Skip to main content
Log in

The functional dependence of canopy conductance on water vapor pressure deficit revisited

  • Original Paper
  • Published:
International Journal of Biometeorology Aims and scope Submit manuscript

Abstract

Current research seeking to relate between ambient water vapor deficit (D) and foliage conductance (gF) derives a canopy conductance (gW) from measured transpiration by inverting the coupled transpiration model to yield gW = m − n ln(D) where m and n are fitting parameters. In contrast, this paper demonstrates that the relation between coupled gW and D is gW = AP/D + B, where P is the barometric pressure, A is the radiative term, and B is the convective term coefficient of the Penman-Monteith equation. A and B are functions of gF and of meteorological parameters but are mathematically independent of D. Keeping A and B constant implies constancy of gF. With these premises, the derived gW is a hyperbolic function of D resembling the logarithmic expression, in contradiction with the pre-set constancy of gF. Calculations with random inputs that ensure independence between gF and D reproduce published experimental scatter plots that display a dependence between gW and D in contradiction with the premises. For this reason, the dependence of gW on D is a computational artifact unrelated to any real effect of ambient humidity on stomatal aperture and closure. Data collected in a maize field confirm the inadequacy of the logarithmic function to quantify the relation between canopy conductance and vapor pressure deficit.

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

Similar content being viewed by others

References

  • Assouline S (2001) A model of soil relative hydraulic conductivity based on the water retention characteristic curve. Water Resour Res 37:265–271

    Article  Google Scholar 

  • Brutsaert W (1982) Evaporation into the atmosphere. D. Reidel Publishing Co., Dordrecht, Boston

    Book  Google Scholar 

  • Campbell GS, Norman JM (1998) An introduction to environmental biophysics. Springer, New York

    Book  Google Scholar 

  • Campbell GS, Papendick RI, Rabie E, Shayo-Ngowi AJ (1979) A comparison of osmotic potential, elastic modulus, and apoplastic water in leaves of dryland winter wheat. Agron J 71:31–36

    Article  Google Scholar 

  • Chahine MT (1992) The hydrological cycle and its influence on climate. Nature 359:373–380

    Article  Google Scholar 

  • Cohen S, Naor A (2002) The effect of rootstocks on water use, canopy conductance and hydraulic parameters of apple trees and predicting canopy from hydraulic conductance. Plant Cell Environ 25:17–28

    Article  Google Scholar 

  • David TS, Feirreira MI, Cohen S et al (2004) Constraints on transpiration from an evergreen oak tree in southern Portugal. Agric For Meteorol 122:193–205

    Article  Google Scholar 

  • Dodd I (2003) Hormonal interactions and stomatal responses. J Plant Growth Regul 22:32–46

    Article  CAS  Google Scholar 

  • Eichinger WE, Parlange MB, Stricker H (1996) On the concept of equilibrium evaporation and the value of the Priestley-Taylor coefficient. Water Resour Res 32:161–164

    Article  CAS  Google Scholar 

  • Fuchs M, Tanner CB (1966) Infrared thermometry of vegetation. Agron J 58:597–601. https://doi.org/10.2134/agronj1966.00021962005800060014x

    Article  Google Scholar 

  • Granier A, Breda N (1996) Modelling canopy conductance and stand transpiration of an oak forest from sap flow measurements. Ann des Sci For 53:537–546

    Article  Google Scholar 

  • Granier A, Loustau D (1994) Measuring and modelling the transpiration of a maritime pine canopy from sap-flow data. Agric For Meteorol 71:61–81

    Article  Google Scholar 

  • Granier A, Huc R, Barigah ST (1996) Transpiration of natural rain forest and its dependence on climatic factors. Agric For Meteorol 78:19–29

    Article  Google Scholar 

  • Hernandez-Santana V, Fernandez JE, Rodriguez-Dominguez CM et al (2016) The dynamics of radial sap flux density reflects changes in stomatal conductance in response to soil and air water deficit. Agric For Meteorol 218(219):92–101

