Skip to main content
Log in

The role of plant water storage and hydraulic strategies in relation to soil moisture availability

  • Regular Article
  • Published:
Plant and Soil Aims and scope Submit manuscript

Abstract

Background and Aims

Plants rely on water storage capacity to increase accessibility of water for transpiration, reduce competition for water with neighboring plants, and buffer water supply during dry periods. The resulting benefits, typically a decrease in plant water stress and increase in productivity, are highly climate dependent and vary with soil moisture, vapor pressure deficit, and solar radiation. This paper analyzes the effects of plant water storage capacity on the relationship between soil moisture and carbon assimilation in woody plants.

Methods

A resistance-capacitance model is used to examine the role of plant water storage at various soil moisture levels. Hydraulic traits are co-varied according to empirical relationships, and effects of sapwood volume and wood density on carbon assimilation are explored. The time scales of plant water storage and withdrawal are analyzed as a function of plant hydraulic capacitance, water storage capacity, and resistance to transport between water storage tissue and xylem.

Results

The effects of plant water storage on carbon assimilation are found to depend strongly on soil moisture levels. The theoretically optimal sapwood volume lies near naturally occurring ranges and increases with increasing soil moisture. The theoretically optimal wood density also lies within expected ranges and decreases with increasing soil moisture.

Conclusions

A large portion of sapwood volume appears to be justified by its role in buffering diurnal variability in evaporative demand. The outlined coordination between soil moisture and optimal hydraulic traits is consistent with observed increases in sapwood capacitance and decreases in wood density across increasing rainfall gradients. This coordination provides support for the drought-tolerance vs. drought-avoidance hypothesis.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5

Similar content being viewed by others

References

  • Ackerly D (2004) Functional Strategies of Chaparral Shrubs in Relation to Seasonal Water Deficit and Disturbance. Ecol Monogr 74(1):25–44

    Article  Google Scholar 

  • Aumann CA, Ford ED (2002) Modeling tree water flow as an unsaturated flow through a porous medium. J Theor Biol 219(4):415–29

    Article  PubMed  Google Scholar 

  • Bartlett MS, Vico G, Porporato A (2014) Coupled carbon and water fluxes in CAM photosynthesis: modeling quantification of water use efficiency and productivity. Plant Soil 383(1-2):111–138

    Article  CAS  Google Scholar 

  • Bohrer G, Mourad H, Laursen TA, Drewry D, Avissar R, Poggi D, Oren R, Katul GG (2005) Finite element tree crown hydrodynamics model (FETCH) using porous media flow within branching elements: A new representation of tree hydrodynamics. Water Resour Res 41(11):1–17

    Article  Google Scholar 

  • Bucci SJ, Goldstein G, Meinzer FC, Scholz FG, Franco AC, Bustamante M (2004) Functional convergence in hydraulic architecture and water relations of tropical savanna trees: from leaf to whole plant. Tree Physiolx 24(8):891–899

    Article  CAS  Google Scholar 

  • Burghardt M, Riederer M (2003) Ecophysiological relevance of cuticular transpiration of deciduous and evergreen plants in relation to stomatal closure and leaf water potential. J Exp Bot 54(389):1941–9

    Article  CAS  PubMed  Google Scholar 

  • Carlson T, Lynn B (1991) The effects of plant water storage on transpiration and radiometric surface temperature. Agric For Meteorol 57(1986)

  • Cermák J, Kucera J, Bauerle WL, Phillips N, Hinckley TM (2007) Tree water storage and its diurnal dynamics related to sap flow and changes in stem volume in old-growth Douglas-fir trees. Tree Physiol 27(2):181–98

    Article  PubMed  Google Scholar 

  • Chapotin SM, Razanameharizaka JH, Holbrook NM (2006a) Baobab trees (Adansonia) in Madagascar use stored water to flush new leaves but not to support stomatal opening before the rainy season. Phytologist 169(3):549–59

    Article  Google Scholar 

  • Chapotin SM, Razanameharizaka JH, Holbrook NM (2006b) Water relations of baobab trees (Adansonia spp. L.) during the rainy season: does stem water buffer daily water deficits? Plant, Cell Environ 29(6):1021–1032

