<<

bs_bs_banner

Plant, and Environment (2015) 38, 7–22 doi: 10.1111/pce.12372

Original Article The contributions of apoplastic, symplastic and gas phase pathways for transport outside the bundle sheath in

Thomas N. Buckley

IA Watson Grains Research Centre, Faculty of Agriculture and Environment, The University of Sydney, Narrabri, New South Wales 2390, Australia

ABSTRACT which light, temperature and dehydration cause changes in the outside- component (Kox) of hydraulic conduct- Water movement from the xylem to stomata is poorly under- ance (Kleaf), to what degree are different leaf tissues stood. There is still no consensus about whether apoplastic or hydraulically coupled to one another, and perhaps most symplastic pathways are more important, and recent work compellingly,how do stomata sense and respond to changes in suggests vapour diffusion may also play a role. The objective water supply and evaporative demand? of this study was to estimate the proportions of hydraulic The prevailing viewpoint that has guided most investiga- conductance outside the bundle sheath contributed by tions into these questions assumes that Kleaf, calculated as the apoplastic, symplastic and gas phase pathways, using a novel ratio of water flow to difference in between the analytical framework based on measurable anatomical and petiole water source and bulk leaf tissues, represents a series biophysical parameters. The calculations presented here of liquid phase pathways:through the xylem,then through the suggest that apoplastic pathways provide the majority of con- bundle sheath and finally through the mesophyll and/or epi- ductance outside the bundle sheath under most conditions, dermis via apoplastic and/or symplastic routes (Tyree et al. whereas symplastic pathways contribute only a small propor- 1999; Zwieniecki et al. 2007). The apoplastic pathway is often tion. The contributions of apoplastic and gas phase pathways implicitly assumed to be dominated by bulk flow rather than vary depending on several critical but poorly known or highly by molecular diffusion of water (Strugger 1943, Tanton & variable parameters namely, the effective Poiseuille radius Crowdy 1972, Byott & Sheriff 1976, Evert et al. 1985, but see for apoplastic bulk flow, the thickness of cell walls and verti- Canny 1990).The possibility that symplastic diffusion may be cal temperature gradients within the leaf. The gas phase con- important in outside-xylem water flow was enhanced by dis- ductance should increase strongly as the leaf centre becomes covery of aquaporins (Agre et al. 1993; Chrispeels & Agre warmer than the – providing up to 44% of vertical 1994), and by experiments demonstrating accumulation of water transport for a temperature gradient of 0.2 K. These tracers at the proximal margin of living cells adjoining the results may help to explain how leaf water transport is influ- xylem (Canny 1986), which strongly indicated symplastic flow enced by light absorption, temperature and differences in of water. As a result, suspicion has grown in recent decades leaf anatomy among species. that much or most flow outside the xylem may be symplastic (Tyree et al. 1981, 1999; Sack et al. 2005). The notion that leaf Key-words: ; leaf hydraulic conductance; stomata; water flow should be strongly limited by high-resistance liquid stream. phase pathways for water diffusion outside the xylem, whether apoplastic or symplastic, was further advanced by INTRODUCTION Brodribb et al. (2007), who showed a strong inverse correla- tion between K and the pathlength for liquid phase water water transport and stomatal regulation depend leaf flow from xylem to stomata. strongly on the distribution of hydraulic resistance and water The intercellular air spaces can also provide a conduit for potential along the transpiration stream. This distribution is water transport outside the xylem, although this pathway has somewhat well characterized within stem and leaf xylem, but received relatively little attention. Boyer (1985), commenting it remains poorly understood for the pathways taken by water on data of Farquhar & Raschke (1978) showing that the after exiting the xylem in the leaf. Questions that are linked to pathlength for vapour diffusion from the sites of evaporation water flow outside the xylem and remain largely unanswered to stomata was roughly half that for helium diffusion across include: what pathways are taken by water after it exits the an amphistomatous leaf, noted that this suggested sites of bundle sheath surrounding the xylem, where does the bulk of evaporation were deep within the leaf and close to the evaporation occur in the leaf, what are the mechanisms by vascular system. He also noted the apparently high resistance Correspondence: T. N. Buckley. Fax: +61 2 6799 2239; e-mail: to long-distance water flow in the mesophyll and concluded [email protected] that most water transport outside the vascular bundles may © 2014 John Wiley & Sons Ltd 7 8 T. N. Buckley occur in the vapour phase.This contradicts the prevailing but (a) (b) (c) tacit assumption that the air spaces within the leaf do not palisade epidermis E e1 represent a physiologically significant pathway for water A P S delivery to the tissues outside the xylem. The notion was not AP p AE pursued further until Pieruschka et al. (2010) and Peak & x AG 8r E Mott (2011) suggested that vapour transport between meso- P VBBS S phyll and epidermal tissues may play a role in the regulation spongy s of stomatal movements. More recently, Rockwell et al. (2014) explored the competition between liquid and gas phase water transport in the leaf, focusing on the role of thermal energy e2 rP rS direction direction transport. That study provided strong theoretical support for of flow of flow the notion that gas phase transport can help deliver water to Figure 1. Diagram illustrating the geometry assumed in cells in the vicinity of the stomatal complex. calculating conductances of alternative water pathways away from The pathways for water movement outside the bundle the bundle sheath. (a) Geometry of cell types; rP, rS and xE are sheath thus remain uncertain, yet knowledge of those path- palisade mesophyll radius, spongy mesophyll radius and epidermal ways is more important than ever, in light of growing aware- cell length, respectively. (b) Diagram of a transverse section ness of their implications – for example, for interpretation of parallel to the direction of water flow away from the bundle sheath (BS), with epidermis (E) and palisade (P) and spongy (S) Kleaf in relation to leaf anatomy and environment (Tyree et al. mesophyll labelled. This is the overall flow path for which relative 1999; Brodribb et al. 2007; Cochard et al. 2007; Scoffoni et al. conductances are calculated in this paper. (c) Diagram of a 2008), patterns of isotopic enrichment (Barbour & Farquhar transverse section normal to the direction of water flow away from 2004) and stomatal behaviour (Buckley 2005; Kaiser & the bundle sheath; this is the view ‘seen’ by water. Total areas of

