Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

Plant Ecology & Diversity

ISSN: 1755-0874 (Print) 1755-1668 (Online) Journal homepage: https://www.tandfonline.com/loi/tped20

Plant functional diversity in tropical Andean


páramos

Fermin Rada, Aura Azócar & Carlos García-Núñez

To cite this article: Fermin Rada, Aura Azócar & Carlos García-Núñez (2019): Plant
functional diversity in tropical Andean páramos, Plant Ecology & Diversity, DOI:
10.1080/17550874.2019.1674396

To link to this article: https://doi.org/10.1080/17550874.2019.1674396

Published online: 17 Oct 2019.

Submit your article to this journal

Article views: 9

View related articles

View Crossmark data

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=tped20
PLANT ECOLOGY & DIVERSITY
https://doi.org/10.1080/17550874.2019.1674396

REVIEW

Plant functional diversity in tropical Andean páramos


Fermin Rada, Aura Azócar and Carlos García-Núñez
Instituto de Ciencias Ambientales y Ecológicas, Facultad de Ciencias, Universidad de Los Andes, Mérida, Venezuela

ABSTRACT ARTICLE HISTORY


Background: Tropical high mountains present extreme daily temperature variations, frequent Received 21 December 2017
high air evaporative demands and seasonal differences in soil water availability. Plants have Accepted 26 September 2019
adapted to these conditions through different avoidance-tolerance mechanisms. This review KEYWORDS
focuses on plant-growth forms and their adaptive strategies. Climate change; freezing
Aims: This integrated review of páramo plant traits aims at contributing to understanding the resistance; gas exchange;
functioning of plant-growth forms and their significance on ecosystem properties under giant rosettes; páramo; plant
environmental climate and land-use changes. growth-forms; plant water
Methods: Plant responses are presented along avoidance-tolerance gradients considering relations; tropical Andes
three main aspects: freezing resistance, water relations and gas exchange characteristics.
Results from 45 herbaceous and 42 woody species along elevational gradients in the
Venezuelan high Andes were analysed.
Results: Leaf supercooling is the common avoidance response of woody plants to night-time
freezing temperatures, while herbaceous plants tolerate frost. Trees and caulescent rosettes
maintain more positive leaf water potentials under water deficit conditions compared to more
tolerant herbaceous species. All plant growth-forms showed strong stomatal control under dry-
season conditions.
Conclusions: Páramo plant growth-forms may be separated according to an avoidance-
tolerance gradient in response to water deficit and low temperature resistance. Woody growth-
forms tend to avoid both freezing and water stress, while herbaceous forms tolerate frost and
resist an unfavourable water status. Grasses and cushion plants are at the tolerant extreme of
the gradient and coincide in that both reach the highest elevations in the páramo. Andean
giant rosettes are freezing avoidant, particularly susceptible to water deficit and the most
vulnerable, of all growth-forms, to changing environmental conditions.

Introduction The páramos, a collective term for the alpine eco-


systems of the humid tropical Andes, occur above the
Tropical high mountains offer environmental condi-
treeline, composed of predominantly low-stature
tions that result in unique ecosystems where plants
vegetation, dominated by different plant growth-
present numerous strategies of adaptation. Although
forms (Figure 1) such as giant rosettes, sclerophyllous
these environments are defined as cold, both low and
shrubs, grasses and forbs; at the highest elevations
high temperatures play a key role in shaping plant
acaulescent rosettes and cushion plants usually
responses. Freezing temperatures may occur any
increase their relative abundance. Páramos occur
night of the year with a higher frequency at higher
between 11° N and 8° S along Ecuador, Colombia
elevations and towards the dry season, while high
and Venezuela, with smaller regions to the north of
incoming radiation during clear sky conditions lead
Peru and into Central America in Panama and Costa
to high temperatures on bare soil and plant surfaces.
Rica (Monasterio 1980; Luteyn 1992). The páramos
Hedberg (1964) has defined high tropical mountains as
hold the highest floristic diversity and the largest
environments where ‘summer’ occurs during
number of endemic species of alpine ecosystems
every day and ‘winter’ at every night. In terms of
worldwide (Luteyn 1999; Sklenář et al. 2014). In
water conditions, all tropical high mountain regions,
Venezuela, páramos are found between the natural
in varying degrees, show some seasonality in precipita-
boundaries of continuous forest at around 3000 m to
tion. In addition, at higher elevations, water may be
maximum elevations near 4900 m a.s.l.
frozen during early morning hours, limiting its absorp-
Ample ecological research has been carried out for
tion by roots. Added to these water conditions, low
decades in the Venezuelan páramos (Llambí and
relative humidity and high soil surface temperature on
Rada 2019). These investigations have included
clear days generate extremely high evaporative
plant adaptive responses to the particular conditions
demands (Ramírez et al. 2015).

CONTACT Fermin Rada frada@ula.ve


© 2019 Botanical Society of Scotland and Taylor & Francis
2 F. RADA ET AL.

Figure 1. Representative species of different plant growth-forms of the Venezuelan páramo from the Páramo de Piedras Blancas at
4200 m. A. Giant rosette-shrub association with Coespeletia spicata and Hypericum laricifolium, respectively. B. The forb Senecio
formosus. C. Cushion plants Azorella julliani and Aciachne pulvinata. D. Acaulescent rosette Acaena cylindristachya. Photographic
credits: A, C, D by F Rada; B by M Araujo.

the páramo presents for organisms and has involved Temperature and water availability along
the study of cold temperature and/or freezing resis- elevational gradients
tance, water relations and gas exchange in a vast
One of the peculiarities of tropical high mountain
number of plant species of different plant-growth
climates is their relatively constant seasonal mean air
forms (Azócar and Rada 2006; Rada 2016).
temperature while daily variations are extreme.
However, an integrated approach to the study of
Differences between the warmest and coldest months
these plant traits has not been made so far.
are below 2 K at the lower limit of the páramos
Therefore, in this review we present an overview of
(3000–3600 m) and reach 2.7 K at higher elevations
the functional diversity of páramo plants from
(4200 m) (Azócar and Rada 2006). On a daily basis,
a plant-growth form perspective. This analysis aids
differences between minimum and maximum air tem-
in our efforts to understand how the functional diver-
peratures may vary from 10 K during the wet to more
sity of páramo plants may influence on ecosystem
than 20 K in the dry season, being larger at higher
properties under environmental change, principally
elevations (Figure 2). However, when absolute mini-
related to climate and/or land-use. First, we briefly
mum and maximum temperatures are compared these
describe two of the most important environmental
differences may exceed 30 K (Table 1). Mean air tem-
drivers, temperature and water availability, along
perature along the gradient from Los Plantíos (2600 m)
elevational gradients in the páramo. Second, we
to Piedras Blancas (4300 m) within the largest páramo
review the available studies on cold temperature
complex in Venezuela (Cordillera de Mérida, period
and freezing resistance, followed by leaf water poten-
2015–2016) decreases at a rate of ca. 5 K km−1.
tial studies and stomatal responses and gas exchange
Minimum temperatures of ca. −10°C have been
characteristics of páramo plants. Finally, we briefly
recorded at the higher elevations (Table 1).
examine the implications climate change may have
Additionally, the occurrence of freezing temperatures
on the functional diversity of páramo plant growth
at higher elevations coincides with the ecological limit
forms and eventually on ecosystem properties.
for the distribution of trees. It is at this boundary, ca.
PLANT ECOLOGY & DIVERSITY 3

Figure 2. Mean maximum (■) and minimum (▲) air temperatures along an elevational gradient (Santo Domingo 2600 m, Los
Plantios 3100 m, Mucubají 3550 m, Trasandina 3900 m, Piedras Blancas 4300 m). Data obtained from April-December 2015 for wet
(left) and January-March for dry (right) seasons. Corresponding regressions for the wet, Tmax = −0.0034x+26.731 (r2 = 0.73) and
Tmin = −0,0061x+25,349 (r2 = 0.97), and dry, Tmax = −0.0024x+26.983 (r2 = 0.88) and Tmin = −0,0061x+18.797 (r2 = 0.90), seasons.
Bars correspond to one standard error of the mean.

3000 m in the Venezuelan Andes, where the typical are close to 20°C (Ramírez et al. 2015). These thermal
formations of the páramo (grasses, caulescent rosette- differences lead to differences in plant physiological
sclerophyllous shrub associations) replace montane functioning when growth-forms that grow close to the
forests (Sarmiento 1986). Frequency and duration of ground (e.g. acaulescent rosettes, cushion plants) are
nights with below zero temperatures increases with compared with taller erect growth forms (e.g. shrubs,
elevation to 80% of the days at elevations above giant rosettes). Tropical alpine plants require adapta-
4000 m (Table 1). Marked seasonal differences in the tion to night-time freezing temperatures and high day-
occurrence of below zero night-time temperatures are time soil and plant surface temperatures throughout
observed at the elevation limit of ca. 3100 m. At this the year.
elevation, freezing air temperatures do not occur in the In the Venezuelan Andes, there is a marked dry
wet season, while they begin to occur during the dry season between December and March, and in some
season. At higher elevations (3500 m), freezing is cases a less intense dry period between June and
observed during both seasons, and occurs every night July. The driest páramos in the country have an
of the year during the dry season (Table 1, Sarmiento average annual rainfall of 650 mm, while the most
1986). humid reach 1800 mm (Sarmiento 1986). Unlike
Absolute maximum air temperatures for the entire temperature, which shows a clear decreasing trend
gradient from Los Plantíos (2600 m) to Piedras Blancas with increasing elevation across the globe, precipi-
(4300 m) range between 20 and 25°C (Table 1). In tation patterns vary according to different factors
contrast to maximum air temperatures, which do not (e.g. latitude, macroslope orientation). In the case of
seem to have a stressful effect on plant life per se, the the gradient from Los Plantios to Piedras Blancas
high incoming solar radiation during the day, espe- (Table 1), there is a clear decrease in precipitation
cially on clear days, causes high temperatures at the with increasing elevation (this is common in many
surface of plants and soils. A steep thermal gradient other areas along the northern tropical Andes)
between the ground and the air is generated during (Sarmiento 1986; Leuschner 2000). Additionally,
daytime hours. At 4200 m, soil temperatures may be low relative air humidity occurring mainly during
above 50°C while air temperatures (1 m above ground) the dry season, determine high water evaporative

Table 1. Mean air (Tmean), absolute minimum (Tminabs) and absolute maximum (Tmaxabs) temperatures (ºC) measured at 50 cm
above soil level, along an elevational gradient in the Venezuelan páramo for the wet (April-December 2015) and dry (January-
March 2016) seasons. Number of days per year with temperatures below 0°C (D < 0ºC), % days below 0°C, mean night frost
duration (NF) during wet (WS) and dry (DS) seasons and precipitation (P, mm).
Elevation % NF
(m a.s.l.) Tmean Tminabs Tmaxabs D < 0ºC WS/DS WS/DS P
Santo Domingo 2600 11.8 1.6 25.4 0 0/0 0/0 1360
Los Plantios 3100 9.7 −5.5 24.6 65 0/60 0/2h25’ 1005
Mucubají 3550 6.7 −7.7 23.0 130 7/100 1h25'/4h35’ 970
Trasandina 3900 5.5 −9.5 22.4 185 30/100 2h15'/8h15’
Piedras Blancas 4300 3.9 −9.2 21.2 270 60/100 2h55'/8h45’ 760
4 F. RADA ET AL.

Figure 3. Main abiotic environmental factors which determine daily and seasonal plant responses in the Andean tropical high
mountains. Upper and lower portions of the figure correspond to wet and dry seasons, respectively. Left- and right-hand sides
correspond to temperature and water availability, respectively. Seasonal differences in cloud formation determine contrasting
incoming/outgoing radiation levels which have a direct effect on daytime/night-time temperatures. Seasonal differences in water
availability and daily differences in air evaporative demand influence ambient/plant water status. Plants are constantly subjected
to night-time low and daytime high temperatures; and to low water availability and high air evaporative demand during the dry
season which result in continuous temperature and water stresses (→). During the wet season, night-time freezing temperature
stress is important only at higher elevations, while in the day both high temperature and water stresses are important under rare
days with clear skies (-->).

demands which affect both soil and leaf surfaces. distribution of plants along elevation gradients (Sakai
These two factors, a marked dry season and high air and Larcher 1987; Woodward 1990). The formation of
evaporative demands, may induce water deficit in ice crystals inside the cells lead to injury through the
plants. The responses of different plant growth rupture of the cell membranes. Plants may survive
forms to these two major factors (cold tempera- freezing temperatures by (1) avoiding tissue freezing
tures; and water conditions, through limitations in through having insulating structures or by a process
the soil due to precipitation seasonality or to high called supercooling, whereby ice nucleation is pre-
air evaporative demands, Figure 3) that constrain vented, or by (2) tolerating frost by forming ice crystals
plant growth are discussed in the following sections. in the intercellular spaces (Levitt 1980; Sakai and
Larcher 1987). In the first case, since plants are poiki-
lothermic organisms, insulating structures are only
Plant responses to cold temperatures and
important in a limited number of plant species.
freezing
Supercooling refers to the ability of plant tissues to
Freezing temperatures have been described as maintain water in a liquid state at temperatures
a determinant factor that influences fitness and below the freezing point. Secondly, freezing-tolerant
PLANT ECOLOGY & DIVERSITY 5

plants resist the formation of ice crystals in their inter- avoidance mechanisms, through protection by insulat-
cellular spaces. The formation of ice nuclei in the ing structures (stem pith and apical bud) and super-
intercellular spaces immediately creates a gradient in cooling in leaf cells. In all cases, ice nucleation
which intracellular water is displaced outwards where corresponded to injury temperatures meaning that
it freezes and thereby prevents the rupture of mem- ice formation produced damage. Mean leaf supercool-
branes and organelles by ice crystals inside the cell. ing capacity for giant rosettes studied along a gradient
This loss of water into the intercellular spaces causes between 3100 m and 4200 m is ca. −9°C, with a low
the cell to undergo a significant dehydration. supercooling capacity at lower elevations (−6°C at
Therefore, a freezing tolerant plant must necessarily 3100 m) and higher at 4200 m (−14°C). Minimum
be tolerant to water deficits (Sakai and Larcher 1987; recorded air temperatures at this latter elevation in the
Larcher 2003; Azócar and Rada 2006). Páramo de Piedras Blancas, where the species of
Early studies on plant freezing resistance mechan- Coespeletia and Espeletia schultzii cohabit, seldom
isms in the páramo were carried out on giant rosettes drop below −5°C (Rada et al. 1985a). Unlike these
of the genera Coespeletia and Espeletia, species with giant rosettes of the Andes, afroalpine rosettes tolerate
pubescent leaves on a single stem with marcescent freezing of their leaves (Beck et al. 1982). Although
leaves (Figure 1). These giant rosettes provide these studies in Africa were carried out at a similar
a typical example of the role of insulating structures elevation as those of the páramos, minimum recorded
in the resistance to freezing temperatures. Hedberg air temperatures frequently fell below −10°C. Under
(1964) first referred to nyctinastic leaf movements in these more unfavourable temperature conditions, the
afroalpine giant rosettes, while Smith (1974) described idea that in less extreme environments avoidance
these movements for Espeletia schultzii in the mechanisms prevail, while in more extreme condi-
Venezuelan páramos. At night, adult rosette leaves tions plant survival is attained mainly through freezing
bend inward, thus protecting the apical bud and slow- tolerance is supported (Sakai and Larcher 1987;
ing down heat loss to the surrounding air. Rada et al. Azócar et al. 1988; Squeo et al. 1991; Rada 2016).
(1985a) have reported apical bud temperatures above Considering other plant growth-forms, some
0°C for Coespeletia timotensis and C. spicata with respond to night-time freezing conditions through
surrounding air temperatures of −4 °C. Likewise, the avoidance by supercooling while others are freezing
marcescent leaves around the stem produce the same tolerant (Figure 4). The tree Polylepis sericea occurs
effect on the stem pith protecting it from low night- well above the continuous forest line (ca. 3100 m) in
time temperatures (Smith 1979; Goldstein and the Venezuelan Andes (Llambí 2015), up to 4200 m
Meinzer 1983; Rada et al. 1985a). Larcher (1975) and elevation and closely associated to rock outcrops
Larcher and Wagner (1976) described supercooling as which act as thermal refuges (see Goldstein et al.
an avoidance mechanism in tropical high mountain 1994 for a further description of the biology of this
plants for the first time in Espeletia semiglobulata species). This species has a moderate supercooling
leaves, reporting an ice nucleation temperature of capacity to ca. −9°C (Rada et al. 1985b). In a further
−12°C. Rada et al. (1985a) have recorded nucleation study, Dulhoste (2010) has shown freezing resis-
temperatures between −14 and −16°C for leaves of tance of 20 tree species growing in an elevational
Coespeletia timotensis and C. spicata; Goldstein et al. gradient from the cloud forest to the páramo at
(1985), studying 11 different species of giant rosettes 3500 m, through supercooling. Similar supercooling
along an elevational gradient, have found that all spe- capacities have been reported by Rehm and Feeley
cies avoided freezing of leaves by different degrees of (2015a) for the Peruvian tropical Andes. Shrubs
supercooling. These authors and Rada et al. (1987) generally avoid freezing (Squeo et al. 1991), but
have reported a direct relationship between supercool- two species, Valeriana parviflora and Monticalia
ing and temperature with increasing elevation, with sclerosa, have been found to be tolerant to freezing
a greater capacity to supercool (lower nucleation tem- (Azócar 2006). On the other hand, forbs, acaules-
peratures) at higher elevations. A relationship between cent rosettes, cushion plants (Azócar et al. 1988;
leaf supercooling and leaf water status has also been Squeo et al. 1991) and herbaceous grasses are freez-
described (Goldstein et al. 1985; Rada et al. 1987). ing tolerant. Márquez et al. (2006) have reported
Supercooling capacity increases as leaf water potentials that all studied grasses along a 2500–4200 m eleva-
decrease, a condition that accentuates during the dry tional gradient were freezing tolerant with injury
season, when minimum night-time temperatures are temperatures down to −18°C. On the contrary,
recorded in the páramo. In summary, all giant rosettes woody grasses (bamboos) in the páramos have
of the Venezuelan Andes studied to date rely solely on been shown to be frost avoidant (Ely et al. 2014,
6 F. RADA ET AL.

Figure 4. Temperature (IT, ■) that causes 50% leaf tissue injury Figure 5. The relationship between injury temperature and
and ice nucleation temperature (INT, □) for all studied plant elevation for all plants (n = 64) studied from different growth-
growth-forms of the páramos. CP, cushion plants (4 species); forms for the Venezuelan páramos (Tinjury = 2*10−6x2 – 0,0175x
Gr, grasses (9); AR, acaulescent rosettes (4); F, forbs (8); Sh tl, + 28,046; r2 = 0,38, P = 0.12). Dotted line represents absolute
frost-tolerant shrubs (2); WG, woody grasses (5); Sh av, frost- minimum temperature -elevation relation (Tmin = −0.0061x
avoiding shrubs (5); CR, caulescent rosettes (10); and Tr, trees +18.797, r2 = 0.90) along a páramo elevational gradient.
(17). Significant differences between mean IT and INT for the Sources: see list in Figure 4.
first five growth-forms listed (CP to Sh tl) can survive ice
formation in leaf tissues, i.e. freezing tolerant. No significant
differences for the remaining four growth-forms indicate that subjected to lower absolute minimum temperatures
ice formation causes leaf tissue injury and are frost avoidant. during establishment. García-Varela and Rada (2003)
Bars correspond to one standard error of the mean. Different
small letters indicate significant differences in IT among plant have studied freezing resistance mechanisms in juve-
growth-forms (One-way Permanova (Pseudo F = 8.54, niles of Coespeletia timotensis and C. spicata, while
P < 0.01). Sources: CP (Squeo et al. 1991; F Rada, E Márquez, Rada et al. (2011) have described freezing resistance
R Dulhoste unpublished data); Gr (Márquez et al. 2006); AR
(Squeo et al. 1991; Alvizu 2004; F Rada, E Márquez, R Dulhoste,
in Polylepis sericea saplings and ramets. Both studies
P Alvizú unpublished data); F (Azócar et al. 1988; Squeo et al. have reported similar avoidance responses as their
1991; Azócar 2006; Rada et al. 2008); Sh tl (Azócar 2006); WG adult counterparts, but survival of saplings was asso-
(Ely et al. 2014); Sh av (Squeo et al. 1991; Alvizu 2004; Azócar ciated to more favourable microsites for establishment,
2006); CR (Goldstein et al. 1985; Rada et al. 1985a, 1987); Tr
(Rada et al. 1985b; Cavieres et al. 2000; Dulhoste 2010). i.e. rocks or ‘nurse’ plants (see Llambí et al. 2013 for
other upper treeline species). Rehm and Feeley (2015a)
have reported that juvenile tree species maintained
2019). In general, with a few exceptions, woody freezing resistance similar to that of their adults at
plants tend to be avoiders, while herbaceous plants a tropical treeline and that rare extreme temperature
are freezing tolerant. Moreover, injury temperatures events might limit the advance of these species into the
are lower in herbs, with cushion plants and grasses open páramo. However, these authors have also
being the most resistant, contributing to explain pointed out that another species, Weinmannia fagar-
why these groups reach the highest elevations oides, has the ability to withstand lower freezing tem-
under natural conditions (Cuesta et al. 2019; peratures, suggesting that extreme low temperatures
Figure 4). These results agree with the hypothesis are not limiting its upper elevational limit.
by Sakai and Larcher (1987) whereby plants under In general, research in the Venezuelan páramos
less extreme temperature conditions, i.e. lower ele- suggest that even though for some species, their
vations, rely on avoidance mechanisms, while plants upper distribution limit may be determined by their
subjected to more extreme conditions tolerate freez- inability to withstand freezing temperatures, there are
ing. Regardless of the mechanism, avoidance or many others which do not seem to be affected. In the
tolerance, there is an increase in resistance with particular case of trees, there are many species
increasing elevation (Figure 5). More importantly, (Dulhoste 2010; Rehm and Feeley 2015a) which clearly
injury temperatures are always below the minimum have the capability to survive freezing night-time tem-
recorded temperatures at the different elevations in peratures but currently occur at lower elevations than
the Venezuelan páramo. the potential temperature limits, e.g. the evident con-
Few studies have been carried out on the freezing tinuous forest line in the cloud forest-páramo ecotone.
resistance of páramo plants in their establishment This suggests that there must be other limiting factors
phases (seedlings or juveniles). As mentioned in the (see below), determining the distribution of trees
previous section, the more extreme temperature con- along elevational gradients in the páramo. Studies
ditions occur at the soil surface. Therefore, plants are have suggested that treelines, on a global scale, are
PLANT ECOLOGY & DIVERSITY 7

associated with a mean root-zone temperature of 6.7°


C, and near 5°C for tropical treelines (Körner and
Paulsen 2004). While seedling or sapling stages may
be critical in the establishment of trees (Hoch and
Körner 2005; Rada et al. 2011), and certain studies
have suggested a temperature related limitation in
carbon investment towards growth instead of
a constraint in the photosynthesis process per se
(Hoch and Körner 2005; Körner 2012); the physiolo-
gical mechanisms that limit the upper tree distribution
is still under debate. Further ecophysiological studies
Figure 6. Mean minimum leaf water potentials (ΨLmin) during
on the genus Polylepis may lead us to the answers, and
wet (■) and dry (□) seasons for the different páramo plant
as Körner (2012) has very accurately suggested, the growth-forms studied. Tr, trees (13 species); CR, caulescent
‘treeline will be understood once we find the func- rosettes (5); Sh, shrubs (6); F, forbs (12); excluded growth-forms,
tional differences between trees and alpine shrubs’. AR, acaulescent rosettes (5); CP, cushion plants (3); Gr, herbaceous
grasses (9). Number of species studied for each plant life-form in
parenthesis. Bars correspond to one standard error of the mean.
Different lower case letters indicate significant differences in dry
Leaf water potentials in páramo plants season ΨLmin between plant growth-forms (One-way Permanova
The water status of a plant is determined by the (Pseudo F = 6.64, P < 0.01). Sources: Tr (Rada et al. 1996; Cavieres
et al. 2000; Dulhoste 2010; F Rada, L Ramírez unpublished data);
balance between water uptake by the roots and water CR (Goldstein et al. 1989; Rada et al. 1998, 2012); Sh (Rada 1993;
loss through transpiration from the leaves. The former Llambí et al. 2003; Pirela 2006); F (Briceño 1992; Llambí et al.
depends on the availability of water in the soil and the 2003; García-Varela 2008); AR (Rada 1993; Pirela 2006;
García-Varela 2008); CP (Cáceres 2011; Ramírez et al. 2015;
latter on air evaporative demand and stomatal control. P Alvizú, F Rada unpublished data); Gr (García-Varela 2008;
The ability of a plant to respond to environmental F Rada, E Márquez, R Dulhoste unpublished data).
water deficit may be determined through its capacity
to survive under low leaf water potentials (ΨL) condi-
tions. Those species that manage to maintain higher demand are adverse. Giant rosettes with the highest
(less negative) ΨL and therefore a more favourable capacitance are found at the highest elevations in the
water status under limiting water conditions rely on driest páramos (Goldstein et al. 1984; Rada 2016).
avoidance mechanisms, while those which survive For example, Coespeletia timotensis, the species with
these adverse conditions reaching lower (more nega- the highest capacitance, grows from 4200 m to
tive) ΨL are more tolerant to water deficit. Woody 4600 m in the subnival páramo of Piedras Blancas.
plants tend to be less tolerant than herbaceous plants Its high capacitance allows it to maintain constantly
when comparing responses of different plant growth- positive leaf ΨL compared to other giant rosettes,
forms along this avoidance-tolerance gradient even in the dry season (Meinzer and Goldstein
(Figure 6). Under wet season conditions no differences 1985). In contrast, Coespeletia moritziana is the
in minimum leaf water potentials (ΨLmin) were found giant rosette most tolerant to water deficit (Rada
between growth-forms, all showing values between et al. 2012). This species grows on poorly developed
−0.9 and −1.3 MPa. This is expected as soil water is soils and fractured rocks, which are associated with
available and air evaporative demand is relatively low low soil water retention. For C. moritziana, ΨLmin
in the wet season, and therefore, water deficit does not during the dry season can reach −1.8 MPa compared
seem to represent a major constraint. Conversely, to the minimum water potentials of other rosettes
during the dry season, under limiting soil water avail- which do not fall below −1.5 MPa (Goldstein et al.
ability and high temperature and air evaporative 1984; Rada et al. 1998). Additionally, unlike other
demand, ΨLmin are lower (−1.4 to −2.7 MPa). During high elevation giant rosettes, C. moritziana shows
this season, trees and giant rosettes show the highest the ability to osmotically adjust from wet to dry
ΨLmin, while grasses show the lowest. season, allowing lower water potentials without tur-
Most giant rosettes have stems with a well- gor loss (Rada et al. 2012). With respect to establish-
developed pith that allows them to have a high water- ing giant rosette seedlings and juveniles, in which
capacitance, i.e. high capacity to store water (Rada pith capacitance does not play an important role in
2016). This pith permits giant rosettes to maintain maintaining a favourable water status, different
a favourable water balance even when conditions of authors have found that the effect of adverse envir-
water availability in the soil or high evaporative onmental conditions at ground level are the main
8 F. RADA ET AL.

cause for their high mortality rates (Smith 1981; For leaves to remain physiologically active, it is
Goldstein et al. 1985; Estrada and Monasterio 1988; necessary that their cells be turgid. The capacity to
Guariguata and Azócar 1988) and attributed it to maintain turgor in plants can be determined by com-
a low physiological capacity to tolerate drought. paring ΨLmin during periods of water deficit with
Additionally, García-Varela (2000) and Mora et al. osmotic potentials at the turgor loss point (ΨLtlp).
(2019) have reported that juveniles of Coespeletia None of the studied growth-forms in the páramo has
timotensis and C. spicata were associated with favour- shown significant differences between these two
able microsites such as those near to adult rosettes or potentials (Figure 7). These results suggest that in
rocks. general, even during periods of peak water deficit,
There are other giant rosettes (e.g. Ruilopezia atro- most plant species maintain turgor. However, at least
purpurea) that differ from those described above by two of the studied species within each growth-form
their morphological characteristics, including the present ΨLmin below ΨLtlp during the dry season, e.g.
absence of leaf pubescence and lack of a well- the cushion plant Aciachne pulvinata (ΨLmin = −2.6
developed stem pith. For R. atropurpurea minimum and ΨLtlp = −2.2 MPa), the grass Vulpia myuros (−3.7
leaf water potentials of −1.5 MPa have been described and −2.2 MPa), the hemiparasite forb Castilleja fissifo-
(Baruch and Smith 1979; Rosquete 2004; Navarro lia (−2.8 and −2.2 MPa), the acaulescent rosette
2013), a value comparable to those for Coespeletias. Calandrinia acaulis (−2.7 and −2.4 MPa) and the
Navarro (2013) has described osmotic adjustments for shrub Monticalia sclerosa (−1.9 and −1.6 MPa) while
adult individuals of this species, which allows them to ΨLmin were always above ΨLtlp in all trees and caules-
withstand water stress during the dry season without cent rosettes (data not presented). Trees and caulescent
relying on a well-developed stem pith. All these rosettes always present more positive ΨLmin and ΨLtlp.
authors have agreed in that this species does not have For trees, this may be explained by their root systems
the necessary adaptations to grow in more extreme that allow them to obtain water from deeper soil layers
environments, being restricted to higher precipitation, and for caulescent rosettes by the high capacitance of
moderate temperatures and relatively nutrient-rich their stem piths. As with herbaceous growth-forms, it
soils found in low elevation páramos near the contin- has been reported that when water deficit becomes
uous forest line (Monasterio and Vuillemuier 1986). extreme during severe dry seasons, aerial biomass is
Trees have been considered as avoiders of water lost, remaining in a dormant state until the following
deficit (Leuschner 2000; Cabrera 2002; Dulhoste wet season when new leaves are produced (Rada 1993).
2010) through different anatomical and functional
characteristics. Dulhoste (2010) has reported mini-
mum water potentials of ca. −1.5 MPa for forest –
páramo ecotone trees, with a minimum of −2.0 MPa
for Diplostephium venezuelense, a small tree able to
colonise páramos above the continuous forest line
(Llambí et al. 2013). Rada et al. (1985b) have reported
minimum water potentials of −2.0 MPa for Polylepis
sericea, distributed between 4200 and 4600 m in the
Venezuelan Andes, growing higher than any other
tree. Important osmotic adjustments have been
described for both tree species, allowing them to main-
tain turgor during periods of water deficit (Rada et al.
1985a; Dulhoste 2010). For shrubs, Hypericum larici-
Figure 7. Leaf water potentials (ΨL), minimum ( ) and at the
folium is the most representative and abundant species turgor loss point ( ), for the different páramo plant growth-forms
growing along a wide elevation range between 2600 m during the dry season. Tr, Trees (13 species); CR, Caulescent
and 4300 m. This species reaches minimum water rosettes (7); Sh, Shrubs (4); F, Forbs (6); excluded life-forms: AR,
Acaulescent rosettes (3); CP, Cushion plants (3); Gr, Grasses (3).
potentials of ca. −1.8 MPa during periods of water Bars correspond to one standard error of the mean. Different
deficit in the dry season (Rada 1993; Llambí et al. lower case letters indicate significant differences in ΨL at the
2003; Pirela 2006). At the other extreme, grasses exhi- turgor loss point between plant growth-forms (One-way
bit the lowest leaf ΨLmin with a mean of −2.8 MPa and Permanova (Pseudo F = 7.80, P < 0.01). Source: Tr (Rada et al.
1985b,1996 ; Dulhoste 2010; R Dulhoste, F Rada, unpublished
values as low as −4.0 MPa for Nassella mexicana or data); CR (Rada et al. 2012; Navarro 2013; F Rada unpublished
Cortaderia hapalotricha (Pirela 2006; García-Varela data); Sh (Alvizu 2004); F (Briceño 1992; Rada 1993; Alvizu2004);
2008; Ramírez et al. 2015). AR (Alvizu 2004); CP (Alvizu 2004); Gr (Alvizu 2004).
PLANT ECOLOGY & DIVERSITY 9

Stomatal control and gas exchange Llambí et al. 2003). The three growth-forms which
show no significant seasonal differences in IWUE
The degree of stomatal opening or closure at the leaf-
are: the giant rosettes, which partly compensate by
air interphase is a key factor which determines both
mobilising water from their large water storage in
the amount of water vapour losses from the leaf
the pith; herbaceous grasses which, as indicated
lamina to the air (transpiration), and of CO2 from
above, lose part of all of their aerial biomass during
the air entering the substomatal cavities of the leaf
the dry season and bamboos which generally live in
and determine the CO2 assimilation rate. Through
constantly humid environments.
stomatal control, plants must optimise the relation-
ship between CO2 assimilation and leaf conductance,
known as intrinsic water use efficiency (IWUE). An
Páramo plants and climate change
unfavourable water status in páramo plants occurs
mostly during the dry season, when a reduction in The specific environmental conditions at high ele-
precipitation leads to reduced soil water availability vations have given rise to the unique alpine eco-
and there is a high air evaporative demand produced systems of the tropical Andes. Variability in
by high incident radiation, high temperature and low climate, topographic, geomorphologic and soil
relative humidity. Such high air evaporative demand characteristics lead to strong spatial heterogeneity
effect can also occur during the wet season, but due to over very short distances and along elevational
the occurrence of fewer days with clear skies, it plays gradients. This close relationship between climate
a less important role. The effect of climatic season- and environment in continuously occurring tran-
ality on water availability is evident when stomatal sitional zones reveal important differences in the
responses of all growth-forms between seasons is structure of vegetation communities (Fariñas and
compared (Table 2). A reduction of 50–60% in sto- Monasterio 1980; Ramírez et al. 2009; Llambí
matal conductance (gs) in all growth forms occurs 2015). Páramos maintain high rates of endemism
between seasons; only bamboos differ, showing due to an ′island′ effect which, through isolation,
a smaller reduction (22%). This may be explained favour speciation (Pouchon et al. 2018). These
by the humid environments with which the bamboos biologically diverse ecosystems (Beniston 2003)
have been associated (Ely 2009; Ely et al. 2019). are considered important biodiversity hotspots
Seasonal stomatal closure is concomitant with (Myers et al. 2000) and have been described as
a reduction in CO2 assimilation in the different highly exposed and sensitive to climate change
plant growth-forms of the páramo. The reductions (Gosling and Bunting 2007; Buytaeart et al. 2010;
in CO2 assimilation rates range from 30–45% for Cuesta et al. 2017).
most growth-forms and up to 65–75% in grasses Plants may respond in three different ways to
and forbs. Many of the species belonging to these changes in climate: adaptation, migration or local
two latter groups usually enter a semi-dormancy extinction (Thuiller et al. 2008). Mountain vegeta-
stage, losing above-ground biomass as the dry sea- tion should migrate upward in response to climate
son progresses. Regarding IWUE, most growth- change; however, Walther et al. (2005) have sug-
forms show an increase towards the dry season gested that vegetation will not move as a ´front´ and
when water becomes a limiting resource (e.g. shifts would occur according to individualistic

Table 2. Mean maximum stomatal conductance (gsmax, mmol m−2s−1), CO2 assimilation (Amax, μmol m−2s−1) and
intrinsic water use efficiency (Amax/gsmax x10−2) during wet (WS) and dry (DS) seasons, for the different plant growth-
forms of the Venezuelan Andean páramo (number of species studied in parenthesis). Tr – Trees, CR – Caulescent
rosettes, Sh – shrubs, AR – Acaulescent rosettes, F – forbs, excluding those herbs categorised in other growth-forms,
Gr – grasses, WG – Woody grasses.
gsmax Amax Amax/gsmax
Growth-form WS DS WS DS WS ES
Tr (7)1 199 ± 54 87 ± 24 9.1 ± 1.9 6.3 ± 0.5 5.5 ± 0.8 7.6 ± 0.9
CR (6)2 115 ± 27 55 ± 6 5.7 ± 1.4 3.7 ± 1.0 5.3 ± 1.2 6.5 ± 1.4
Sh (7)3 120 ± 10 51 ± 7 8.7 ± 1.2 4.9 ± 0.6 4.1 ± 0.6 12.6 ± 2.2
AR (6)4 165 ± 18 68 ± 19 7.4 ± 1.3 4.6 ± 0.7 4.7 ± 1.2 7.3 ± 0.9
F (15)5 156 ± 19 57 ± 15 10.2 ± 1.3 3.5 ± 0.7 5.2 ± 0.6 8.2 ± 2.0
Gr (10)6 139 ± 33 61 ± 8 8.2 ± 0.6 2.3 ± 0.8 7.0 ± 1.5 5.6 ± 2.4
WG (3)7 89 ± 22 69 ± 9 6.3 ± 1.1 4.8 ± 0.8 7.3 ± 0.8 7.1 ± 1.1
Source: 1 Rada et al. (1996), Cavieres et al. (2000), Dulhoste (2010); 2 Goldstein et al. (1989), Rada et al. (1998, 2012), Navarro (2013), F Rada,
R Dulhoste unpublished data; 3 Rada (1993), Pirela (2006), Cáceres, Llambí et al. (2003); 4 Rada (1993), Cabrera et al. (2000), Pirela (2006),
Cáceres and Rada (2011); 5 Cabrera et al. (2000), Pirela (2006), García-Varela (2008), Llambí et al. (2003); 6 Pirela (2006), García-Varela (2008),
F Rada, E Márquez, R Dulhoste unpublished data; 7 Ely (2009).
10 F. RADA ET AL.

responses by species and the disaggregation and when models are compared to observational data.
emergence of other species associations under Different reports suggest a 300 mm/year increase in
environmental change (Williams and Jackson the next hundred years for Ecuador and Southern
2007). On the other hand, Körner (2003) has sug- Colombia, while Northern Colombia and Venezuela
gested that projections that assume that vegetation show a decreasing precipitation trend (Urrutia and
will move up along isotherms overestimate this Vuille 2009; Buytaeart et al. 2011).
shift, while the effect on single species may be There is no clear evidence of the direct effect of
underestimated. Unless micro-habitat conditions daytime higher temperatures on photosynthetic
are considered, climate warming projections may processes of páramo plants. Different authors have
be misleading (Scherrer et al. 2011). Additionally, shown that alpine plants present a particularly wide
Scherrer and Körner (2011) have suggested that, range of temperatures in which photosynthesis
due to the thermal heterogeneity at a micro- remains at or near optimum, suggesting that CO2
habitat scale in alpine habitats there may be more assimilation would not be affected by an increase in
suitable sites available than have been assumed temperature per se (Rada et al. 1992; Rada et al.
under climatic change scenarios. Evidence for 2008; Körner 2012). However, higher temperatures
migration has been described for mountain envir- will have an effect on air evaporative demands and
onments and supported through an increase in spe- soil water availability. Higher evaporative demands,
cies richness at mountain summits as an effect of lower soil water content and projected decreases in
global warming (Grabherr et al. 1994; Pauli et al. precipitation for the páramos suggest an accentu-
1996, 2012). Pauli et al. (1996) and Rehm and Feeley ated effect on plant water relations and gas
(2015b) have reported that migration of vascular exchange. Considering all plant growth-forms stu-
plants has occurred at a lower rate when compared died, shrubs and herbaceous growth-forms tend to
to temperature changes, an evidence of a time lag be more tolerant to water deficit (i.e. include species
between temperature increases and vegetation which present the lowest ΨLmin) than trees (forest –
responses. In the particular case of the cloud forest- páramo ecotone species and Polylepis sericea) and
páramo ecotone in the Andes, Rehm and Feeley caulescent rosettes. In addition, shrubs and herbac-
(2015b) have described a dsiplacement upslope of eous growth-forms seem to have a greater capacity
cloud forest trees three times slower than expected to respond to a broader range of water availability
from concurrent warming. Other factors, e.g. conditions (larger standard errors within these
decreasing water availability, high incoming radia- growth-forms, Figure 6). For example, the shrub
tion, competition with low strata vegetation of the genus Hypericum presents extremely avoidant spe-
open páramo, which may be responsible for this cies, H. juniperinum, restricted to very humid habi-
fairly stable forest treeline need to be studied. tats; and H. laricifolium, a shrub which largely
Therefore, how plants will respond in terms of dominates the páramos along a wide elevational
adaptation or extinction will depend largely on their gradient (2200–4500 m) including humid and
physiological and ecological characteristics. How phy- desert páramos (Fariñas et al. 2008; Cáceres et al.
siological responses change via plasticity or evolution 2015). Therefore, these patterns suggest that even
are determinant in the projections of the effects of though some species within these growth-forms
climate change on vegetation distribution. Cuesta may disappear, others may functionally replace
and Becerra (2012) and Cuesta et al. (2017) have them, giving this ecosystem a certain degree of
suggested that for tropical Andean plant communities, resilience with regard to shrub-herbaceous plant
species with restricted ranges and/or at higher eleva- communities.
tions will show a large reduction in their habitats and Trees, with the exception of Polylepis sericea, are
many of them may be driven towards extinction (see restricted to lower elevations and transition zones
also Mavárez et al. 2019). between upper cloud forests and low páramo
The tropical high Andes stands out as a region of (Ramírez et al. 2009; Arzac et al. 2011, 2019).
high projection uncertainty, limiting the reliability of Considering that all tree species studied to date along
future predictions. Projections suggest a 3 ± 1.5°C forest – páramo gradients have shown avoidance
rise for the Andes in the next century (Buytaeart et al. responses to water deficit (higher ΨLmin, Figure 6),
2011) while Vuille et al. (2003) have reported a rise of under expected scenarios of greater water stress, only
0.34°C per decade in the last 30 years for the northern those species tolerating lower leaf water potentials
Andes. With respect to precipitation, Buytaeart et al. would be fit to colonise páramos at higher elevations.
(2010) have described large errors in simulations Dulhoste (2010) has described Diplostephium
PLANT ECOLOGY & DIVERSITY 11

venezuelense as the woody species which presents the On a tolerance gradient and on a plant growth-
most similar water relations to those of P. sericea, form basis; caulescent rosettes, trees and woody
which grows in small patches under sheltered condi- grasses are less resistant to water deficit, while the
tions above the continuous forest line. Coincidently, two former growth-forms are less resistant to freezing
D. venezuelense was found to be the dominant species temperatures (Figure 8). In general, even though direct
in continuous forest borders in the forest – páramo effects of night-time freezing temperatures are
ecotone, and the only tall woody species present in the a determinant factor in the survival and distribution
adjacent open páramo (Ramírez et al. 2009; Llambí of specific páramo plant species, they seem to play
et al. 2013). However, other factors such as the high a less important role in the composition and vegeta-
radiation loads characteristic of páramos outside the tion dynamics, at a plant growth-form level, towards
continuous forest canopy need to be considered. Bader higher elevations. Additionally, all climate change sce-
et al. (2007) have reported that excess solar radiation narios indicate increasing temperatures and an
limits tree establishment above the continuous forest ascending frost line towards higher elevations (Vuille
line, with Diplostephium species showing the highest et al. 2003; Braun and Bezada 2013; IPCC 2013;
survival when plant cover was experimentally Rabatel et al. 2013) which would favour those less
removed from saplings growing in páramos adjacent resistant species. However, it should be kept in mind
to the continuous forest line. that even though mean temperatures are on the rise,
The giant rosettes may be considered the most extreme temperature events, including lower mini-
dominant growth-form of the páramo. Nevertheless, mum freezing temperatures are occurring (Inouye
most studied species of this group, with the exception 2008; Sierra-Almeida and Cavieres 2010). High day-
of Coespeletia moritziana, maintain relatively high ΨL time temperatures seem to be more important in the
(> −1.5 MPa) in order to survive. Presently, in order to case of plants growing at ground level, i.e. acaulescent
maintain high ΨL, giant rosettes maintain nearly closed rosettes, prostrate herbs, cushion plants or establishing
stomates during water deficit periods which, in turn, seedlings, where extremely high temperatures fre-
limit processes such as CO2 assimilation. Under quently occur. However, indirect high temperature
a scenario of severe lack of water, giant rosettes effects may be determining survival and distribution
would not have the capacity to control stomates of plant growth-forms in the páramos (Figure 8).
further which will affect survival. Consequently, this These effects are mainly related to higher ambient
growth-form seems to be the most susceptible to chan- water evaporative demands, which would have
ging environmental conditions. Azócar et al. (2000) a direct effect on plant gas exchange and water rela-
have suggested that giant rosettes in high elevation tions, e.g. limiting stomatal opening and affecting CO2
ecosystems may display low resilience and a fragile assimilation rates. On a species-specific basis, some
functional stability. Additionally, habitat loss due to plants may be restricted by other factors, such as
climate change in giant rosettes is expected to be more high radiation inputs including UV radiation, nutrient
severe on species living at higher elevations and found relations, plant-plant and plant-pollinator interac-
in restricted habitats (Mavárez et al. 2019), i.e. endemic tions, seed dispersal mechanisms, which have an effect
species such as Coespeletia timotensis or C. moritziana. on current status of plant communities in the páramo
and which will undoubtedly have an influence on
future vegetation patterns (Llambí and Rada 2019).
Conclusions
Finally, there are two research aspects which need to
Based on the gathered evidence on plant functional be further studied in order to have a better under-
traits for the Venezuelan páramos, we suggest that standing of plant functioning and responses under
due to the functional diversity of species and exist- the high elevation extreme conditions and to compre-
ing redundancy within plant-growth forms, the per- hend the effects of future climate change on this eco-
petuation of most existing plant-growth forms and system. First, considering that establishment
the stability in terms of the general ecosystem prop- represents a filter phase determining the success of
erties would seem to be guaranteed. However, one any plant species in the páramo, especially since con-
of the main components of this ecosystem, the giant ditions at the soil surface are more extreme, studies on
rosettes, would be affected under projected limiting factors affecting seedling and juvenile fitness are
water availability conditions for the Northern required. Second, intra and interspecific relations
Andes and would place them under a vulnerable between plant species and growth-forms need to be
status. considered to better understand distribution patterns
12 F. RADA ET AL.

Figure 8. Schematic diagram showing the effects of temperature and water deficit on all studied plant growth-forms of the
Venezuelan Andean páramos. CR, Caulescent rosettes; Tr, Trees; WG, Woody grasses; Sh, Shrubs; AR, Acaulescent rosettes; F, Forbs;
CP, Cushion plants; Gr, Grasses. * Most shrub species studied avoided frost, although there were two which were frost tolerant.
This was the only growth-form which presented both avoidance and tolerance. Under predicted scenarios of increased water
deficit; caulescent rosettes, trees and woody grasses would seem to be most affected. Temperature per se seems to be less
important in determining vegetation dynamics in the páramos (see text for details).

in the páramos, especially at higher elevations. Under CRN2-2005]; Institute of International Education (IIE);
climate change scenarios, biotic pressures, e.g. compe- CDCHT-ULA.
tition, surely play an important role in the lower
páramo boundaries, as lower elevation plant species
move up. While at higher elevations, under harsher Notes on contributor
conditions, the existence of favourable microhabitats, F Rada, A Azócar and C García-Núñez specialise in plant
through facilitation; could allow associations between physiological ecology, focusing on plant responses to differ-
species which will, in the end, determine the occur- ent stresses, mainly related to temperature and water, in
rence of different plant communities (Cáceres et al. diverse tropical ecosystems.
2015; Hupp et al. 2017; Mora et al. 2019).

ORCID
Acknowledgements
Aura Azócar http://orcid.org/0000-0003-4486-2414
Financial support by the Consejo de Desarrollo Científico,
Humanístico, Tecnológico y de las Artes (CDCHTA-ULA),
Fondo Nacional de Ciencia, Tecnología e Innovación
References
(FONACIT) and the Inter-American Institute for Global
Change Research (IAI) throughout the years is gratefully Alvizu P. 2004. Complejidad y respuesta funcional de la
appreciated. FR wishes to thankfully acknowledge the vegetación de páramo a lo largo de gradientes altitudi-
Institute of International Education (IIE) and the nales [Doctoral Thesis]. Mérida (Venezuela): Facultad de
Universidad de Los Andes (UniAndes) for their significant Ciencias, Universidad de Los Andes, p. 115.
support commencing this year. The authors appreciate help- Arzac A, Chacón-Moreno E, Llambí LD, Dulhoste R. 2011.
ful comments on the manuscript made by the editor-in- Distribución de formas de vida de plantas en el límite
chief and two anonymous reviewers. superior del ecotono bosque-páramo en los Andes
Tropicales. Ecotrópicos. 24(1):26–46.
Arzac A, Llambí LD, Dulhoste R, Olano JM,
Disclosure statement Chacón-Moreno E. 2019. Modelling the effect of tem-
perature changes on plant life-form distribution in
No potential conflict of interest was reported by the authors.
a treeline ecotone of the tropical Andes. Plant Ecol
Divers. 12(this volume).
Funding Azócar A, Rada F. 2006. Ecofisiología de Plantas de Páramo.
Mérida (Venezuela): Publicaciones ICAE; p. 182.
This work was supported by the Fondo Nacional de Ciencia Azócar A, Rada F, García-Núñez C. 2000. Aspectos
Tecnología e Innovación [P12-0387]; Interamerican ecofisiológicos para la conservación de ecosistemas tro-
Institute for Global Change Studies (IAI) [CRN-040, picales contrastantes. B Soc Bot Mex. 65:89–94.
PLANT ECOLOGY & DIVERSITY 13

Azócar A, Rada F, Goldstein G. 1988. Freezing tolerance in glacier change. Reg Environ Change. 1–20. doi:10.1007/
Draba chionophylla, a ´miniature´ caulescent rosette spe- s10113-019-01499-3
cies. Oecologia. 75:156–160. Cuesta F, Muriel P, Llambí LD, Halloy S, Aguirre N, Beck S,
Azócar C. 2006. Relación entre anatomía, forma de vida Carilla M, Meneses RI, Cuello S, Grau A, et al. 2017.
y mecanismos de resistencia a temperaturas congelantes Latitudinal and altitudinal patterns of plant community
en diferentes especies de plantas de páramo diversity on mountain summits across the tropical Andes.
[Undergraduate Thesis]. Mérida (Venezuela): Facultad Ecography. 40:1–14.
de Ciencias, Universidad de Los Andes, p.64. Dulhoste R. 2010. Estrés hídrico y térmico en especies
Bader MY, van Geloof I, Rietkerk M. 2007. High solar leñosas de la zona de transición selva húmeda-páramo
radiation hinders tree regeneration above the alpine tree- [Doctorate Thesis]. Mérida (Venezuela): Facultad de
line in northern Ecuador. Plant Ecol. 191:33–45. Ciencias, Universidad de Los Andes. p. 141.
Baruch Z, Smith AP. 1979. Morphological and physiological Ely F. 2009. Comportamiento ecofisiológico y diversidad
correlates of niche breadth in two species of Espeletia genética de Chusquea (Bambuosoideae, Poaceae) en la
(Compositae) in the Venezuelan Andes. Oecologia. Cordillera de Mérida [Doctorate Thesis]. Mérida
38:71–82. (Venezuela): Facultad de Ciencias, Universidad de Los
Beck E, Senser M, Scheibe R, Steiger H, Pongratz P. 1982. Andes, p. 211.
Frost avoidance and freezing tolerance in afroalpine Ely F, Kiyota S, Rada F. 2014. Freezing avoidance in tropical
“giant rosette” plants. Plant Cell Environ. 5:215–222. Andean bamboos. Bamboo science and culture. J Amer
Beniston M. 2003. Climatic change in mountain regions: Bamboo Soc. 27:1–10.
a review of possible impacts. Clim Change. 59:5–31. Ely F, Rada F, Fermin G, Clark L. 2019. Ecophysiology and
Braun C, Bezada M. 2013. The history and disappearance of genetic diversity in high altitude woody bamboos of the
glaciers in Venezuela. J Lat Am Geogr. 12:85–124. Venezuelan Andes. Plant Ecol Divers. (this volume).
Briceño B. 1992. Intercambio de gases y relaciones hídricas Estrada C, Monasterio M. 1988. Ecología poblacional de una
en poblaciones del género Lupinus en un gradiente alti- roseta gigante, Espeletia spicata Sch Bip (Compositae) del
tudinal [Masters Thesis]. Mérida (Venezuela): Facultad páramo desértico. Ecotropicos. 1:25–39.
de Ciencias, Universidad de Los Andes, p. 162. Fariñas MR, Monasterio M. 1980. La vegetación del páramo
Buytaeart W, Cuesta-Camacho F, Tobón C. 2011. Potential de Mucubají. Análisis de ordenamiento y su
impacts of climate change on the environmental services interpretación ecológica. In: Monasterio M, editor.
of humid tropical alpine regions. Glob Ecol Biogeogr. Estudios Ecológicos en los Páramos Andinos. Mérida
20:19–33. (Venezuela): Universidad de Los Andes; p. 263–307.
Buytaeart W, Vuille M, Dewulf A, Urrutia R, Karmalkar A, Fariñas MR, Lázaro N, Monasterio M. 2008. Ecología compar-
Celleri R. 2010. Uncertainties in climate change projec- ada de Hypericum laricifolium Juss. Y de H. juniperinum
tions and regional downscaling in the tropical Andes: Kunth en el valle fluvioglacial del Páramo de
implications for water resources management. Hydrol Mucubají. Mérida, Venezuela. Ecotropicos. 21:75–88.
Earth Syst Sci. 14:1247–1258. García-Varela S. 2000. Mecanismos de resistencia en plantas
Cabrera HM. 2002. Respuestas ecofisiológicas de plantas en juveniles de Espeletia spicata y Espeletia timotensis en los
ecosistemas de zonas con clima mediterráneo Altos Andes venezolanos [Undergraduate Thesis].
y ambientes de alta montaña. Rev Chil Hist Nat. Mérida (Venezuela): Facultad de Ciencias, Universidad
75:625–637. de Los Andes, p. 105.
Cáceres Y. 2011. Relaciones espaciales y mecanismos de García-Varela S. 2008. Características adaptativas de difer-
interacción entre un arbusto dominante (Hypericum lar- entes formas de vida a lo largo de un gradiente sucesional
icifolium) y otras especies de plantas en el páramo en el páramo venezolano [Masters Thesis]. Mérida
Altiandino [Masters Thesis]. Mérida (Venezuela): (Venezuela): Facultad de Ciencias, Universidad de Los
Facultad de Ciencias, Universidad de Los Andes, p. 108. Andes, p. 80.
Cáceres Y, Llambí LD, Rada F. 2015. Shrubs as foundation García-Varela S, Rada F. 2003. Freezing avoidance mechan-
species in a high tropical alpine ecosystem: a multi-scale isms in juveniles of giant rosette plants of the genus
analysis of plant spatial interactions. Plant Ecol Divers. Espeletia. Acta Oecol. 24:165–167.
8:147–161. Goldstein G, Meinzer FC, Rada F. 1994. Environmental
Cavieres L, Rada F, Azócar A, García-Núñez CCabrera HM. biology of a tropical treeline species, Polylepis sericea.
2000. Gas exchange and low temperature resistance in In: Rundel PW, Smith AP, Meinzer FC, editors.
two tropical high mountain tree species from the vene- Tropical Alpine Environments: plant Form and
zuelan andes. Acta Oecologica. 21:203-211. doi:10.1016/ Function. Cambridge: Cambridge University Press; p.
S1146-609X(00)01077-8 129–149.
Cuesta F, Becerra MT. 2012. Biodiversidad y cambio Goldstein G, Meinzer FC. 1983. Influence of insulating dead
climático en los Andes: Importancia del monitoreo y el leaves and low temperatures on water balance in an
trabajo regional. Rev Virtual REDESMA. 6:19–27. Andean giant rosette plant. Plant Cell Environ.
Cuesta F, Llambí LD, Huggel C, Drenkhan F, Gosling WD, 6:649–656.
Muriel P, Jaramillo R, Tovar C. 2019. New land in the Goldstein G, Meinzer FC, Monasterio M. 1984. The role of
Neotropics: a review of biotic community, ecosystem and capacitance in the water balance of Andean giant rosette
landscape transformations in the face of climate and species. Plant Cell Environ. 7:179–186.
14 F. RADA ET AL.

Goldstein G, Rada F, Canales J, Zabala O. 1989. Leaf gas Llambí LD. 2015. Estructura, diversidad y dinámica de la
exchange of two giant caulescent rosette species. Acta vegetación en el ecotono bosque-páramo: revisión de la evi-
Oecol, Oecol Plant 10: 359-370. dencia en la Cordillera de Mérida. Acta Biolo Colomb.
Goldstein G, Rada F, Azócar A. 1985. Cold hardiness and 20:5–20.
supercooling along an altitudinal gradient in Andean Llambí LD, Fonataine M, Rada F, Saugier B, Sarmiento L.
giant rosette species. Oecologia. 68:147–152. 2003. Ecophysiology of dominant plant species during
Gosling WD, Bunting MJ. 2007. A role for palaeoecology in secondary succession in a high Andean páramo ecosys-
anticipating future change in mountain regions? tem. Arct Antarct Alp Res. 35:447–453.
Palaeoclimatol Palaeoecol. 259:1–5. Llambí LD, Rada F. 2019. Ecological research in tropical
Grabherr G, Gottfried M, Pauli H. 1994. Climate effects on alpine ecosystems of the Venezuelan páramo: past, pre-
mountain plants. Nature. 369:448–448. sent and future. Plant Ecol Divers. (this volume).
Guariguata MR, Azócar A. 1988. Seed bank dynamics and Luteyn JL. 1992. Páramos: why study them?. In: Balslev H,
germination ecology in Espeletia timotensis Luteyn JL, editors. Páramo, an Andean ecosystem under
(Compositae), an Andean giant rosette. Biotropica. human influence. London, UK: Academic Press; p. 1–14.
20:54–59. Luteyn JL. 1999. Páramos: a checklist of plant diversity,
Hedberg O. 1964. Features of Afroalpine plant ecology. Acta geographical distribution and botanical literature. Mem
Phytogeogr Suec. 49:1–44. N Y Bot Gard. 84: 1-278.
Hoch G, Körner C. 2005. Growth, demography and carbon Márquez EJ, Rada F, Fariñas MR. 2006. Freezing tolerance
relations of Polylepis trees at the world’s highest treeline. in grasses along an altitudinal gradient in the Venezuelan
Funct Ecol. 19:941–951. Andes. Oecologia. 150:393–397.
Hupp N, Llambí LD, Ramírez L, Callaway R. 2017. Alpine Mavárez J, Bézy S, Goeury T, Fernández A, Aubert S. 2019.
cushion plants have species-specific effects on microha- Current and future distributions of Espeletiinae (Asteraceae)
bitats and community structure in the tropical Andes. in the Cordillera de Mérida, Venezuela. A fine resolution
J Veg Sci. 28(5):928–938. analysis based on statistical downscaling of climatic variables
Inouye DW. 2008. Effects of climate change on phenology, and niche modeling. Plant Ecol Divers. (this volume).
frost damage, and floral abundance of montane Meinzer FC, Goldstein G. 1985. Some consequences of leaf
wildflowers. Ecology. 89:353–362. pubescence in the Andean giant rosette plant Espeletia
IPCC. 2013. Climate Change 2013: the physical Science timotensis. Ecology. 66:512–520.
Basis: working Group I contribution to the Fifth Monasterio M. 1980. Las formaciones vegetales de los
Assessment Report of the Intergovernmental Panel on Páramos de Venezuela. In: Monasterio M, editor.
Climate Change. New York: Cambridge University Estudios Ecológicos en los Páramos Andinos. Mérida
Press, p. 1523. (Venezuela): Universidad de Los Andes; p. 93–158.
Körner C. 2003. Alpine plant life, functional plant ecology of Monasterio M, Vuillemuier F. 1986. Introduction: high tro-
high mountain ecosystems. Berlin: Springer; p. 344. pical mountain biota of the world. In: Vuillemuier F,
Körner C. 2012. Alpine treelines. Functional ecology of the Monasterio M, editors. High altitude tropical biogeogra-
global high elevation tree limits. Berlin: Springer; p. 220. phy. Oxford: Oxford University Press; p. 3–7.
Körner C, Paulsen J. 2004. A world-wide study of high Monasterio M, Sarmiento L. 1991. Adaptive radiation of
altitude treeline temperatures. J Biogeogr. 31:713–732. Espeletia in the cold Andean tropics. Trends Ecol Evol.
Larcher W. 1975. Pflanzenokologische Beobachtungen in 6:387–391.
der páramostufe der Venezolanischen Anden. Anz Math- Mora MA, Llambí LD, Ramírez L. 2019. Giant stem rosettes
Naturw Kl Oest Akad Wissensch. 112:194–213. have strong facilitation effects on plant communities of
Larcher W. 2003. Physiological plant ecology. the high tropical Andes. (this volume).
Ecophysiology and stress physiology of functional Myers N, Mittermeirer RA, Mittermeier CG, Da
groups. 4th ed. Berlin: Springer; p. 513. Fonseca GAB, Kent J. 2000. Biodiversity hotspots for
Larcher W, Wagner J. 1976. Temperaturgrenzen der CO2- conservation priorities. Nature. 403:853–858.
aufnahme und temperaturresistenz der blätter von Navarro A. 2013. Relaciones hídricas en Ruilopezia atropur-
gebirgspflanzen in vegetationsaktiven Zustand. Oecol purea (A.C. Sm.) Cuatrec. a diferentes condiciones
Plant. 11:361–374. microambientales en el Páramo de San José,
Leuschner C. 2000. Are high elevations in tropical moun- estado Mérida [Masters Thesis]. Mérida (Venezuela):
tains arid environments for plants? Ecology. Facultad de Ciencias, Universidad de Los Andes, p.91.
81:1425–1436. Pauli H, Gottfried G, Dullinger S, Abdaladze O,
Levitt J. 1980. Responses of plants to environmental stresses, Akhalkatsi M, Alonso JLB, Coldea G, Dick J,
vol 1. Chilling, freezing and high temperature stresses. Erschbamer B, Fernández Calzado R, et al. 2012. Recent
New York: Academic Press; p. 510. plant diversity changes on Europe´s mountain summits.
Llambí LD, Ramírez L, Schwartzkopf T. 2013. Patrones de Science. 336:353–355.
distribución de planas leñosas en el ecotono bosque- Pauli H, Gottfried M, Grabherr G. 1996. Effects of climate
páramo de la Sierra Nevada de Mérdia: ¿Qué nos sugieren change on mountain ecosystems – upward shifting of
sobre la dinámica del límite del bosque? In: Cuesta F, alpine plants. World Resour Rev. 8:382–390.
Sevink J, Llambí LD, De Bievre B, Posner J, editors. Pirela M. 2006. Análisis funcional de la comunidad de
Avances en Investigación para la Conservación en los plantas en tres unidades geomorfológicas en el Páramo
Páramos Andinos. Quito: CONDESAN; p. 486–502. de Mucubají [Undergraduate Thesis]. Mérida
PLANT ECOLOGY & DIVERSITY 15

(Venezuela): Facultad de Ciencias, Universidad de Los Rehm EM, Feeley KJ. 2015b. The inability of tropical cloud
Andes, p. 95. forest species to invade grasslands above treeline during
Pouchon C, Fernández A, Nassar JM, Boyer F, Aubert S, climate change: potential explanations and consequences.
Lavergne S, Mavárez J. 2018. Phylogenomic analysis of the Ecography. 38:1167–1175. doi:10.1111/ecog.01050
explosive adaptive radiation of the Espeletia complex Rosquete C. 2004. Balance térmico en hojas de diferentes
(Asteraceae) in the tropical Andes. Syst Biol. 67:1041–1060. especies de los géneros Espeletia y Ruilopezia
Rabatel A, Francou B, Soruco A, Gomez J, Cáceres B, [Undergraduate Thesis]. Mérdia (Venezuela): Facultad
Ceballos JL, Scheel M. 2013. Current state of glaciers in de Ciencias, Universidad de Los Andes, p. 62.
the tropical Andes: a multi-century perspective on glacier Sakai A, Larcher W. 1987. Frost survival of plants: responses
evolution and climate change. Cryosphere. 7:81–102. and adaptations to freezing stress. Berlin: Springer; p. 321.
Rada F. 1993. Respuesta estomática y asimilación de CO2 en Sarmiento G. 1986. Ecologically crucial features of climate
plantas de distintas formas de vida en ambientes de baja in high tropical mountains. In: Vuilleumier F,
disponibilidad de CO2 en las altas montañas tropicales Monasterio M, editors. High altitude tropical biogeogra-
[Doctorate Thesis. Facultad de Ciencias]. Mérida phy. Oxford: Oxford University Press; p. 11–45.
(Venezuela): Universidad de Los Andes, p.135. Scherrer D, Körner C. 2011. Topography controlled
Rada F. 2016. Functional diversity in tropical high elevation thermal-habitat differentiation buffers alpine plant
giant rosettes. In: Goldstein G, Santiago LS, editors. diversity against climate warming. J Biogeogr.
Tropical tree physiology: adaptations and responses in 38:406–416.
a changing environment. Berlin: Springer; p. 181–202. Scherrer D, Schmid S, Körner C. 2011. Elevational species
Rada F, Azócar A, Briceño B, González J. 1998. Leaf gas shifts in a warmer climate are overestimated when based
exchange in Espeletia schultzii Wedd, a giant caulescent on weather station data. Int J Biometeorol. 55:645–654.
rosette, along an altitudinal gradient in the Venezuelan Sierra-Almeida A, Cavieres LA. 2010. Summper freezing
Andes. Acta Oecol. 19:73–79. resistance decreased in high-elevation plants exposed to
Rada F, Azócar A, Rojas-Altuve A. 2012. Water relations experimental warming in the central Chilean Andes.
and gas exchange in Coespeletia moritziana (Sch. Bip) Oecologia. 163:267–276.
Cuatrec., a giant rosette species of the high tropical Sklenář P, Hedberg I, Cleef AM. 2014. Island biogeography
Andes. Photosynthetica. 50:429–436. of tropical alpine floras. J Biogeogr. 41:287–297.
Rada F, Briceño B, Azócar A. 2008. How do two Lupinus Smith AP. 1974. Bud temperature in relation to nyctinastic
species respond to temperatura along an altitudinal gra- leaf movement in an Andean giant rosette plant.
dient in the Venezuelan Andes? Rev Chil De Hist Nat. Biotropica. 6:263–266.
81:335–343. Smith AP. 1979. The function of dead leaves in Espeletia
Rada F, García-Núñez C, Rangel S. 2011. Microclimate and schultzii (Compositae) an Andean giant rosette species.
regeneration patterns of Polylepis serícea in a treeline Biotropica. 11:43–47.
forest of the Venezuelan Andes. Ecotropicos. 24:113–122. Smith AP. 1981. Growth and population dynamics of
Rada F, Goldstein G, Azócar A, Meinzer F. 1985a. Freezing Espeletia (Compositae) of the Venezuelan Andes.
avoidance in Andean giant rosette plants. Plant Cell Smithsonian Contribut Bot. 48:1–45.
Environ. 8:501–507. Squeo F, Rada F, Azócar A, Goldstein G. 1991. Freezing
Rada F, Goldstein G, Azócar A, Meinzer F. 1985b. Daily and tolerance and avoidance in high tropical Andean plants:
seasonal osmotic changes in a tropical treeline species. is it equally represented in species with different plant
J Exp Bot. 36:989–1000. height? Oecologia. 86:378–382.
Rada F, Goldstein G, Azócar A, Torres F. 1987. Supercooling
Thuiller W, Albert C, Araújo MB, Berry PM, Cabeza M,
along an altitudinal gradient in Espletia schultzii, a caulescent
Guisan A, Hickler T, Midgley GF, Paterson J, Schurr FM,
giant rosette species. J Exp Bot. 38:491–497.
et al. 2008. Predicting global change impacts on plants´
Rada F, Azócar A, Briceño B, González J, García-Nuñez C.
species distribution: future challenges. Perspect Plant
1996. Carbon and water balance in Polylepis sericea, a
Ecol Evol Systemat. 9:137–152.
tropical treeline species. Trees 10: 218-222.
Rada F, González J, Azócar A, Briceño B, Jaimez R. 1992. Urrutia R, Vuille M. 2009. Climate change projections for
Net photosyntesis-leaf temperature relations in plant spe- the tropical Andes using a regional climate model: tem-
cies with different height along an altitudinal gradient. perature and precipitation simulations for the end of the
Acta Oecologica, 13: 535-542. 21st century. J Geophys Res. 114: D02108.
Ramírez L, Llambí LD, Schwarzkopf T, Gámez L, Márquez N. Vuille M, Bradley RS, Werner M, Keimig F. 2003. 20th
2009. Vegetation structure along the forest-páramo transition century climate change in the tropical Andes: observa-
belt in the Sierra Nevada de Mérida: implications for under- tions and model results. Clim Change. 59:75–99.
standing treeline dynamics. Ecotrópicos. 22:83–98. Walther G, Beiβner S, Pott R. 2005. Climate change and
Ramírez L, Rada F, Llambí LD. 2015. Linking patterns and high mountain vegetation shifts.
processes through ecosystem engineering: effects of shrubs Williams JW, Jackson ST. 2007. Novel climates, no-analog
on microhabitat and water status of associated plants in the communities, and ecological surprises. Front Ecol
high tropical Andes. Plant Ecol. 216:213–225. Environ. 5:475–482.
Rehm EM, Feeley KJ. 2015a. Freezing temperatures as a Woodward FI. 1990. The impact of low temperatures in
limit to forest recruitment above tropical andean tree- controlling the geographical distribution of plants. Phil
lines. Ecology. 96:1856–1865. doi:10.1890/14-1992.1 Trans R Soc Lond B. 326:585–593.

You might also like