    Article  Google Scholar 

  • Chave J, Muller-landau HC, Baker TR, Easdale TA, Webb CO (2006) Regional and Phylogenetic Variation of Wood Density across 2456 Neotropical Tree Species. Ecol Appl 16(6):2356–2367

    Article  PubMed  Google Scholar 

  • Chuang YL, Oren R, Bertozzi AL, Phillips N, Katul GG (2006) The porous media model for the hydraulic system of a conifer tree: Linking sap flux data to transpiration rate. Ecol Modell 191 (3-4):447–468

    Article  Google Scholar 

  • Daly E, Porporato A, Rodriguez-Iturbe I (2004) Coupled Dynamics of Photosynthesis, Transpiration, and Soil Water Balance. Part I: Upscaling from Hourly to Daily Level. J Hydrometeorol 5(3):546–558

    Article  Google Scholar 

  • Domec JC, Gartner BL (2001) Cavitation and water storage capacity in bole xylem segments of mature and young Douglas-fir trees. Trees - Struct Funct 15(4):204–214

    Article  Google Scholar 

  • Farquhar G, Caemmerer SV, Berry J (1980) A biochemical model of photosynthetic CO2 assimilation in leaves of C3 species. Planta 149:78–90

    Article  CAS  PubMed  Google Scholar 

  • Flexas J, Ribas-Carbó M, Diaz-Espejo A, Galmés J, Medrano H (2008) Mesophyll conductance to CO2: Current knowledge and future prospects. Plant, Cell Environ 31(5):602–621

    Article  CAS  Google Scholar 

  • Hacke UG, Sperry JS, Ewers BE, Ellsworth DS, Schäfer K V R, Oren R (2000) Influence of soil porosity on water use in Pinus taeda. Oecologia 124(4):495–505

    Article  CAS  PubMed  Google Scholar 

  • 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 negative pressure. Oecologia 126(4):457–461

    Article  PubMed  Google Scholar 

  • Hartzell S, Bartlett MS, Virgin L, Porporato A (2015) Nonlinear dynamics of the CAM circadian rhythm in response to environmental forcing. J Theor Biol 368:83–94

    Article  CAS  PubMed  Google Scholar 

  • Hubbard RM, Bond BJ, Ryan MG (1999) Evidence that hydraulic conductance limits photosynthesis in old Pinus ponderosa trees. Tree Physiol 19(3):165–172

    Article  PubMed  Google Scholar 

  • Hunt RE, Running SW, Federer C (1991) Extrapolating plant water flow resistances and capacitances to regional scales. Agric For Meteorolx 54(2–4):169–195

    Article  Google Scholar 

  • Ishii HR, Azuma W, Kuroda K, Sillett SC (2014) Pushing the limits to tree height: could foliar water storage compensate for hydraulic constraints in Sequoia sempervirens?, pp 1087–1093

  • Jones H (1992) Plants and microclimate: a quantitative approach to environmental plant physiology. Cambridge University Press

  • Kattge J, Knorr W (2007) Temperature acclimation in a biochemical model of photosynthesis: a reanalysis of data from 36 species. Plant, Cell Environ 30(9):1176–90

    Article  CAS  Google Scholar 

  • Leuning R (1995) A critical appraisal of a combined stomatal-photosynthesis model for C3 plants

  • Lhomme JP, Elguero E, Chehbouni A, Boulet G (1998) Stomatal control of transpiration: Examination of Monteith’s formulation of canopy resistance. Water Resour Res 34(9):2301–2308

    Article  CAS  Google Scholar 

  • Long SP, Postl WF, Bolhár-Nordenkampf HR (1993) Quantum yields for uptake of carbon dioxide in C3 vascular plants of contrasting habitats and taxonomic groupings. Planta 189(2):226–234

    Article  CAS  Google Scholar 

  • Manzoni S, Vico G, Katul G, Palmroth S, Porporato A (2014) Optimal plant water-use strategies under stochastic rainfall. Water Resources Research, pp 1–16