    Article  Google Scholar 

  • Igarashi Y, Kumagai T, Yoshifuji N et al (2015) Environmental control of canopy stomatal conductance in a tropical deciduous forest in northern Thailand. Agric For Meteorol 202:1–10

    Article  Google Scholar 

  • Jagtap SS, Jones JW (1989) Stability of crop coefficients under different climate and irrigation management practices. Irrig Sci 10:231–244

    Article  Google Scholar 

  • Jones HG (1992) Plants and microclimate. Cambridge University Press, Cambridge

    Google Scholar 

  • Kelly G, Moshelion M, David-Schwartz R, Halperin O, Wallach R, Attia Z, Belausov E, Granot D (2013) Hexokinase mediates stomatal closure. Plant J 75:977–988

    Article  CAS  Google Scholar 

  • Lange OL, Schulze E-D, Kappen L (1971) Responses of stomata to changes in humidity. Planta 100:76–86

    Article  CAS  Google Scholar 

  • Langensiepen M, Fuchs M, Bergamaschi H, Moreshet S, Cohen Y, Wolff P, Jutzi SC, Cohen S, Rosa LMG, Li Y, Fricke T (2009) Quantifying the uncertainties of transpiration calculations with the Penman-Monteith equation under different climate and optimum water supply conditions. Agric For Meteorol 149:1063–1072

    Article  Google Scholar 

  • List RJ (1966) Smithsonian meteorological tables. The Smithsonian Institution, Washington, DC

    Google Scholar 

  • Lohammar T, Larsson S, Linder S, Falk SO (1980) FAST: simulation models of gaseous exchange in scots pine. Ecol Bull 32:505–523

  • Martin TA, Brown KJ, Cermák J, Ceulemans R, Kucera J, Meinzer FC, Rombold JS, Sprugel DG, Hinckley TM (1997) Crown conductance and tree and stand transpiration in a second-growth Abies amabilis forest. Can J For Res 27:797–808

    Article  Google Scholar 

  • McAdam SAM, Brodribb TJ (2015) The evolution of mechanisms driving the stomatal response to vapor pressure deficit. Plant Physiol 167:833–843

    Article  CAS  Google Scholar 

  • McNaughton KG, Black TA (1973) A study of evapotranspiration from a Douglas fir forest using the energy balance approach. Water Resour Res 9:1579–1590

    Article  Google Scholar 

  • McNaughton KG, Jarvis PG (1983) Predicting effects of vegetation changes on transpiration and evaporation. In: Kozlowski TT (ed) Water deficit in plant growth. Academic Press, New York, pp 1–47

    Google Scholar 

  • Monteith JL (1965) Evaporation and environment. In: Fogg GE (ed) The state and movement of water in living organisms. Cambridge University Press, Cambridge, pp 205–234

    Google Scholar 

  • Monteith JL (1995) A reinterpretation of stomatal responses to humidity. Plant Cell Environ 18:357–364

    Article  Google Scholar 

  • Mott KA, Peak D (2010) Stomatal responses to humidity and temperature in darkness. Plant Cell Environ 33:1084–1090

    Google Scholar 

  • Ocheltree TW, Nippert JB, Prasad PVV (2014) Stomatal responses to changes in vapor pressure deficit reflect tissue-specific differences in hydraulic conductance. Plant Cell Environ 37:132–139

    Article  CAS  Google Scholar 

  • Oren R, Aasamaa K, Sperry JS et al (1999) Survey and synthesis of intra- and interspecific variation in stomatal sensitivity to vapour pressure deficit. Plant Cell Environ 22:1515–1926

    Article  Google Scholar 

  • Overdieck D, Strain BR (1981) Effects of atmospheric humidity on net photosynthesis, transpiration, and stomatal resistance. Int J Biometeorol 25:29–38. https://doi.org/10.1007/BF02184435

    Article  Google Scholar 

  • Peak D, Mott KA (2011) A new, vapour-phase mechanism for stomatal responses to humidity and temperature. Plant Cell Environ 34:162–178