Legner 2007; Peak & Mott 2011).The overarching goal of the each cell type (AS, AP and AE; spongy, palisade and epidermis, present study was to place hypotheses about restrictions on respectively) and of the gas phase pathway (AG) correspond to the leaf water flow outside the xylem on an explicitly anatomical total areas of each colour given in the legend of panel c. Values and biophysical basis. More specifically, the objectives of this shown at right represent tissue thicknesses; tissue thickness fractions are: z (epidermis) = (e1+e2)/(e1+e2+p + s), z paper were (1) to derive a flexible analytical framework for E P (palisade) = p/(e1+e2+p + s) and zS (spongy) = s/(e1+e2+p + s). computing the conductances of alternative pathways for water transport from the bundle sheath to stomata (gas phase, symplastic and apoplastic), given the values for meas- bulk flow through the apoplast and gas diffusion through the urable anatomical and biophysical parameters; (2) to use this intercellular air spaces. This section presents derivations of framework to estimate the relative contributions of each generic expressions for the molar conductivities of each path pathway to total conductance for horizontal and vertical (water flow per unit cross-sectional area of flow, per unit water movement through various tissues; and (3) to deter- time, per unit difference in water potential; mol m−2 s−1 Pa−1). mine which parameters are most important for partitioning Each conductivity is based, either implicitly or explicitly, on a flow among pathways. finite pathlength for water movement. In the expressions derived below, membrane conductivity applies to movement across a single membrane; transcellular and apoplastic con- METHODS ductivities apply to movement across or around a single cell; Overview of approach and gas phase conductivity applies to an arbitrary, generic pathlength. In Appendix 1, these conductivities are scaled to The approach of this paper will be to derive expressions for whole-tissue conductances based on estimates of leaf ana- the molar conductivities of water movement in different tomical dimensions. pathways, then to derive expressions for molar conductances in each pathway for different flow directions and tissue types, and finally to apply parameter estimates to the latter expres- Symplastic diffusion sions to estimate how conductance is partitioned among pathways. Conductivities are derived in the following section, Water travelling symplastically will encounter two sets of and then scaled to conductances in Appendix 1, using param- membranes per cell, and it will also have to diffuse across the eter values estimated in Appendix 2 and the generic model of interior of cell. Each membrane has an osmotic water per- −1 3 −2 −1 leaf anatomy shown in Fig. 1. meability of Pm (m s ; equivalently, m m s ). The molar −2 −1 flux through one membrane, Jm (mol m s )is

Molar conductivities of alternative pathways for Pm Jm =Δψ (1) water outside the bundle sheath RT

This analysis will consider four alternative pathways, or where Δψ is the water potential difference across the mem- mechanisms, for water movement to the vicinity of stomatal brane (Pa), R is the gas constant (8.3144621 Pa m3 mol−1) and pores after exiting the bundle sheath: liquid diffusion through T is temperature in kelvins. (A list of mathematical symbols the , liquid diffusion through the apoplast, liquid is given in Table 1.) The molar hydraulic conductivity, km © 2014 John Wiley & Sons Ltd, Plant, Cell and Environment, 38, 7–22 Leaf water pathways 9

Table 1. List of mathematical quantities that appear in this paper

Description Symbol Units

2 Cross-sectional area of apoplast Aa m 2 Total area available for horizontal transport AH m 2 Area available for horizontal transport in tissue j AH,j m 2 Total area available for vertical transport AV m 2 Area available for vertical transport in tissue j AV,j m Water vapour concentration c mol m−3 −3 Water vapour concentration at vascular plane c1 mol m −3 Water vapour concentration near stomata c2 mol m Water vapour concentration difference Δc mol m−3 −3 Saturated water vapour concentration cs mol m −3 Saturated water vapour concentration at vascular plane cs1 mol m −3 Saturated water vapour concentration near stomata cs2 mol m 2 −1 Molecular diffusivity of water vapour in air Dwa m s 2 −1 Molecular diffusivity of liquid water in water Dww m s −1 Apoplastic bulk flow around one cell FaB mol s Fraction of mesophyll surface area that adjoins other cells fc – Dynamic viscosity of water η Pa s −2 −1 Water vapour flux in intercellular air spaces Jg mol m s −2 −1 Water flux into one cell Jm mol m s −2 −1 −1 Total conductivity of apoplast ka mol m s Pa −2 −1 −1 Conductivity of apoplast for bulk flow kaB mol m s Pa −2 −1 −1 Conductivity of apoplast for diffusive transport kaD mol m s Pa −2 −1 −1 Conductivity of cell interior for one cell kc mol m s Pa −2 −1 −1 Isothermal conductivity of gas phase pathway kg mol m s Pa −2 −1 −1 Anisothermal conductivity of gas phase pathway kg′ mol m s Pa −2 −1 −1 Conductivity of one km mol m s Pa −1 −1 Conductance for horizontal apoplastic flow around a cell of type j KaH,j mol s Pa −1 −1 Conductance for vertical apoplastic flow around a cell of type j KaV,j mol s Pa −1 −1 Conductance for horizontal flow through the interior of a cell of type j KcH,j mol s Pa −1 −1 Conductance for vertical flow through the interior of a cell of type j KcV,j mol s Pa −1 −1 Total conductance for horizontal gas phase flow in tissue type j KgasH,j mol s Pa −1 −1 Total conductance for vertical gas phase flow in tissue type j KgasV,j mol s Pa −2 −1 −1 Leaf hydraulic conductance Kleaf mol m s Pa −1 −1 Conductance for horizontal entry into a cell of type j KmH,j mol s Pa −1 −1 Conductance for vertical entry into a cell of type j KmV,j mol s Pa −2 −1 −1 Leaf hydraulic conductance outside xylem Kox mol m s Pa −1 −1 Total conductance for horizontal symplastic flow in tissue type j KsymH,j mol s Pa −1 −1 Total conductance for vertical symplastic flow in tissue type j KsymV,j mol s Pa Permeability of apoplast for bulk flow (used in Darcy’s Law) κ m2 Length of path across one cell interior lc m Length of gas phase pathway lg m Total horizontal liquid phase pathlength lLH m Total vertical liquid phase pathlength in tissue j lLV,j m Number of Poiseuille conduits in apoplast area Aa m – Number of cells in horizontal liquid phase flow path of tissue j nH,j – Number of cells in vertical liquid phase flow path of tissue j nV,j – Water vapour pressure p Pa Hydrostatic pressure difference apoplastic flow path ΔP Pa Porosity of cell walls pa – Total pressure of air in intercellular air spaces pi Pa −1 Osmotic water permeability of cell membranes pm ms Palisade mesophyll tissue porosity pP – Spongy mesophyll tissue porosity pS – Saturated water vapour pressure psat Pa Saturated water vapour pressure at vascular plane psat1 Pa Saturated water vapour pressure near stomata psat2 Pa Gas constant R Pa m3 mol−1 K−1 Effective Poiseuille radius of bulk flow paths Ra m Palisade mesophyll cell radius rP m Spongy mesophyll cell radius rS m Absolute temperature T K Absolute temperature at vascular plane T1 K Absolute temperature near stomata T2 K Temperature difference between vascular plane and stomata ΔT K Cell wall thickness ta m Tortuosity of flow pathways through cell walls τa – Celsius temperature Tc °C 3 −1 Molar volume of water Vw m mol Water vapour mole fraction w mol mol−1 −1 Saturated water vapour mole fraction ws mol mol Epidermal cell width xE m Ratio of apoplastic to transcellular pathlength y – Water potential at vascular plane ψ1 Pa Water potential near stomata ψ2 Pa Water potential difference Δψ Pa Fraction of leaf thickness occupied by tissue j zj –