  • Matheny AM, Fiorella RP, Bohrer G, Poulsen CJ, Morin TH, Wunderlich A, Vogel CS, Curtis PS (2017) Contrasting strategies of hydraulic control in two codominant temperate tree species. Ecohydrology 10(3):1–16

    Article  Google Scholar 

  • Mcculloh KA, Johnson DM, Meinzer FC, Woodruff DR (2014) The dynamic pipeline: Hydraulic capacitance and xylem hydraulic safety in four tall conifer species. Plant, Cell Environ 37(5):1171–1183

    Article  Google Scholar 

  • McNaughton KG, Spriggs TW (1986) A mixed-layer model for regional evaporation. Bound-Layer Meteorol 34(3):243–262

    Article  Google Scholar 

  • Meinzer FC, Johnson DM, Lachenbruch B, McCulloh KA, Woodruff DR (2009) Xylem hydraulic safety margins in woody plants: Coordination of stomatal control of xylem tension with hydraulic capacitance. Funct Ecol 23(5):922–930

    Article  Google Scholar 

  • Mirfenderesgi G, Bohrer G, Matheny AM, Fatichi S, de Moraes Frasson RP, Schäfer KVR (2016) Tree-level hydrodynamic approach for modeling aboveground water storage and stomatal conductance illuminates the effects of tree hydraulic strategy. J Geophys Res: Biogeosci 121:1792–1813

    Article  Google Scholar 

  • Nobel PS, Jordan PW (1983) Transpiration stream of desert species: resistances and capacitances for a C3, a C4 and a CAM plant. J Exp Bot 34(147):1379–1391

    Article  Google Scholar 

  • Nolf M, Creek D, Duursma R, Holtum J, Mayr S, Choat B (2015) Stem and leaf hydraulic properties are finely coordinated in three tropical rain forest tree species. Plant, Cell Environ 38 (12):2652–2661

    Article  CAS  Google Scholar 

  • Ogburn RM, Edwards EJ (2010) The ecological water-use strategies of succulent plants, vol 55, 1st edn. Elsevier Ltd.

  • Onoda Y, Richards AE, Westoby M (2010) The relationship between stem biomechanics and wood density is modified by rainfall in 32 Australian woody plant species. Phytologist 185(2):493–501

    Article  Google Scholar 

  • Phillips N, Nagchaudhuri A, Oren R, Katul G (1997) Time constant for water transport in loblolly pine trees extimated from time series of evaporative demand and stem sapflow. Trees, pp 412–419

  • Phillips NG, Oren R, Licata J, Linder S (2004) Time series diagnosis of tree hydraulic characteristics. Tree Physiol 24(8):879–890

    Article  PubMed  Google Scholar 

  • Preston KA, Cornwell WK, DeNoyer JL (2006) Wood density and vessel traits as distinct correlates of ecological strategy in 51 California coast range angiosperms. Phytologist 170(4):807–818

    Article  Google Scholar 

  • Richards AE, Wright IJ, Lenz TI, Zanne AE (2013) Sapwood capacitance is greater in evergreen sclerophyll species growing in high compared to low-rainfall environments. Funct Ecol 28(3):734–744

    Article  Google Scholar 

  • Rodriguez-Iturbe I, Porporato A (2004) Ecohydrology of water-controlled ecosystems. Cambridge University Press

  • Rodriguez-Iturbe I, Porporato A, Ridolfi L, Isham V, Coxi DR (1999) Probabilistic modelling of water balance at a point: The role of climate, soil and vegetation. Proc Royal Soc Lond Ser A: Mathe Phys Eng Sci 455(1990):3789–3805

    Article  Google Scholar 

  • Ryan MG (1990) Growth and maintenance respiration in stems of Pinus contorta and Picea engelmannii. Canadian Journal of Forest Research

  • Scholz FG, Bucci SJ, Goldstein G, Meinzer FC, Franco AC, Miralles-Wilhelm F (2007) Biophysical properties and functional significance of stem water storage tissues in Neotropical savanna trees. Plant, Cell Environ 30(2):236–248

    Article  Google Scholar 

  • Scholz FG, Phillips NG, Bucci SJ, Meinzer FC, Goldstein G (2011) Size- Age-Related Chang Tree Struct Funct 4: 341–361

    Article  Google Scholar 

  • Sperry JS (2004) Coordinating stomatal and xylem functioning - an evolutionary perspective. Phytologist 162(3):563–567

    Article  Google Scholar 

  • Sperry JS, Adler FR, Campbell GS, Comstock JP (1998) Limitation of plant water use by rhizosphere and xylem conductance: Results from a model. Plant, Cell Environ 21(4):347–359

    Article  Google Scholar 

  • Thomas SC, Winner WE (2000) Leaf area index of an old-growth Douglas-fir forest estimated from direct structural measurements in the canopy. Can J For Res 30(12):1922–1930

    Article  Google Scholar 

  • Tyree M, Ewers F (1991) The hydraulic architecture of trees and other woody plants. Phytologist 119(34):345–360

    Article  Google Scholar 

  • Waring R, Running S (1978) Sapwood water storage: its contribution to transpiration and effect upon water conductance through the stems of old growth Douglas fir. Plant, Cell and Environment, pp 131–140

  • Williams M, Bond BJ, Ryan MG (2001) Evaluating different soil and plant hydraulic constraints on tree function using a model and sap flow data from ponderosa pine. Plant, Cell Environ 24(7):679–690

    Article  Google Scholar 

Download references

Acknowledgments

This work was supported through the USDA Agricultural Research Service cooperative agreement 58-6408-3-027 and National Institute of Food and Agriculture (NIFA) grant 12110061; and National Science Foundation (NSF) grants CBET-1033467, EAR-1331846, FESD-1338694, EAR-1316258, and the Duke WISeNet Grant DGE-1068871. We thank Gil Bohrer and one anonymous reviewer for their useful suggestions.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Amilcare Porporato.

Additional information

Responsible Editor: Susan Schwinning.

Appendices

Appendix A: Conductances

Following Daly et al. (2004), the soil-root conductance, g s r , is assumed to be proportional to the soil hydraulic conductivity, K(s), divided by the average distance from the soil to root surface, i.e.,

$$ g_{sr}(s) = \frac{K(s)\sqrt{RAI_{w} s^{-d}}}{\pi g\rho_{w} Z_{r}}, $$
(16)

where R A I w is the root area index under well-watered conditions, s d is a term introduced to model root growth under water-stressed conditions, and K(s) = s (2b+3). See Table 6.

The decrease in plant conductance under water stress is modeled by a vulnerability curve so that g p is close to g p m a x for high ψ l and is close to 0 for low ψ l due to xylem cavitation (Sperry et al. 1998; Daly et al. 2004), i.e.,

$$ g_{p} = g_{pmax}\left[1+\left( \frac{\psi_{x}}{\psi_{50}}\right)^{a}\right]^{-1}. $$
(17)

where a is a parameter between 1 and 10, here assumed to be equal to 4, and ψ 50 is the xylem potential at 50% loss of hydraulic conductivity. Following Waring and Running (1978) and Carlson and Lynn (1991), the storage conductance, or the conductance between the water storage and transport pathways, is assumed to decrease with the fraction of stored water as

$$ g_{w} = g_{wmax} w^{m}, $$
(18)

where g w m a x is the maximum storage conductance (Table 1).

Appendix B: Water potentials

The soil water potential, ψ s , is related to the soil moisture through a strongly nonlinear function given by Rodriguez-Iturbe and Porporato (2004) and Daly et al. (2004) as

$$ \psi_{s}(s) = \overline{\psi}_{s} s^{-b}. $$
(19)

The relationship between the stored water potential, ψ w , and the stored water volume is most commonly represented by a logarithmic or sigmoidal curve (e.g. Chuang et al. 2006) and the specific plant water capacitance c, is defined as the change in relative stored water volume per unit change in water potential (c = d w/d ψ w ). In this study we have chosen to approximate the stored water potential as a linear function of the relative water storage volume following Hunt et al. (1991), i.e.,

$$ \psi_{w}(w) = \frac{w - 1}{c}. $$
(20)

While a simplification, this linear relationship is a good approximation in the physically relevant regime examined here, where the relative stored water volume remains above 40%. In this range, it compares very well to data (see Fig. 2, data from Waring and Running 1978 and Domec and Gartner 2001). Moreover, this approximation allows us to more clearly tease out the effects of plant hydraulic capacitance on the dynamics of water storage.

Appendix C: Leaf temperature and specific humidity

The leaf temperature and specific humidity, which influence plant transpiration and carbon assimilation, are are obtained by coupling the plant model to a boundary layer model as in Daly et al. (2004). The leaf temperature, T l , can be obtained through the plant energy balance as

$$ \phi-H-\lambda_{w}\rho_{w}E = 0, $$
(21)

where ϕ is the net flux of radiation per unit leaf area, λ w is the latent heat of water, ρ w is the density of water, and H is the sensible heat flux.

$$ H = c_{p}\rho g_{a}(T_{l}-T_{a}), $$
(22)

where c p is the specific heat of air at constant pressure and g a is the atmospheric conductance. Substituting (22) into (21) and rearranging leads to an explicit formula for T l as

$$ T_{l} = T_{a} + \frac{\phi-\rho_{w} \lambda_{w} E}{c_{p} \rho g_{a}}. $$
(23)

The specific humidity of the leaf mesophyll can be calculated as

$$ q_{l}(T_{l},\psi_{l}) = q_{l,sat}(T_{l})\exp\left[\frac{V_{w}\psi_{l}}{R T_{l}}\right], $$
(24)

where V w is the partial molar volume of water, and the specific humidity at saturation, q l,s a t is related to T l as

$$ q_{l,sat}(T_{l}) = \frac{0.622}{p_{a}}a_{sat}\exp\left[\frac{b_{sat}(T_{l}-273)}{c_{sat}+T_{l}-273}\right], $$
(25)

where p a is the atmospheric pressure (Pa) and a s a t , b s a t and c s a t are constants (Jones 1992). See Table 4.

Table 4 Atmospheric parameters

Appendix D: Boundary layer model

The boundary layer is modeled according to McNaughton and Spriggs (1986), Daly et al. (2004) as a well-mixed slab of air with height h, specific humidity q, and potential temperature 𝜃. Within the boundary layer, the equations for heat and water balance, respectively, are

$$ \rho c_{p} h \frac{d\theta}{dt} = H + \rho c_{p}(\theta_{s} - \theta)\frac{dh}{dt}, $$
(26)
$$ \rho h \frac{dq}{dt} = \rho_{w} E + \rho(q_{s}-q)\frac{dh}{dt}, $$
(27)

where 𝜃 s and q s are the potential temperature and humidity, respectively, at height h. The growth rate of the boundary layer is given by

$$ \frac{dh}{dt} = \frac{H}{\rho c_{p} h \gamma_{\theta}}, $$
(28)

where γ 𝜃 is the gradient of the potential temperature at height h. The linear profiles of temperature and humidity in the atmosphere above the boundary layer respectively are expressed as

$$ \theta_{s} = \gamma_{\theta}z + \theta_{s0}, $$
(29)

and

$$ q_{s} = \gamma_{q}z + q_{s0}. $$
(30)

The solar radiation, ϕ, is modeled as

$$ \phi(t) = \frac{4\phi_{max}}{\delta^{2}}[-t^{2} + (\delta + 2t_{0})t - t_{0}(t_{0} + \delta)], $$
(31)

where ϕ m a x is the maximum solar radiation, δ is the day length, and t 0 is the time of sunrise (Lhomme et al. 1998). See Table 4.

Appendix E: Details of the carbon assimilation and stomatal conductance models

R, the ratio of mesophyll to atmospheric CO 2 concentration and can be related to the vapor pressure deficit, D, is given by

$$ R = 1 -\frac{1}{a_{1}}\left( 1+\frac{D}{D_{x}}\right), $$
(32)

where a 1 and D x are empirical constants (Leuning 1995).

We assume that, under well-watered conditions, there is no control of carbon assimilation by the leaf water potential up to the point ψ l,A1, at which point the assimilation decreases linearly to 0 at the point ψ l,A0, i.e.,

$$\begin{array}{@{}rcl@{}} A_{\psi_{l}}(\psi_{l})=\left\{ \begin{array}{ll} \text{0} & \quad \psi_{l} < \psi_{lA0} \\ \frac{(\psi_{l}-\psi_{lA0})}{(\psi_{lA1}-\psi_{lA0})} & \quad \psi_{lA0} < \psi_{l} \leq \psi_{lA1} \ \\ \text{1} & \quad \psi_{l} > \psi_{l_{A1}}. \ \end{array} \right. \end{array} $$
(33)

The function \(A_{\phi ,c_{i},T_{l}}(\phi , c_{i}, T_{l})\) is given by

$$ A_{\phi,c_{i},T_{l}}(\phi, c_{i}, T_{l}) = \min{(A_{c}, A_{q})}, $$
(34)

and describes the dependence of carbon assimilation on ϕ, c i , and T l given by the Farquhar model (Farquhar et al. 1980) for C 3 plants. The Rubisco-limited rate of carbon assimilation, A c , is given by

$$ A_{c}=V_{c,max}\frac{c_{i}-{\Gamma}^{*}(T_{l})}{c_{i}+K_{c}(T_{l})(1+o_{i}/K_{o}(T_{l}))}, $$
(35)

where V c,m a x is the maximum carboxylation rate, K c and K o are the Michaelis-Menten coefficients for CO 2 and O 2, respectively, o i is the oxygen concentration, and Γ is the CO 2 compensation point. The light-limited assimilation rate, A q , is

$$ A_{q} = \frac{J(\phi, T_{l})}{4}\frac{(c_{i}-{\Gamma}^{*}(T_{l}))}{(c_{i}+2{\Gamma}^{*}(T_{l}))}, $$
(36)

where J, the electron transport rate, is equal to min (J m a x (T l ),Q). Q is the absorbed photon irradiance (mol photons m −2 s −1) and is given by

$$ Q(\phi)=\frac{\phi\lambda\kappa_{2}}{2N_{a}hc}, $$
(37)

where 50 percent of the incoming radiation is considered photosynthetically active radiation (PAR) (Jones 1992), λ is the average wavelength (m) for PAR (assumed to be 550 nm), h is Planck’s constant (Js), c is the speed of light (m/s), N a is Avogadro’s constant (mol −1), and κ 2 is the quantum yield of photosynthesis in mol CO 2 mol −1 photons. J m a x (T l ) is given by

$$ J_{max}(T_{l})=J_{max0}\frac{\exp[\frac{H_{vJ}}{RT_{0}}(1-\frac{T_{0}}{T_{l}})]}{1+\exp(\frac{S_{vQ}T_{l}-H_{dJ}}{RT_{l}})}. $$
(38)

The maximum carboxylation rate, V c,m a x , and the CO 2 compensation point, Γ, are given by

$$ V_{c,max}(T_{l})=V_{c,max0}\frac{\exp[\frac{H_{vJ}}{RT_{0}}(1-\frac{T_{0}}{T_{l}})]}{1+\exp(\frac{S_{vC}T_{l}-H_{dJ}}{RT_{l}})} $$
(39)

and

$$ {\Gamma}^{*}(T_{l})={\Gamma}_{0}[1+{\Gamma}_{1}(T_{l}-T_{0})+{\Gamma}_{2}(T_{l}-T_{0})^{2}], $$
(40)

where the temperature dependence of the Michaelis-Menten constants K c and K o is described by a modified Arrhenius equation, i.e.,

$$ K_{x}(T_{l})=K_{x0}\exp\left[\frac{H_{Kx}}{RT_{0}}\left( 1-\frac{T_{0}}{T_{l}}\right)\right], $$
(41)

where x stands for either c or o. Parameter values are given in (Tables 5 and 6).

Table 5 Plant photosynthetic parameters
Table 6 Soil parameters

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Hartzell, S., Bartlett, M. & Porporato, A. The role of plant water storage and hydraulic strategies in relation to soil moisture availability. Plant Soil 419, 503–521 (2017). https://doi.org/10.1007/s11104-017-3341-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11104-017-3341-7

Keywords

Navigation