    Article  Google Scholar 

  • Penman HL (1948) Natural evaporation from open water, bare soil, and grass. Proc R Soc Lond A Math Phys Sci 193:120–146

    Article  CAS  Google Scholar 

  • Petersen KL, Moreshet S, Fuchs M (1991) Stomatal responses of field-grown cotton to radiation and soil moisture. Agron J 83:1059–1065

    Article  Google Scholar 

  • Renninger HJ, Carlo NJ, Clark KL, Schäfer KVR (2015) Resource use and efficiency, and stomatal responses to environmental drivers of oak and pine species in an Atlantic Coastal Plain forest. Front Plant Sci 6(297):1–16. https://doi.org/10.3389/fpls.2015.00297

  • Schoppach R, Fleury D, Sinclair TR, Sadok W (2016) Transpiration sensitivity to evaporative demand across 120 years of breeding of Australian wheat cultivars. J Agron Crop Sci 203:219–226. https://doi.org/10.1111/jac.12193

    Article  CAS  Google Scholar 

  • Sinclair TR, Zwieniecki MA, Holbrook NM (2008) Low leaf hydraulic conductance associated with drought tolerance in soybean. Physiol Plant 132:446–451

    Article  CAS  Google Scholar 

  • Slatyer RO, McIlroy IC (1961) Practical Microclimatology. Commonwealth Scientific and Industrial Research Organization, Melbourne, Australia

    Google Scholar 

  • Sweet KJ, Peak D, Mott KA (2017) Stomatal heterogeneity in responses to humidity and temperature: testing a mechanistic model. Plant Cell Environ 40:2771–2779. https://doi.org/10.1111/pce.13051

    Article  CAS  Google Scholar 

  • Tardieu F, Simonneau T (1998) Variability among species of stomatal control under fluctuating soil water status and evaporative demand: modelling isohydric and anisohydric behaviours. J Exp Bot 49:419–432

    Article  Google Scholar 

  • Tardieu F, Zhang J, Davies WJ (1992) What information is conveyed by an ABA signal from maize roots in drying field soil? Plant Cell Environ 15:185–191

    Article  CAS  Google Scholar 

  • Tetens O (1930) Über einige meteorologische Begriffe. Z Geophys 6:297–309

    Google Scholar 

Download references

Acknowledgements

We thank Dr. Shabtai Cohen for providing us the original data used in Fig. 3.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Marcel Fuchs.

Appendix

Appendix

The calculation of transpiration uses the approach and numerical values given in (Campbell and Norman 1998). Sunlit and shaded leaves were treated separately because of differences in the way they intercept radiation and differences in leaf conductance values (Jagtap and Jones 1989). In the following development index, j designates the class to which the considered leaves belong:

$$ {\displaystyle \begin{array}{l}j=0;\kern1em \mathrm{shaded}\kern0.5em \mathrm{leaves}\\ {}j=1;\kern1em \mathrm{sunlit}\kern0.5em \mathrm{leaves}\end{array}} $$
(A.1)

Radiation balance

The global radiation intercepted by the foliage in class j, Rg j is:

$$ {R}_{g\;j}=j\kern0.5em f\kern0.5em {R}_r\;{L}_j+x\frac{L_j}{L}\left({R}_d+{S}_d\right) $$
(A.2)

The symbols of Eq. (A.2) are defined here below:

The ray interception factor f by leaves with random inclination and orientation is:

$$ f=0.5\;\sec \left(\eta \right) $$
(A.3)

where η is the solar zenith angle.