Nominal values and ranges for parameters that varied in sensitivity analyses are given in Table 2.The generic subscript j is a placeholder for tissue type (j = E, P or S for epidermis, palisade mesophyll and spongy mesophyll, respectively). Note: area units for Kleaf and Kox represent leaf area, whereas area units for conductivities represent cross-sectional areas of flow paths.

© 2014 John Wiley & Sons Ltd, Plant, Cell and Environment, 38, 7–22 10 T. N. Buckley

−2 −1 −1 2 (mol m s Pa ), is the flux divided by the water potential the total conduit area, m·πRa , equal to the total cell wall gradient, or void area, which is the product of the porosity pa and total 2 apoplast cross-sectional area Aa. Thus, mπRa = paAa,or Pm km = . (2) π 4 2 RT m Ra = paAaRa ,so 2 pARaaa The effective permeability for the diffusive path across the FaB = ΔP. (7) 8ητ yl V cell interior equals the diffusion coefficient for water in acw water, D (m2 s−1), divided by the pathlength across the cell ww Assuming osmotic gradients in the apoplast water are neg- in the direction of flow, lc (m), so the molar hydraulic con- ligible, ΔP ≈Δψ. The apoplastic bulk flow conductivity kaB is ductivity for diffusion across the interior of one cell, kc,is the flow divided by the gradient Δψ and by the flow area Aa,or = Dww 2 kc . (3) = pRaa lRTc kaB . (8) 8ητacwyl V Apoplastic diffusion Because diffusion and advection (bulk flow) are additive, An apoplastic route will flow along the cell walls forming kaD and kaB can be combined into a single term for the molar the perimeter around each cell in the symplastic pathway. hydraulic conductivity of the apoplast, ka: ⎛ 2 ⎞ The molar conductivity for water diffusion along this =+pa DwwR a ka ⎝⎜ ⎠⎟ . (9) pathway, kaD, is similar to that for the cell interior in the τηacyl RT 8 Vw symplastic route. However, in the apoplast, the conductivity must be modified in two ways: (1) the conductivity must be Gas phase diffusion multiplied by a factor pa/τa to account for the porosity pa of the cell wall (void fraction, which is less than 1) and the The intercellular air spaces represent a conduit through tortuosity τa of the water paths through those voids (the which water can move outside the bundle sheath, in precisely ratio of actual mean path length to the shortest possible the same sense that the adjacent cells and cell walls are path length; this ratio is greater than 1); and (2) the conduits. Thus, gas phase water transport can increase the pathlength lc must be multiplied by a factor y (unitless) to total conductance for water transport outside the bundle account for the fact that the pathlength around a cell may be sheath, so it can also affect the water potential at any par- greater than that directly across the cell. (y depends on cell ticular point outside the bundle sheath. A consequence of geometry; see Appendix 1.) For an apoplastic route span- this fact is that the quantity described as ‘leaf hydraulic con- ning one cell, kaD is thus ductance’ may include a gas phase component. However, because the word ‘hydraulic’ is generally reserved for prop- pD k = aww. (4) erties involving liquid water, this word should be omitted aD τ acyl RT when describing the gas phase component. It is important to recognize that a molar conductivity for Apoplastic bulk flow vapour flux can be defined and calculated in precisely the −1 Bulk flow through the apoplast, FaB (mol s ), can be same terms as for liquid water fluxes. For liquids, molar described using Darcy’s Law: hydraulic conductivities were defined earlier as ratios of molar fluxes (with units of mol m−2 s−1) to water potential κ A F = a ΔP (5) gradients (with units of Pa). Although gas flux is convention- aB η ylcw V ally described in terms of gradients in concentration, partial

2 pressure or mole fraction, water vapour concentrations can where κ is the intrinsic permeability of the cell wall (m ), Aa also be expressed as water potential gradients, as shown is the total apoplastic cross-sectional flow area, ΔP (Pa) is the below. To compute the molar conductivity for gas phase difference in hydrostatic pressure in the apoplast across the transport within the leaf, one can begin with Fick’s first law of effective flow path, η is the dynamic viscosity (Pa s), Vw is 3 −1 diffusion, the molar volume of water (m mol ) and ylc is the apoplastic pathlength around the cell. The permeability κ can be esti- dc Dwa JDg =≈Δc (10) mated by analogy to the Hagen-Poiseuille equation, which dx lg describes laminar bulk flow through cylindrical tubes.Assum- where Jg is the water vapour flux (subscript g for gas phase) ing that the area of the apoplast available for flow can be −2 −1 (mol m s ), Dwa is the molecular diffusivity of water vapour modelled as a number m of conduits of radius Ra, the 2 −1 in air (m s ), lg is the length of the gas phase flow pathway Poiseuille flow rate is (m) and Δc is the difference in water vapour concentration −3 4 (mol m ) across that pathway. To relate concentration to mRπ a FaB = ΔP (6) water potential, it is useful to first relate water potential to 8ητ yl V acw water vapour mole fraction (w, mol mol−1), and then to relate which accounts for the tortuosity (τa) of cell wall flow paths, w to concentration.The relationship between water potential as in Eqn 4. The number of conduits is computed by setting (ψ) and w is © 2014 John Wiley & Sons Ltd, Plant, Cell and Environment, 38, 7–22 Leaf water pathways 11

