The structure variables were more sensitive because they produced a greater number of boundaries. These boundaries were arranged forming a “diagonal” (in the lower and upper areas of the ecotone for big and small individuals, respectively) because of the sequential location of progressively bigger unistemmed individuals descending across the ecotone. The shrubby individuals were associated with rocky substrates, that decreased in the forest, where organic matter predominated. The snow-wind interaction can explain the location of the studied timberline that could be considered a local phenomenon. Krummholz can buffer seedlings against the harsh climatic conditions of this ecotone (strong wind, reduced snowpack, low temperature). The spatial location of the different classes of individuals, the spatial interaction between seedlings and krummholz individuals, and changes of growth form (from shrubby to vertical growth form or vice versa) can cause some inertia in the response of ecotone P. uncinata populations to environmental changes.
... A reduction in snowpack also reduces the amount of insulation value of snow and decreases the temperature in the subnivum, causing prey to suffer higher mortality rates. In addition, many landscape-scale ecological processes are driven by the amount of snow in the subalpine zone including the development of alpine meadows and subalpine parklands (Henderson 1973, Rochefort et al. 1994. Persistent snow cover during a significant portion of the year contributes to the development of the habitat conditions that support many of the prey species on which the fox relies, whereas climatic warming results in meadows being invaded by conifer species (Millar et al. 2004). ...
Technical Report
Full-text available
The North Cascades Ecosystem is one of the largest and most intact wilderness areas in the contiguous United States. It spans 34,965 km2 across the U.S.-Canada border between central Washington State and southern British Columbia and is bisected north to south by the Cascade Mountain range. The North Cascades National Park Service Complex (hereafter, the Park) lies in the heart of the ecosystem and is comprised of North Cascades National Park (2,044 km2), Ross Lake National Recreation Area (NRA) (473 km2), and Lake Chelan NRA (251 km2). This report focuses on terrestrial wildlife species in the Park that are federally listed under the Endangered Species Act of 1973 (ESA), state-listed under Washington Administrative Code, and/or designated by NPS as Management Priority species. We provide a detailed synthesis of information around NPS records from 1995–2020 for nine bird and nine mammal species and one mammal Order (Chiroptera [bats]), as well as a suite of invertebrate pollinators in the Park where data are available. Information for each species includes life history information, occurrence in the Park, protective status, trends when known, a summary of known threats, and a summary of conservation and research needs. We also provide brief summaries for an additional five bird and two mammal species, and one taxa group (woodpeckers) where data are more limited.
... Previous studies suggested that tree growth at alpine regions was cold-limited, resulting in a positive correlation between tree-ring width and temperature (Briffa et al., 1998;Körner, 2004;Rochefort et al., 1994). However, Guo et al. (2022) found that radial growth of 13 treeline sites over the Tibatan Plateau were predominantly explained by CO 2 rather by climate change over the last 30 years. ...
Article
Full-text available
Tree growth at alpine regions is sensitive to climate change. However, there is still uncertainty about the spatial and temporal stability of tree growth in response to warming. Herein, through collecting 302 cores from 162 conifer trees at 4 sites of paired timberline and subalpine forests on the southeastern Tibetan Plateau, we analyzed the spatial and temporal pattern of radial growth in response to warming. Response analysis result of standard and first-order difference chronology showed that the variation of radial growth was mainly affected by summer temperature in the Sergyemla Mountains (SGM), Baima Snow Mountains (BSM), and Meri Snow Mountains (MSM), by winter temperature and early summer drought stress at Chagyab (CY). The radial growth in timberline forests is more sensitive to temperature than those in subalpine forests. Growth-temperature sensitivity generally decreased during the past 60 years. The analysis of the Standardized Precipitation Evapotranspiration Index (SPEI) indicated that the drought stress was one of the main reasons for the variation of temperature insensitivity, suggesting that the moisture condition may play an increasingly important role on tree growth in warming future.
... A cooling since the mid-Holocene has also been widely documented in a number of palaeoceanographic and palaeoclimatic studies outside the circum-Iceland region around the North Atlantic region, e.g. by lower sea-surface temperature (Calvo et al. 2002), lighter ice-sheet isotopes (Dahl-Jensen et al. 1998), re-advances of glaciers in Scandinavia (Dahl & Nesje 1996), change of vegetation cover (Heikkik¨a & Sepp¨a 2003), decline in mountain tree limits (Rochefort et al. 1994) and expansion of permafrost areas (Payette et al. 2002). The mid-Holocene climatic reversal has been discussed by, for example, Sandweiss et al. (1999), Steig (1999), Porter (2000), Hodell et al. (2001), Mullins & Halfman (2001), Magny & Haas (2004) and Blaauw et al. (2004) as a probable global event at c. 6000-5000 cal. ...
... For example, highresolution SCA observations, when employed in a data-assimilation context, may improve runoff projections and similar hydrologic simulations by capturing fine-scale variability in snow cover (Clark et al., 2011;Luce et al., 1999;Lundquist and Dettinger, 2005). In mountain ecosystems, the composition of plant communities and associated phenological events such as flowering and growth vary as a function of snow cover extent and snow duration (Choler, 2005;Ford et al., 2013;Theobald et al., 2017;Venn et al., 2011), which exhibit significant heterogeneity over small spatial and temporal scales (1-10 m; hours to days during ablation) (Clark et al., 2011;Little et al., 1994;Rochefort et al., 1994). ...
Article
Full-text available
Snow cover affects a diverse array of physical, ecological, and societal systems. As such, the development of optical remote sensing techniques to measure snow-covered area (SCA) has enabled progress in a wide variety of research domains. However, in many cases, the spatial and temporal resolutions of currently available remotely sensed SCA products are insufficient to capture SCA evolution at spatial and temporal resolutions relevant to the study of fine-scale spatially heterogeneous phenomena. We developed a convolutional neural network-based method to identify snow covered area using the ~3 m, 4-band PlanetScope optical satellite image dataset with ~daily, near-global coverage. By comparing our model performance to snow extent derived from high-resolution airborne lidar differential depth measurements and satellite platforms in two North American sites (Sierra Nevada, CA, USA and Rocky Mountains, CO, USA), we show that these emerging image archives have great potential to accurately observe snow-covered area at high spatial and temporal resolutions despite limited radiometric bandwidth and band placement. We achieve average snow classification F-Scores of 0.73 in our training basin and 0.67 in a climatically-distinct out-of-sample basin, suggesting opportunities for model transferability. We also evaluate the performance of these data in forested regions, suggesting avenues for further research. The unparalleled spatial and temporal coverage of CubeSat imagery offers an excellent opportunity for satellite remote sensing of snow, with real implications for ecological and water resource applications.
... Subalpine forests are limited by cold temperatures and a short growing season, therefore warmer temperatures and more atmospheric CO 2 may increase tree growth, productivity, and potentially carbon sequestration rates of some species at high elevations (Case and Peterson, 2007;Latta et al., 2010;Peterson and Peterson, 2001, Fig. 4). Climate change would also reduce snowpack depth and increase soil temperatures, two limiting factors for growth at high elevations (Ettl and Peterson, 1995;Peterson et al., 2002;Peterson, 2001, 1994;Rochefort et al., 1994), especially for encroaching lower elevation tree species (Franklin et al., 1971;Harsch et al., 2009). However, snowmelt during the dry, summer months is critical for tree growth and seedling establishment at high elevations (Burns and Honkala, 1990) and warming temperatures will intensify summer drought conditions, especially during extreme years (Marshall et al., 2019a,b;Vose et al., 2016). ...
Article
Rising greenhouse gases are changing the Earth’s climate and adversely affecting ecosystems that currently provide a suite of invaluable benefits, from cleaning water to sequestering carbon. Some of the world’s most productive forests grow in the Pacific Northwest region of North America, but our understanding of climate change effects on these forests and their carbon is still emerging. Here, we synthesize the current state of research (including empirical, paleo, and modeling studies), discuss the implications on forest growth and carbon storage in Pacific Northwest forests, and identify key knowledge gaps and future research opportunities based on a combination of published studies and expert opinion. Two case studies are presented that illustrate the expected effects of climate change on moist and dry forest ecology and carbon storage. In response to these impacts, we highlight a number of appropriate regional forest restoration and management adaptation strategies. Filling in knowledge gaps will improve the accuracy of forest carbon accounting, a crucial part of the strategy to meet climate mitigation targets and prevent the most severe impacts of climate change.
Article
Natural climatic treelines are relatively discrete boundaries in the landscape established at a certain elevation within an otherwise continuous gradient of environmental change. By studying tree rings along elevational transects at and below the upper treeline in the European Alps, we (1) determine whether radial stem growth declines abruptly or gradually, and (2) test climatic influences on trees near treeline by investigating transects for climatically different historical periods. While tree height decreases gradually toward the treeline, there is no such general trend for radial tree growth. We found rather abrupt changes which imply threshold effects of temperature which moved upslope in a wave-like manner as temperatures increased over the past 150 yr. Currently radial tree growth at treeline in the Alps is the same magnitude as at several hundred meters below current treeline. Over short intervals, tree-ring width is more dependent on interannual climatic variability than on altitudinal distance to treeline. We conclude that (1) the elevational response of tree-rings includes a threshold component (a minimal seasonal temperature) and that (2) radial growth is more strongly correlated with year to year variation in climate than with treeline elevation as such. Our data indicate that the current treeline position reflects influences of past climates and not the current climate.
Book
Full-text available
Trees and other woody plants, such as shrubs and lianas, form the principal components in forests and many other ecosystems on our planet. Being among the largest and longest-living organisms, they support an immense share of the Earth’s terrestrial biodiversity, providing food and habitats for innumerable microorganisms, epiphytes, invertebrate, and vertebrate species. Woody species are perfect study objects, giving us a link between the past, present, and future. Woody species have also accompanied our own species throughout its evolution. Even today, billions of people depend on trees and shrubs for fuel, medicine, food, tools, fodder for livestock, shade, and watershed maintenance. Woody species, therefore, have a high scientific, economic, social, cultural, and aesthetic value. However, the future of many trees and shrubs is uncertain. Ten of thousands of species are threatened by overharvesting, non-native pests and diseases, changes in accelerated land use, and climate warming. Many aspects of their biology, ecology, and biogeography are still unexplored or insufficiently understood. These knowledge shortfalls, concerning their genetic diversity, for example, significantly hinder the development of protection strategies and the elaboration of efficient action plans. This book, dedicated to this very diverse group of plants, aims to encourage ongoing research and conservation efforts worldwide.
Article
Full-text available
Invasive limber pine Pinus flexilis gave way to Douglas-fir Pseudotsuga menziesii and lodgepole pine Pinus contorta during the period of major climate-caused invasion around 1895-1915. Invasions into the meadows in 1920-1940 were a result of drier climatic conditions, the possible influence of grazing, and a forest fire. Heavy grazing prevents present-day invasion in at least one meadow. -from Author
Article
The use of analysis of variance and correlation ratios reveal that the three dominant processes observed on talus of southwest Devon Island produce characteristic arrangements in size of surface fragments. The correlation ratios show that on 25 of the 27 surfaces zonal (upslope-downslope) variance accounts for a much higher proportion of size variation than does lateral (cross-slope) variation. This characteristic supports the hypothesis of fall-sorting and reverse fall-sorting of rock fragments as the fundamental mode of talus formation. Analysis of variance indicates that at present 10 of 15 (66%) rockfall taluses have statistically significant zonal size arrangements, compared to 1 of 2 (50%) on talus with basal erosion and only 4 of 10 (40%) on alluvial talus. Basal erosion and meltwater activity may either reinforce or obscure the original downslope arrangement, depending upon the form of the process, its magnitude and frequency. Comparison of similar studies by way of statistical power analysis reveals considerable support for these findings.