L is the leaf area index, Lj is:

$$ {\displaystyle \begin{array}{l}{L}_j\kern0.5em \mathrm{for}\kern0.5em j=1\to {L}_1=\frac{\left[1-\exp \left(-f\;L\right)\right]}{f}\\ {}{L}_j\kern0.5em \mathrm{for}\kern0.5em j=0\to {L}_0=L-{L}_1\end{array}} $$
(A.4)

The diffuse interception x of uniformly distributed hemispherical rays is:

$$ x=\frac{1}{\pi }{\int}_0^{2\pi }{\int}_0^{\raisebox{1ex}{$\pi $}\!\left/ \!\raisebox{-1ex}{$2$}\right.}\left[1-\exp \left(-f\;L\right)\right]\sin \left(\eta \right)\cos \left(\eta \right)\partial \eta\;\partial \varphi $$
(A.5)

(η = zenith angle, φ = azimuth angle, of an incident ray).

The beam Rr and diffuse Rd components of global radiation through a cloudless sky are calculated from the extraterrestrial solar constant Ro = 1356 W m−2 (List 1966):

$$ {\displaystyle \begin{array}{l}{R}_r={R}_o\cos \left(\eta \right)\times {0.65}^{\raisebox{1ex}{$P\sec \left(\eta \right)$}\!\left/ \!\raisebox{-1ex}{$101.3$}\right.}\\ {}{R}_d={R}_o\cos \left(\eta \right)\times 0.34\times \left(1-{0.65}^{\raisebox{1ex}{$P\sec \left(\eta \right)$}\!\left/ \!\raisebox{-1ex}{$101.3$}\right.}\right)\\ {}P=\mathrm{baromertic}\ \mathrm{pressure},\left(\mathrm{kPa}\right)\end{array}} $$
(A.6)

The foliage also intercepts Sd, the secondary scattering of global radiation approximated as:

$$ {S}_d\approx 0.22\times \left\{{R}_r\left[1-\exp \left(-f\;L\right)\right]+x\;{R}_d\right\} $$
(A.7)

Assigning a value of 0.55 to leaf solar radiation absorptivity, \( {\varepsilon}_s=0.767\times {e}_a^{1/7} \) (ea = air vapor pressure in kPa) to sky emissivity of terrestrial radiation (Brutsaert 1982) and assuming the canopy emissivity of terrestrial radiation is close to unity (Fuchs and Tanner 1966), Rn j the net radiation of leaves in class j is:

$$ {\displaystyle \begin{array}{l}{R}_{n\;j}=0.55{R}_{g\;j}+x\frac{L_j}{L}\left({\varepsilon}_s\sigma {T}_{a\kern0.24em K}^4-\sigma {T}_{F\;K\;j}^4\right)\\ {}\sigma {T}_{F\;K\;j}^4\approx \sigma {T}_{a\kern0.24em K}^4+4\sigma {T}_{a\;K}^3\left({T}_{F\;j}-{T}_a\right)\end{array}} $$
(A.8)

Here, σ = 5.67 × 10−8 W m‐2 K‐4; Ta K and TF K j are the Kelvin air and foliage temperatures; Ta and TF j are the Celsius air and foliage temperatures.

We define an intermediate radiation term Rm j:

$$ {R}_{m\;j}=0.55{R}_{g\;j}+x\frac{L_j}{L}\left({\varepsilon}_s-1\right)\sigma {T}_{a\;K}^4 $$
(A.9)

Equation (A.8) becomes:

$$ {R}_{n\;j}={R}_{m\;j}-4x\frac{L_j}{L}\sigma {T}_{a\;K}^3\left({T}_{F\;j}-{T}_a\right) $$
(A.10)

Energy balance

We define a radiative conductance gr j as:

$$ {g}_{r\;j}=4x\frac{L_j}{L}\frac{\sigma {T}_{a\;K}^3}{c_p} $$
(A.11)

Here, cp is the specific heat of air (= 29.3 J mol−1 C−1). Setting radiative flux density Hr j as:

$$ {H}_{r\;j}={c}_p{g}_{r\;j}\left({T}_{F\;j}-{T}_a\right) $$
(A.12)

leads to the energy balance of the leaves in class j:

$$ {R}_{m\;j}=\lambda {E}_j+{H}_{c\;j}+{H}_{r\;j} $$
(A.13)

where λ is the latent of vaporization (= 43,384 J mol−1 at T = 300 K) and λEj is the latent heat flux density:

$$ \lambda {E}_j=\frac{\lambda }{P}{g}_{V\;j}\left[e\left({T}_{F\;j}\right)-{e}_a\right] $$
(A.14)

where e(TFj) is the saturated vapor pressure at TFj and ea is the vapor pressure in the air.

Hc j is the convective sensible heat flux density:

$$ {H}_{c\;j}={c}_p{g}_{H\;j}\left({T}_{F\;j}-{T}_a\right) $$
(A.15)

Transport coefficients

In Eq. (A.14), gV j is the conductance of the vapor path from the evaporating locations within the plant tissues to the free air above the canopy:

$$ {g}_{V\;j}=\frac{L_j}{r_{V\;c}+{r}_{L\;j}} $$
(A.16)

rV c is the convective resistance per unit leaf area (rV c = 1/gV c) of air for water vapor transport from the outer surface of a leaf to freely moving air above the canopy.

rL j is the parallelly connected diffusive leaf resistance per unit leaf area through stomata and cuticle of the adaxial and abaxial faces of the leaves (rL j = 1/gL j, gLj is the leaf conductance).

The foliage conductance gFj is then defined as:

$$ {\displaystyle \begin{array}{l}{g}_{Fj}={L}_j{g}_{Lj}=\frac{L_j}{r_{Lj}}\\ {}{g}_F={g}_{F0}+{g}_{F1}\end{array}} $$
(A.17)

The convective resistance rV c is made of:

$$ {r}_{V\;c}={r}_a+0.5\kern0.5em {r}_{V\;b} $$
(A.18)

Here, ra is the convective resistance of the turbulent air layer above the canopy assigned to a unit leaf area:

$$ {r}_a=\frac{\ln \left(\frac{z-d}{z_0}\right)\ln \left(\frac{z-d}{0.2\;{z}_0}\right)L}{\rho\;{k}^2u(z)} $$
(A.19)

where z is the level above the canopy taken from the ground at which wind speed u(z) is measured, d = 0.66 × Z is the wind profile displacement level, z0 = 0.13 × Z is the roughness length, Z is the canopy height, k is the von Kármán constant (= 0.40), and ρ is the air molar density [=44.6 × (Ta K/273.2) × (P/101.3) mol m−3.]

rV b in Eq. (A.18) is the one-sided convective resistance per unit leaf area for vapor transport in the laminar air boundary layer of a leaf surface. The factor 0.5 halves the resistance because leaves are two-sided.

$$ {r}_{V\;b}=6.8\left(\raisebox{1ex}{${l}^{0.5}$}\!\left/ \!\raisebox{-1ex}{$\overline{u_{\mathrm{\ell}}^{0.5}}$}\right.\right) $$
(A.20)

l is the characteristic length of the leaf and \( \overline{u_{\mathrm{\ell}}^{0.5}} \) is defined in Eq. (A.25).

Heat convective conductance gH j in Eq. (A.15) is:

$$ {g}_{H\;j}=\frac{L_j}{r_a+0.5\;{r}_{H\;b}} $$
(A.21)

where rH b is the one-side convective resistance per unit leaf area for heat transport in the laminar air boundary layer surrounding a leaf:

$$ {r}_{H\;b}=7.4\left(\raisebox{1ex}{${l}^{0.5}$}\!\left/ \!\raisebox{-1ex}{$\overline{u_{\mathrm{\ell}}^{0.5}}$}\right.\right) $$
(A.22)

The combined conductance of heat gT j is:

$$ {g}_{T\;j}={g}_{H\;j}+{g}_{r\;j} $$
(A.23)

The boundary-layer convective resistance is proportional to \( 1/\overline{u_{\mathrm{\ell}}^{0.5}} \). The mean value \( \overline{u_{\mathrm{\ell}}^{0.5}} \) in the canopy is derived from the shear that foliage skin friction and form drag exert on wind:

$$ \frac{\partial {u}_{\mathrm{\ell}}}{\mathrm{\partial \ell }}=a\;{u}_{\mathrm{\ell}} $$
(A.24)

Here, ℓis the leaf area index (LAI) depth at which the wind has penetrated (at the top of the canopy ℓ = L = LAI; ℓ = 0 at the bottom of the canopy), uis the wind speed at ℓ, and a is the combined drag and friction coefficient set at 0.4. The mean \( \overline{u_{\mathrm{\ell}}^{0.5}} \) adjusted to the wind profile in the canopy is:

$$ \overline{u_{\mathrm{\ell}}^{0.5}}=\frac{2{u}_Z^{0.5}}{a\;L}\left[1-\exp \left(-a\;L/2\right)\right] $$
(A.25)

Here, uZ is the wind speed at canopy height Z:

$$ {u}_Z=u(z)\frac{\ln \left(\frac{Z-d}{z_0}\right)}{\ln \left(\frac{z-d}{z_0}\right)} $$
(A.26)

Additional definitions

The additional elements needed to complete the calculation of Ej are listed here.

The saturated water vapor pressure function of temperature T is (Tetens 1930):

$$ e(T)=0.6018\times \exp \left(\frac{17.27\times T}{T+237.3}\right);\kern1em \left(\mathrm{kPa}\right) $$
(A.27)

The slope of the molar fraction vapor saturation function of temperature is:

$$ \Delta =\frac{1}{P}\frac{\partial e\left({T}_a\right)}{\partial {T}_a}=\frac{e\left({T}_a\right)}{P}\times \frac{17.27\times 237.3}{{\left({T}_a+237.3\right)}^2}\approx \frac{e\left({T}_{F\;j}\right)-e\left({T}_a\right)}{P\left({T}_{F\;j}-{T}_a\right)};\kern1em \left({\mathrm{C}}^{-1}\right) $$
(A.28)

Finalizing

Combining Eqs (A.12), (A.13), (A.14), (A.15), and (A.28) leads to the P-M equation:

$$ {E}_j=\frac{\raisebox{1ex}{$\Delta\;{R}_{m\kern0.24em j}$}\!\left/ \!\raisebox{-1ex}{$\lambda\;{\Gamma}_j$}\right.+{g}_{V\;j}\raisebox{1ex}{$D$}\!\left/ \!\raisebox{-1ex}{$P$}\right.}{1+\raisebox{1ex}{$\Delta $}\!\left/ \!\raisebox{-1ex}{${\Gamma}_j$}\right.} $$
(A.29)

Here, Γj is the psychrometric constant (C−1) adjusted to the conductance of heat gT j and vapor gVj for the leaves in class j.

$$ {\Gamma}_j=\raisebox{1ex}{${c}_p\;{g}_{T\;j}$}\!\left/ \!\raisebox{-1ex}{$\lambda\;{g}_{V\;j}$}\right. $$
(A.30)

The resulting value of foliage transpiration is:

$$ E={E}_0+{E}_1 $$
(A.31)

The coupled model of transpiration is obtained when u(z) → ∞. This condition draws in Eqs. (A.20), (A.19), (A.18), (A.17), (A.16), rVb, ra and rVc → 0, and gVj → gFj.

Coupling also implies that in Eq. (A.22), rHb → 0 and in Eq. (A.21) gHj → ∞. Consequently in Eq.(A.23), gTj → ∞ leading to Γj → ∞ in Eq.(A.30). The implications on convection deriving from coupling transforms Eq. (A.29) into:

$$ {E}_j\to {E}_{\infty j}={g}_{Fj}\frac{D}{P} $$
(A.32)

leading through Eqs. (A.17) and (A.31) to Eq. (6).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Fuchs, M., Stanghellini, C. The functional dependence of canopy conductance on water vapor pressure deficit revisited. Int J Biometeorol 62, 1211–1220 (2018). https://doi.org/10.1007/s00484-018-1524-4

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00484-018-1524-4

Keywords

Navigation