RT w RT ⎛ w ⎞ ψ =≈− respectively, lg is the distance between those two locations, ln⎝ 1⎠ (11) Vwswsw V w kg(T1) means kg evaluated at T1, and Δψ = ψ1–ψ2. Equation 15 where ws is the saturation water vapour mole fraction of air is a general (not isothermal) expression for the molar con- at a temperature T (kelvins) (Nobel 1991). The approximate ductivity for intercellular water vapour diffusion. Although equality on the right hand side of Eqn 11 uses ln(x) ≈ x – 1 for this expression has somewhat limited utility because one Δψ x ≈ 1, which results in an error of less than 1% for water must supply values of , T1 and T2, it can nevertheless be potential at the evaporating site above −2.2 MPa and tem- used to assess the effect of temperature gradients on gas perature between 0 and 40 °C (not shown).To relate w and ws transport within leaves. In the Results section, Eqns 14 and Δψ to water vapour concentrations, note that w = p/Pi, where p is 15 will be compared at several values for and across a vapour pressure (Pa) and Pi is total pressure of the intercel- range of values for the difference between T1 and T2 Δ lular air (Pa). The ideal gas law (p = cRT) then implies (expressed below as T = T1–T2). w/ws =[cRT/Pi]/[csRT/Pi]=c/cs, where cs is the value of c at saturation. This gives Simulation procedures ψ ≈−RT ⎛ c ⎞ ⎝ 1.⎠ (12) For each tissue type (epidermis, palisade mesophyll and Vwsc spongy mesophyll) and flow direction (horizontal and verti- Δψ ≈Δ Therefore, c(RT/Vwcs), provided temperature does cal), the anatomical model illustrated in Fig. 1 was used to not vary along the flow path (this assumption will be relaxed scale the conductivities derived above to whole-path con- below). Solving this for Δc and applying the result to 10 gives ductances, based on the areas available for flow in each pathway, the pathlengths for flow, and for liquid phase path- =ΔDVcwa w s ψ Jg . (13) ways, the numbers of cells encountered along the pathway in lRT g each case. These calculations are explained in detail in

Finally, replacing cs with psat/RT, where psat is the saturation Appendix 1. This resulted in conductance values for three vapour pressure, and dividing the flux by the water potential parallel pathways – apoplastic, symplastic and gas phase – gradient Δψ gives the isothermal molar conductivity of the within each tissue type and for each flow direction. These −2 −1 −1 gas phase pathway to water vapour, kg (mol m s Pa ), as three parallel conductances were then summed within each tissue type and flow direction, and each of the three values = DVpwa w sat kg . (14) was expressed as a percentage of the resulting sum. Appen- lRT()2 g dix 2 describes estimation of parameter values required for It can be shown (see Appendix 3) that, if a vertical tem- these calculations, as well as ranges used in parameter sensi- perature gradient occurs between the vascular plane and the tivity analysis; the resulting values and ranges are given in epidermal plane (i.e. near stomata), then the ratio of flux to Table 2. water potential gradient between those two planes becomes To assess the sensitivity of these calculations to variation in parameters other than temperature, key parameters were D ⎛ p p ⎞ ⎛ V ψ ⎞ kkT′ = ()+−wa sat1 sat2 w 2 + 1 varied systematically within the limits given in Table 2. The gg1 Δψ ⎝ ⎠ ⎝ ⎠ (15) lRg T1 T2 RT1 effect of temperature on these calculations was then assessed where the subscripts 1 and 2 refer to the conditions near in two ways: by calculating the effect of uniform variations in the vertical centre of the leaf and near the epidermis, temperature between 0 and 40 °C on the conductivities for

Table 2. List of parameters that varied in sensitivity analyses, with nominal values and ranges

Description Symbol Units Nominal value Range

Fraction of mesophyll surface area that adjoins other cells fc – 0.275 0.15–0.40 Ratio of vertical gas to liquid phase pathlengths (palisade) lg/lLV,P – 1.43 0.9–1.96 Ratio of horizontal gas to liquid phase pathlengths (spongy) lg/lLH,S – 1.13 0.69–1.57 Ratio of vertical gas to liquid phase pathlengths (spongy) lg/lLV,S – 1.13 0.69–1.57 Effective porosity of cell walls pa/τa – 0.2 0.1–0.5 Palisade mesophyll tissue porosity pP – 0.19 0.07–0.40 Spongy mesophyll tissue porosity pS – 0.39 0.27–0.63 −1 −6 −6 Osmotic water permeability of cell membranes Pm ms 40·10 (10–70)·10 −9 −9 Effective Poiseuille radius of cell wall bulk flow paths Ra m 3·10 (1.5–10)·10 −6 −6 Palisade mesophyll cell radius rP m 6.5·10 (3–10)·10 −6 −6 Spongy mesophyll cell radius rS m 12·10 (5–19)·10 −6 −6 Cell wall thickness ta m 0.3·10 (0.2–0.4)·10 −6 −6 Epidermal cell width xE m 27·10 (11–43)·10 Epidermal tissue thickness fraction zE – 0.19 0.05–0.38 Ratio of palisade to spongy tissue thickness fractions zP/zS – 0.75 0.5–2.0

© 2014 John Wiley & Sons Ltd, Plant, Cell and Environment, 38, 7–22 12 T. N. Buckley each flow pathway,and second, by using Eqn 15 to assess how 5 vertical temperature gradients may affect the inferred gas (a) phase conductivity. This required a range of values for the 4 temperature gradient (ΔT) and water potential gradient (Δψ) from the vascular plane to the epidermis. Simulations by

(unitless) 3

Rockwell et al. (2014) suggested that ΔT could be as large as g k /

6 g 0.15 K, and that Δψ could be as large as 0.7 10 Pa. The sen- ' k sitivity analysis presented here assessed the effect of ΔT 2 o Δψ T = 5 C between 0 and 0.20 K for four values of : 0.1, 0.25, 0.5 and T = 25oC 6 1.0 10 Pa, at three baseline temperatures (5, 25 and 45 °C). 1 T = 45oC Water potential and temperature at the bundle sheath were ( = 0.5 MPa) set to be −1.5 106 Pa and 25 °C, respectively, unless otherwise 0 noted, and nominal values given in Table 2 were used for (b) other parameters. All calculations were performed in = 0.1 MPa 20 Microsoft Excel; the spreadsheet is included as Supporting = 0.25 MPa Information Table S1. = 0.50 MPa 15 = 1.0 MPa (T = 25oC) RESULTS 10 Effects of temperature gradients on gas phase conductivity 5 Gas phase conductivity ratio (actual/isothermal), phase conductivity Gas Figure 2 shows how the effective gas phase conductivity (kg′, Eqn 15; the ratio of molar gas flux to water potential gradi- 0 0.0 0.1 0.2 ent) differed from the isothermal gas phase conductivity (kg, Eqn 14) in relation to the assumed temperature decrease Temperature gradient, T /K between the vascular plane and the epidermis (ΔT), at three (vascular plane minus epidermis) temperatures (Fig. 2a) and for four values of the water poten- Figure 2. Effect of temperature difference between the vascular tial decrease from the vascular plane to the epidermis plane and the epidermis (ΔT) (the epidermis being cooler) on the ′ (Fig. 2b). For small water potential gradients (0.1 MPa), kg ratio of the effective molar conductivity for intercellular vapour was over 16 times greater than kg at a temperature gradient diffusion (kg′) to the isothermal value of that conductivity (kg), at of 0.2 K. Even for large water potential gradients (1.0 MPa), (a) three different baseline temperatures, T, and assuming a water kg′ was over 2.5 times greater than kg. The ratio of kg′/kg potential decrease of 0.5 MPa from the vascular plane to the varied by only 2.8% across a wide range of values for the epidermis, and (b) four different values of the water potential decrease from the vascular plane to the epidermis, all at a baseline water potential at the bundle sheath (from 0 to −4 MPa), and temperature of 25 °C. Note the different vertical axes in (a) and (b). by only 15% across a range of values for the temperature at the bundle sheath (5 to 45 °C) (not shown), which indicates that the ratio kg′/kg is far more sensitive to gradients in tem- perature and water potential than to the absolute magnitudes In the absence of a vertical temperature gradient of either temperature or water potential. (ΔT = 0 K) and using the smaller value for cell wall thickness These results suggest that the isothermal approximation (0.3 m), these calculations suggest that the conductance for for kg is not justified when temperature gradients greater water movement outside the bundle sheath is predominantly ∼ than 0.01 K exist between the vascular plane and epidermis. apoplastic. Apoplastic conductance was 72–88% of the total All subsequent calculations were therefore repeated at two within each tissue type and flow direction (horizontal flow, Δ values for T (0 and 0.2 K). Fig. 3a; vertical flow, Fig. 4a,b). The remainder of conduct- ance was contributed in varying proportions by symplastic Proportions of total conductance in and gas phase pathways: 5.9% symplastic and 7.3% gas phase each pathway for horizontal palisade transport (Fig. 3a), 11.4% symplastic and 13.7% gas phase for vertical palisade transport (Fig. 4a), Figures 3 and 4 show the calculated proportions of total con- and 11.6% symplastic and 16.2% gas for both horizontal and ductance for each pathway and tissue, for horizontal (Fig. 3) vertical spongy transport (Figs 3a & 4a). Using the larger and vertical transport (Fig. 4). Two sets of results are shown value for cell wall thickness (1.4 m), apoplastic transport in each case, using the nominal value for cell wall thickness of was predicted to be even more dominant, contributing over 0.3 m (Figs 3a & 4a,b) and the average value (1.4 m) from 92% of total conductance in all cases. a recent survey of 14 species by John et al. (2013) (Figs 3b & When a 0.2 K vertical gradient in temperature was 4c,d). For vertical transport, these calculations were also assumed to occur between the vascular plane and the repeated assuming a vertical temperature gradient (ΔT)of stomata, the predicted contribution of the gas phase pathway 0.2 K (Fig. 4b,d). to total vertical conductance increased greatly, to 39.2% for © 2014 John Wiley & Sons Ltd, Plant, Cell and Environment, 38, 7–22 Leaf water pathways 13

palisade and 44.0% for spongy mesophyll (Fig. 4b). These (a) ta = 0.3 mm numbers were somewhat smaller when ΔT = 0.2 K was com- 100 bined with the larger value for cell wall thickness (Fig. 4d): in this case, gas phase transport was 13.3% of the total for 75 palisade and 15.9% for spongy mesophyll, respectively.

50 Systematic parameter sensitivity analysis gas Due to the large number of input parameters and output 25 sym variables, only the strongest parameter sensitivities are apo described here and presented in Figs 5–8. The strongest 0 effect, by far, was that of the effective Poiseuille radius of (b) ta = 1.4 mm

100 100 gas 80 75 sym 60 apo % of total conductance for horizontal flow 50 40

25 20 (a) epidermal 0 0 epidermis palisade spongy 80

Figure 3. Proportions of molar conductance for horizontal water 60 transport contributed by gas phase, symplastic and apoplastic 40 pathways, in the epidermis, palisade mesophyll and spongy mesophyll. (a) Using the nominal value for cell wall thickness, ta 20 (0.3 m); (b) using the mean value of cell wall thickness reported (b) palisade, horizontal in a survey of 14 species by John et al. (2013) (1.4 m). Nominal 0 values given in Table 2 were used for other parameters. 80

60 (a) ta = 0.3 mm, ΔT = 0 (b) ta = 0.3 mm, ΔT = 0.2 K % of total conductance 100 100 40

75 75 20 (c) palisade, vertical 50 50 0

gas 25 sym 25 80 apo 60 0 t m ΔT 0 (c) ta = 1.4 mm, ΔT = 0 (d) a = 1.4 m, = 0.2 K 100 100 40

75 75 20 (d) spongy

% of total conductance for vertical flow for% of vertical total conductance 0 % of total conductance for vertical flow % of for vertical total conductance 50 50 246810 25 25 Effective Poiseuille radius of R the apoplast, a /nm 0 0 palisade spongy palisade spongy Figure 5. Effect of variation in the effective Poiseuille radius of

Figure 4. Proportions of molar conductance for vertical water cell wall bulk flow paths (Ra) on the % of total conductance in gas transport contributed by gas phase, symplastic and apoplastic phase, symplastic and apoplastic pathways for (a) horizontal flow pathways, in the palisade and spongy mesophyll. (a, b) Using the through the epidermis, (b) horizontal flow through the palisade nominal value for cell wall thickness, ta (0.3 m); (c, d) using the mesophyll, (c) vertical flow through the palisade mesophyll and mean value of ta (1.4 m) reported in a survey of 14 species by (d) flow through the spongy mesophyll. (Vertical and horizontal John et al. (2013). (a, c) Assuming no vertical temperature gradient flow are not distinguished for spongy mesophyll because % between the vascular plane and the epidermis; (b, d) assuming a conductances are identical for both directions in the absence of a temperature difference of 0.2 K (with the epidermis being cooler). vertical temperature gradient.) Nominal parameter values given in Nominal values given in Table 2 were used for other parameters. Table 2 were used for all other parameters. © 2014 John Wiley & Sons Ltd, Plant, Cell and Environment, 38, 7–22 14 T. N. Buckley

100 100

80 80

60 60 40 gas 40 gas sym 20 sym (a) epidermis apo 20 apo

0 for spongy mesophyll % of total conductance 0 80 51015 Spongy mesophyll 60 radius, r /mm % of total conductance S 40 Figure 8. Effect of variation in the radius of spongy mesophyll

20 cells (rS) on the % of total conductance in symplastic and (b) spongy apoplastic pathways for water flow through the spongy mesophyll. 0 Nominal parameter values given in Table 2 were used for all other 0.1 0.2 0.3 0.4 0.5 parameters. Effective porosity of the apoplast, p t a/ a [unitless] mesophyll cell radius (rS,Fig.8)from5to19 m caused the apoplastic share of total conductance to decline from 90 to Figure 6. Effect of variation in the effective porosity of the 59% for horizontal transport through the spongy mesophyll, τ apoplast (pa/ a) on the % of total conductance in gas phase, and from 89 to 57% for vertical transport. symplastic and apoplastic pathways for (a) horizontal flow through the epidermis and (b) flow through the spongy mesophyll. Nominal parameter values given in Table 2 were used for all other Effects of uniform temperature changes parameters. on conductivities

Responses of intracellular, apoplastic and gas phase conduc- apoplastic bulk flow pathways, Ra. This is due to the depend- tivities (kc, ka and kg, respectively) to temperature are shown ence of apoplastic bulk flow conductivity (kaB) on the square in Fig. 9a. [The response of membrane conductivity (km)is of Ra (Eqn 8). For example, variation in Ra between 1.5 and not shown because the effect of temperature on membrane 10 m caused the apoplastic proportions of conductance in permeability (Pm) in mesophyll cells is not known, yet is the spongy mesophyll to vary by nearly 60% (Fig. 5d). The likely to be substantial based on measurements in (Lee τ effective porosity of the apoplast (pa/ a, Fig. 6) also had large et al. 2005; Murai-Hatano et al. 2008); see the Discussion effects on the partitioning of conductance among pathways: section.] The temperature responses of kc and ka are similar; for example, the contribution of apoplastic pathways to total the latter response is slightly stronger due to the greater vertical conductance in the spongy mesophyll increased from effect of temperature on viscosity (η) than on the molecular τ 57 to 87% as pa/ a increased from 0.1 to 0.5. Increases in diffusivity of water in water (Dww) (Fig. 9b).The temperature epidermal cell size (xE, Fig. 7) from 11 to 43 m caused the response of kg is much stronger than those of kc and ka apoplastic share of total conductance in the epidermis to (Fig. 9a) despite the dependence of kg on the inverse square decline from 95 to 60%, with a commensurate increase in the of absolute temperature (Eqn 14) and the weaker tempera- symplastic share from 5 to 40%. Finally, increases in spongy ture dependence of the diffusion coefficient for water vapour

in air (Dwa) as compared with that of Dww. The greater

response of kg to temperature results from its dependence on 100 the saturation vapour pressure (psat) (Eqn 14). 80

60 DISCUSSION The calculations presented here provide novel insights 40 into the potential contributions of different modes of water sym 20 transport to flow outside the bundle sheath. Most notably, apo

% of total conductance these calculations predicted that apoplastic transport domi- 0 nates symplastic transport in most conditions, and that gas 20 30 40 phase pathways contributed up to 44% of total conductance Epidermal cell size, x /mm E for vertical transport. The latter percentage was strongly dependent upon the assumed vertical temperature gradient Figure 7. Effect of variation in the size of epidermal cells (xE)on the % of total conductance in symplastic and apoplastic pathways between the centre of the leaf (the vascular plane) and the for horizontal flow through the epidermis. Nominal parameter epidermis (ΔT): for ΔT = 0.2 K, gas phase conductance values given in Table 2 were used for all other parameters. was 16–44% of total vertical conductance in the spongy © 2014 John Wiley & Sons Ltd, Plant, Cell and Environment, 38, 7–22 Leaf water pathways 15

2.5 traditional conception of leaf water flow, in which water (a) moves from the xylem to sites of evaporation along liquid k g phase flow paths in order to replace water lost from cells to C) o 2.0 k c the intercellular air spaces. Those air spaces are not typically k a perceived to support a substantial fraction of water flow; 1.5 indeed, they have usually been viewed as sinks, not sources, for water in adjacent cells.The results presented here suggest that the conductance of gas phase pathways for water trans- (unitless) (relative to 25 1.0 g

k port is large enough to support a significant fraction of water , a

k flow through tissues outside the bundle sheath.The gas phase ,

c 0.5

k conductance was also predicted to increase dramatically as the leaf interior heats up relative to the epidermis as a result 0.0 p of light absorption. These results, together with the demon- (b) sat stration by Rockwell et al. (2014) that thermally driven

C) D

o ww 2.0 D vapour transfer within the leaf contributes to the apparent wa value of Kleaf, show that the traditional perspective on water 1/h movement outside the xylem in leaves clearly needs to be 1.5 revised to include the intercellular air spaces as an important conduit for water transport. (relative to 25 1.0 sat

(unitless) An interesting feature of the simulations by Rockwell et al. p

, (2014) was that, in general, evaporation should occur pri- wa

D 0.5 marily from either the ‘perivascular’ region (i.e. the bundle ,

ww sheath or other cells near the vascular bundles) or the D ‘peristomatal’ region (i.e. the surfaces of cells lining the 0.0 substomatal chamber that subtends the stomatal pore), but 0 10203040 that very little net evaporation should occur from tissues o Temperature / C lining the pathway between those two regions. It can be argued that this should be the case even under isothermal Figure 9. Effect of temperature on factors of hydraulic conditions, and that it is simply a result of the fact that the conductance for water flow away from the bundle sheath. intercellular spaces are a conduit for water transport just as (a) Molar conductivities of the transcellular (kc), apoplastic (ka) and gas phase (kg) pathways. (b) Molecular diffusivity of water in the adjacent tissues are conduits. When the air spaces are liquid water (Dww); molecular diffusivity of water vapour in air viewed in this manner, it becomes apparent that water will (Dwa): saturation water vapour pressure (psat); and the inverse of generally flow in parallel in the gas and liquid phase path- dynamic viscosity (η). All values are expressed relative to absolute ways until the two pathways converge near the stomatal pore, values at 25 °C. and that little or no net transfer should necessarily occur between the pathways along the way (i.e. little or no net evaporation or condensation should occur along the way). Δ mesophyll, whereas for T = 0, it was just 4–16% of the total. This is best understood by considering an electrical analog Although the predicted proportions of total conductance (Fig. 10). If two parallel pathways share the same ‘ground’ contributed by apoplastic and gas phase pathways varied (i.e. low-potential sink, which, in the leaf, is the air space widely in relation to the values of some key parameters below the stomatal pore), they will also share the same gra- Δ including T, the contribution of symplastic pathways was dient of potential along the way, provided the two pathways quite small across a wide range of most parameter values. accumulate resistance steadily with distance (Fig. 10a). As a These calculations suggest the tentative conclusion that result, there will in general be no potential differences to symplastic transport does not represent a major restriction drive flow between the pathways (Fig. 10b). In the leaf, this for water flow outside the bundle sheath. means that the tissues along the flow path will be in a state of dynamic water potential equilibrium with the adjacent air Vapour transport in leaves spaces, and that net evaporation will not occur along most of that path. This mirrors one of the conclusions of Rockwell The possibility that gas phase transport may contribute sub- et al. (2014), and it suggests that this particular conclusion stantially to water flow in the leaf was suggested by Boyer may be robust to differences in the distribution of radiation (1985), and more recently explored in detail by Rockwell absorption: all that matters is that the gas and liquid phase et al. (2014) in relation to its implications for heat transport pathways accumulate resistance steadily with distance. and the sites of evaporation within leaves. As noted by the Where will evaporation occur, then, according to this view? latter authors, if gas phase water transport is significant, then It will occur in relation to flow: if 15% of flow occurs as differences in the outside-xylem component (Kox) of leaf vapour diffusion from near the bundle sheath, then 15% of hydraulic conductance (Kleaf) may partly reflect differences in evaporation will occur from near the bundle sheath. Like- the properties of gas phase transport. This contrasts with the wise, if 85% of flow occurs in the liquid phase through the © 2014 John Wiley & Sons Ltd, Plant, Cell and Environment, 38, 7–22 16 T. N. Buckley

Figure 10. Diagram of parallel liquid and gas phase pathways from the bundle sheath to the substomatal chamber, either (a) excluding the possibility of evaporation at intermediate locations along the pathway, or (b) allowing it to occur by providing connections between the two pathways. If resistance accumulates in a steady fashion in each pathway, then the gradient of water potential will be the same in both pathways (e.g. from −1 to −1.6 MPa in this example). Therefore, no net flow will occur between the two pathways (arrow with ‘X’ in [b]), regardless of the relative magnitude of liquid and gas phase hydraulic resistances (RLIQ and RGAS, respectively). mesophyll and epidermis, then 85% of evaporation will occur by over an order of magnitude among published measure- from the most distal sites along those pathways. Although ments (e.g. Nobel 1991, Rezvani Moghaddam & Wilman this ignores the effect of localized reductions in temperature 1998, Hanba et al. 2001, Hanba et al. 2002, Scafaro et al. at the sites where water moves from a liquid phase pathway 2011, John et al. 2013). Across the range of values used for to a gas phase pathway, the important insight is the recogni- each of these parameters in this study, the apoplastic con- tion that the air spaces are conduits in which water will flow, tribution to total conductance varied from 14.0% of the at a rate proportional to the share of total conductance con- total (for vertical transport in spongy mesophyll, using low tributed by those pathways. This partially reconciles the values for both Ra and ta and assuming a vertical tempera- hypothesis of Boyer (1985) that much evaporation occurs ture gradient of 0.2 K) to 99.8% of the total (for horizontal from the bundle sheath, with the view that most evaporation transport in palisade mesophyll, using high values for both occurs from cells lining the substomatal chamber (e.g. Tyree parameters and no temperature gradient). This highlights & Yianoulis 1980). the importance of better knowledge of these parameters, as well as the possibility that the role of apoplastic flow may Factors affecting the distribution of conductance differ among species. among pathways The effect of uniform increases in temperature also dif- fered among pathways, with gas phase conductivity in- Sensitivity analysis identified several critical parameters that creasing much more than liquid phase conductivities. strongly influence the partitioning of conductance among Conceptually, this is because molecular diffusion is directly pathways, most notably the effective radius of hydraulic driven by differences in concentration, not by differences in pathways for bulk flow through the apoplast (Ra) and the water potential per se, and the concentration gradient (Δc) thickness of cell walls (ta). Ra is poorly known, and ta varies corresponding to a given water potential gradient (Δψ) © 2014 John Wiley & Sons Ltd, Plant, Cell and Environment, 38, 7–22 Leaf water pathways 17 increases strongly with temperature in the gas phase but not bundle sheath, and of the magnitude of vertical temperature in the liquid phase. As a result, the vapour diffusion flux per gradients within leaves. unit water potential difference (the gas phase conductivity) increases far more with temperature than does the liquid diffusion flux per unit water potential difference (the liquid ACKNOWLEDGMENTS phase conductivities).This does not necessarily imply that the The author thanks Grace John for providing anatomical gas phase pathway transports a greater share of water at data; Lawren Sack, Christine Scoffoni, Helen Bramley, higher temperatures, however, because it is likely that mem- Tarryn Turnbull, Margaret Barbour, Antonio Díaz-Espejo brane conductivity Pm also increases with temperature, as and Kevin Simonin for helpful discussions during the prepa- found in roots (e.g. Lee et al. 2005, Murai-Hatano et al. 2008). ration of this paper; and two anonymous reviewers and the The present study did not attempt to simulate responses of Associate Editor, Professor Steve Long, for constructive Pm to T because they are difficult to generalize and have not comments on earlier drafts. This work was supported by been adequately characterized in mesophyll cells. funding from the US National Science Foundation (Award # Leaf hydraulic conductance (Kleaf) has previously been 1146514), the Australian Research Council (Linkage Grant reported to increase with temperature (Sack et al. 2004; LP130101183) and the Grains Research and Development Cochard et al. 2007). These responses have been attributed Corporation. to effects on aquaporins in the bundle sheath, and on the viscosity for bulk flow in the leaf xylem. The results pre- sented here suggest that the component of Kleaf that resides REFERENCES outside of the bundle sheath should also increase strongly Agre P., Sasaki S. & Chrispeels M. (1993) Aquaporins: a family of water with temperature. This might help to explain the observa- channel proteins. American Journal of Physiology – Renal Physiology 265, tion that stomata open substantially as temperature is F461–F461. increased while the evaporative gradient is held constant Anderson C.T., Carroll A., Akhmetova L. & Somerville C. (2010) Real-time (Mott & Peak 2010): if temperature increases the conduct- imaging of cellulose reorientation during cell wall expansion in Arabidopsis roots. Plant Physiology 152, 787–796. ance for water transport to the stomatal complex, then for a Barbour M.M. & Farquhar G.D. (2004) Do pathways of water movement and given transpiration rate, warming should increase water leaf anatomical dimensions allow development of gradients in H218O potential near the guard cells, favouring stomatal opening in between veins and the sites of evaporation within leaves? Plant, Cell & Environment 27, 107–121. the steady state (Mott et al. 1997; Buckley 2005). This would Boyer J.S. (1985) Water transport. Annual Review of Plant Physiology 36, also explain why temperature has little effect on stomata 473–516. when the evaporative gradient is negligible, as also shown Brodribb T.J., Feild T.S. & Jordan G.J. (2007) Leaf maximum photosynthetic by Mott & Peak (2010), because for hydraulic conductance rate and venation are linked by hydraulics. Plant Physiology 144, 1890–1898. Buckley T.N. (2005) The control of stomata by water balance (Tansley to influence water potential, substantial water loss must be Review). The New Phytologist 168, 275–292. occurring. Peak & Mott (2011) explained the temperature Byott G.S. & Sheriff D.W. (1976) Water movement into and through Trades- effect by positing significant temperature gradients between cantia virginiana (L.) leaves: II. Liquid flow pathways and evaporative sites. Journal of Experimental 27, 634–637. the mesophyll and epidermis, combined with the hypothesis Canny M. (1986) Water pathways in wheat leaves. III. The passage of the that guard cells are hydraulically sequestered from the rest mestome sheath and the function of the suberised lamellae. Physiologia of the leaf yet coupled to vapour in the adjacent air spaces. Plantarum 66, 637–647. Canny M.J. (1990) Tansley review no. 22. What becomes of the transpiration A strong temperature dependence of Kleaf could help to stream? The New Phytologist 114, 341–368. explain the results of Mott & Peak (2010) independent of Carpita N.C. & Gibeaut D.M. (1993) Structural models of primary cell walls in the latter hypothesis, and also in situations where such gra- flowering : consistency of molecular structure with the physical prop- dients might be reduced – for example, in leaves whose erties of the walls during growth. The Plant Journal 3, 1–30. stomata are open in darkness, so there is no light absorption Chrispeels M.J. & Agre P. (1994) Aquaporins: water channel proteins of plant and animal cells. Trends in Biochemical Sciences 19, 421–425. to preferentially warm the centre of the leaf, or for adaxial Cochard H., Venisse J.S., Barigah T.S., Brunel N., Herbette S., Guilliot A., . . . stomata in amphistomatous leaves, in which large tempera- Sakr S. (2007) Putative role of aquaporins in variable hydraulic conductance ture gradients between the upper epidermis and adjacent of leaves in response to light. Plant Physiology 143, 122–133. Cosgrove D.J. (2005) Growth of the plant cell wall. Nature Reviews. Molecular mesophyll are less likely. Cell 6, 850–861. Evans J., Caemmerer S., Setchell B. & Hudson G. (1994) The relation-

ship between CO2 transfer conductance and leaf anatomy in transgenic CONCLUSION tobacco with a reduced content of Rubisco. Functional Plant Biology 21, 475–495. The calculations presented here suggest that the conductance Evert R.F., Botha C.E.J. & Mierzwa R.J. (1985) Free-space marker studies on for apoplastic water transport from the bundle sheath to the leaf of Zea mays L. