You are on page 1of 4

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/322959719

Xylem dysfunction in fires: towards a hydraulic theory of plant responses to


multiple disturbance stressors

Article  in  New Phytologist · February 2018


DOI: 10.1111/nph.15013

CITATIONS READS

15 923

1 author:

Sean Michaletz
University of British Columbia - Vancouver
121 PUBLICATIONS   2,082 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Leaf temperatures, traits, and rates View project

Trait Driver Theory View project

All content following this page was uploaded by Sean Michaletz on 06 February 2018.

The user has requested enhancement of the downloaded file.


Forum

Commentary

Xylem dysfunction in fires: How do fires injure plants?

towards a hydraulic theory of Heat transfer from a fire into a plant can injure the roots, stem, or
crown (Fig. 1; Michaletz & Johnson, 2007). Root injuries occur via
plant responses to multiple conduction heat transfer through soil and into the roots, while
stems and crowns are heated by radiation and convection to the
disturbance stressors plant and conduction within the plant. Heating by fire can cause
necrosis of living cells if they exceed a threshold temperature of
c. 60°C. Necrosis of meristematic tissues can limit or prevent
It is often thought that a wildfire will consume and kill all of the growth of other critical tissues and organs such as phloem, xylem,
vegetation within its perimeter, but this is more an exception than a roots and leaves. Heating may also cause dysfunction of xylem
rule. Indeed, heterogeneity of fuels and microclimate leads to tissue, which limits xylem water flow. This can reduce stomatal
heterogeneity of fire behavior and effects, so that injured but conductance and rates of carbon assimilation.
surviving plants often remain after a wildfire (Bond & Van Wilgen, Heat injuries may then interact to influence whole-plant function
1996). This has important emergent outcomes spanning levels of and mortality. There are two main hypotheses for post-fire plant
biological organization, from cellular photosynthesis and respira- mortality: the cambium necrosis hypothesis and the xylem
tion to ecosystem production and evapotranspiration. However, dysfunction hypothesis (Balfour & Midgley, 2006; Kavanagh
despite more than half a century of research, the mechanisms by et al., 2010; Midgley et al., 2011; Michaletz et al., 2012). According
which fire injuries occur and interact are not well understood to the cambium necrosis hypothesis, phloem and cambium necrosis
(Michaletz & Johnson, 2007). limits carbon translocation to roots, so that root growth must rely
upon stored carbon reserves. When these reserves are depleted, fine-
root production ceases and plant mortality occurs as a result of
hydraulic failure (McDowell et al., 2008). According to the xylem
‘They also show for the first time that wildfires
dysfunction hypothesis, heating reduces the hydraulic conductivity
can permanently alter xylem vulnerability to of the xylem, which increases xylem water tensions, increases
periods of stomatal closure, and limits carbon assimilation and
cavitation, rendering plants more susceptible to future growth. Plant mortality may then result from hydraulic failure or
carbon starvation (McDowell et al., 2008; Sevanto et al., 2014).
disturbances . . .’
Experimental support for the xylem dysfunction
In this issue of New Phytologist, B€ar et al. (pp. 1484-1493) hypothesis
provide a critical step forward in our understanding of fire effects on Plant mortality in fires has traditionally been thought to result from
plants by presenting the first quantitative evidence that wildfires cambium necrosis (Michaletz & Johnson, 2007), but there is
cause xylem dysfunction. Using branches collected from trees that growing support for a more hydraulics-based view of fire effects on
survived a recent wildfire, they show that fires can permanently plants. Stem heating experiments have shown that heat impairs
reduce xylem conductivity via conduit wall deformation. This had plant hydraulic function. For example, stem heating caused
previously only been shown in laboratory experiments (Michaletz reductions in sap flux density, stomatal conductance, and net
et al., 2012; West et al., 2016). They also show for the first time that photosynthesis (Ducrey et al., 1996). While these results are
wildfires can permanently alter xylem vulnerability to cavitation, consistent with the xylem dysfunction hypothesis, they are not
rendering plants more susceptible to future disturbances such as fire conclusive since the experiments also caused phloem and cambium
(Brando et al., 2014) and drought (Anderegg et al., 2013). This had necrosis. Thus, it is unclear whether the results reflect limitation of
not been observed in previous studies that tested for it (Michaletz fine-root growth by phloem necrosis, limitation of xylem growth by
et al., 2012; Battipaglia et al., 2016). The findings of B€ar et al. shed cambium necrosis, or heat-induced xylem dysfunction. In another
new light on the complex set of mechanisms governing fire effects study, stem heating reduced the cross-sectional area of functional
on plants, and contribute to a growing framework for understand- xylem, which by definition reduces the hydraulic conductivity of the
ing and predicting plant responses to multiple interacting distur- stem (Balfour & Midgley, 2006). These reductions were semi-
bance stressors. permanent following reflushing to remove air embolisms, suggest-
ing that heating reduced the xylem conductivity by one or more
This article is a Commentary on B€ar et al., 217: 1484–1493. mechanisms that were not fully consistent with air seed cavitation.

Ó 2018 The Author New Phytologist (2018) 217: 1391–1393 1391


New Phytologist Ó 2018 New Phytologist Trust www.newphytologist.com
New
1392 Forum Commentary Phytologist

common they are in real wildfires. Several studies provide indirect


Crown injury evidence for xylem dysfunction in wildfires. For example, heat
• Leaf necrosis
• Bud meristem necrosis transfer simulations forced with wildfire temperature data pre-
• Phloem & cambium necroses dicted that tree stems can experience substantial reductions in the
• Xylem dysfuncon cross-sectional area of functional xylem (Michaletz et al., 2012).
Consistent with these predictions, analyses of xylem anatomy in
fire-injured tree stems have revealed large areas of discolored and
presumably nonfunctional wood (Fig. 2; Smith et al., 2016).
Convecon Stem injury
• Phloem & cambium necroses
Wildfires have also been shown to reduce stomatal conductance and
• Xylem dysfuncon predawn water potential (Thompson et al., 2017), and also cause
Radiaon
higher mortality rates compared with stems that had their phloem
and cambium removed (Midgley et al., 2011). Post-fire mortality
rates also vary inversely with wood density (Brando et al., 2012),
Conducon Root injury
• Fine root necrosis which likely reflects the role of wood density in prevention of heat-
• Phloem & cambium necroses induced cavitation and conduit wall deformation (Hacke et al.,
• Xylem dysfuncon 2001; Michaletz et al., 2012). While all of these results are
Fig. 1 Heat transfer from a wildfire to a plant, and key injuries that may result consistent with xylem dysfunction in fires, they also likely reflect the
from heat conduction within the plant. influence of other fire injuries and thus cannot confirm the xylem
dysfunction hypothesis (Fig. 1).
Building on this work, Michaletz et al. (2012) used laboratory air
injection experiments to demonstrate that heating reduces the
hydraulic conductivity of xylem via at least two mechanisms: (1) air Towards a hydraulic theory of plant responses to
seed cavitation resulting from temperature-dependent changes in disturbance
sap surface tension; and (2) conduit wall deformation resulting
The results of B€ar et al. provide the first direct quantitative evidence
from thermal softening of viscoelastic cell wall polymers (lignin,
for xylem dysfunction in wildfires. This is an important contribution
hemicelluloses, and cellulose). Both of these mechanisms were
to our understanding of plant responses to disturbance and helps set
subsequently observed to reduce xylem conductivity in laboratory
an agenda for future research. For example, B€ar et al. observed
heat plume experiments (West et al., 2016). While air seed
reduced xylem conductivity in angiosperms but not gymnosperms,
cavitation can be repaired, conduit wall deformation is permanent
and these results were associated with differences in the severity of
once the xylem cools and viscoelastic polymers return to a glassy
conduit wall deformation between angiosperms and gymnosperms.
state. Thus, conduit wall deformation is especially injurious.
This suggests that angiosperms and gymnosperms may differ in the
Thermal softening effects on pit membrane structure would also
kinetics of conduit wall polymer softening, or that softened vessels
permanently alter cavitation vulnerability, but only marginal and
and tracheids may differ in their responses to stresses imposed by
nonsignificant changes were observed in these experiments
tensile sap water. Further work is needed to identify how variation in
(Michaletz et al., 2012).
xylem traits operates via these mechanisms to drive variation in fire
effects. B€ar et al. also provide the first documentation of changes in
Can wildfires cause xylem dysfunction? vulnerability to cavitation following wildfire. While all tested species
Although air seed cavitation and conduit wall deformation have experienced increases in vulnerability, the magnitude of these
been observed in laboratory experiments, it has been less clear how changes varied among species in a manner more complex than

(a) (b)

Fig. 2 Anatomical evidence is consistent with heat-induced xylem dysfunction following wildfire (Smith et al., 2016). (a) Cross-section of western larch (Larix
occidentalis Nutt.) containing healed fire scars and areas of xylem discoloration. (b) Detail from white box in (a) shows a substantial area of discolored wood
(DW), heartwood (HW), and sapwood (SW; functional xylem). The areas of discoloration are likely nonfunctional xylem, but this has not been confirmed by
hydraulic conductivity measurements. Photographs courtesy of Kevin Smith.

New Phytologist (2018) 217: 1391–1393 Ó 2018 The Author


www.newphytologist.com New Phytologist Ó 2018 New Phytologist Trust
New
Phytologist Commentary Forum 1393

simple angiosperm/gymnosperm or vessel/tracheid dichotomies. overview of drought and heat-induced tree mortality reveals emerging climate
Along with previous experiments reporting nonsignificant increases change risks for forests. Forest Ecology and Management 259: 660–684.
Anderegg WRL, Plavcova L, Anderegg LDL, Hacke UG, Berry JA, Field CB.
in vulnerability (Michaletz et al., 2012), this suggests future work is 2013. Drought’s legacy: multiyear hydraulic deterioration underlies widespread
needed to understand the mechanisms by which variation in thermal aspen forest die-off and portends increased future risk. Global Change Biology 19:
softening and pit membrane morphology leads to variation in post- 1188–1196.
fire cavitation vulnerability. Balfour DA, Midgley JJ. 2006. Fire induced stem death in an African acacia is not
Despite our growing knowledge of how plant physiology responds caused by canopy scorching. Austral Ecology 31: 892–896.
B€ar A, Nardini A, Mayr S. 2018. Post-fire effects in xylem hydraulics of Picea abies,
to disturbances such as fire and drought, we still have a limited Pinus sylvestris and Fagus sylvatica. New Phytologist 217: 1484–1493.
understanding of how the responses interact to control whole-plant Battipaglia G, Savi T, Ascoli D, Castagneri D, Esposito A, Mayr S, Nardini A.
function and mortality, especially over longer periods of time 2016. Effects of prescribed burning on ecophysiological, anatomical and stem
(Michaletz & Johnson, 2007; Allen et al., 2010; van Mantgem et al., hydraulic properties in Pinus pinea L. Tree Physiology 36: 1019–1031.
2013). Conduit wall deformation and increased cavitation vulner- Bond WJ, Van Wilgen BW. 1996. Fire and plants (population and community biology
series 14). Population and community biology. London, UK: Chapman & Hall.
ability are permanent fire injuries that may contribute to cumulative Brando PM, Balch JK, Nepstad DC, Morton DC, Putz FE, Coe MT, Silverio D,
xylem dysfunction and make plants more susceptible to future Macedo MN, Davidson EA, Nobrega CC et al. 2014. Abrupt increases in
disturbances (Anderegg et al., 2013). Indeed, post-fire mortality rates Amazonian tree mortality due to drought–fire interactions. Proceedings of the
are strongly influenced by the number of fires a tree experiences National Academy of Sciences, USA 111: 6347–6352.
(Brando et al., 2012), and are higher for water-stressed plants subject Brando PM, Nepstad DC, Balch JK, Bolker B, Christman MC, Coe M, Putz FE.
2012. Fire-induced tree mortality in a neotropical forest: the roles of bark traits,
to drought (van Mantgem et al., 2013; Brando et al., 2014). This tree size, wood density and fire behavior. Global Change Biology 18: 630–641.
indicates an important role for fire-induced xylem dysfunction in the Ducrey M, Duhoux F, Huc R, Rigolot E. 1996. The ecophysiological and growth
long-term responses of plants to climate and disturbance, and responses of Aleppo pine (Pinus halepensis) to controlled heating applied to the
suggests that understanding and predicting these responses is more base of the trunk. Canadian Journal of Forest Research 26: 1366–1374.
complex than previously thought. Additional work is required to Hacke UG, Sperry JS, Pockman WT, Davis SD, McCulloh KA. 2001. Trends in
wood density and structure are linked to prevention of xylem implosion by
understand the relative roles of cavitation and deformation in negative pressure. Oecologia 126: 457–461.
reduction of xylem conductivity, and how xylem dysfunction in Kavanagh KL, Dickinson MB, Bova AS. 2010. A way forward for fire-caused tree
roots, stems, and branches correlates and interacts with other injuries mortality prediction: modeling a physiological consequence of fire. Fire Ecology 6:
to influence plant function and mortality (Fig. 1; Michaletz et al., 80–94.
2012). For example, necrosis and cavitation (hydraulic segmentation) van Mantgem PJ, Nesmith JC, Keifer M, Knapp EE, Flint A, Flint L. 2013.
Climatic stress increases forest fire severity across the western United States.
of leaves may act as a ‘hydraulic fuse’ that limits tension and cavitation Ecology Letters 16: 1151–1156.
in proximal parts of the xylem network (Kavanagh et al., 2010; West McDowell N, Pockman WT, Allen CD, Breshears DD, Cobb N, Kolb T, Plaut J,
et al., 2016; Wolfe et al., 2016). Understanding how such processes Sperry J, West A, Williams DG et al. 2008. Mechanisms of plant survival and
may correlate and interact with those demonstrated by B€ar et al. is a mortality during drought: why do some plants survive while others succumb to
next step towards incorporating cumulative disturbance impacts into drought? New Phytologist 178: 719–739.
Michaletz ST, Johnson EA. 2007. How forest fires kill trees: a review of the
climate–vegetation models (Anderegg et al., 2013). fundamental biophysical processes. Scandinavian Journal of Forest Research 22:
500–515.
Acknowledgements Michaletz ST, Johnson EA, Tyree MT. 2012. Moving beyond the cambium
necrosis hypothesis of post-fire tree mortality: cavitation and deformation of
The author thanks Kevin Smith for kindly providing the xylem in forest fires. New Phytologist 194: 254–263.
photographs used in Fig. 2, and acknowledges the support from Midgley JJ, Kruger LM, Skelton R. 2011. How do fires kill plants? The hydraulic
the Thomas R. Brown Family Foundation. death hypothesis and Cape Proteaceae ‘fire-resisters’. South African Journal of
Botany 77: 381–386.
Sevanto S, McDowell NG, Dickman LT, Pangle R, Pockman WT. 2014. How do
ORCID trees die? A test of the hydraulic failure and carbon starvation hypotheses. Plant,
Sean T. Michaletz X http://orcid.org/0000-0003-2158-6525 Cell & Environment 37: 153–161.
Smith KT, Arbellay E, Falk DA, Sutherland EK. 2016. Macroanatomy and
compartmentalization of recent fire scars in three North American conifers.
Sean T. Michaletz1,2 X Canadian Journal of Forest Research 46: 535–542.
Thompson MTC, Koyama A, Kavanagh KL. 2017. Wildfire effects on physiological
1 properties in conifers of central Idaho forests, USA. Trees 31: 545–555.
Biosphere 2, University of Arizona, Tucson, AZ 85721, USA; and
2 West AG, Nel JA, Bond WJ, Midgley JJ. 2016. Experimental evidence for heat
Department of Ecology and Evolutionary Biology, University of
plume-induced cavitation and xylem deformation as a mechanism of rapid post-
Arizona, Tucson, AZ 85721, USA fire tree mortality. New Phytologist 211: 828–838.
(tel+1 520 626 3336; email michaletz@email.arizona.edu) Wolfe BT, Sperry JS, Kursar TA. 2016. Does leaf shedding protect stems from
cavitation during seasonal droughts? A test of the hydraulic fuse hypothesis. New
Phytologist 212: 1007–1018.

References Key words: cavitation, disturbance, drought, embolism, fire, global change,
hydraulic conductivity, mortality.
Allen CD, Macalady AK, Chenchouni H, Bachelet D, McDowell N, Vennetier M,
Kitzberger T, Rigling A, Breshears DD, Hogg EH (T.) et al. 2010. A global

Ó 2018 The Author New Phytologist (2018) 217: 1391–1393


New Phytologist Ó 2018 New Phytologist Trust www.newphytologist.com

View publication stats

You might also like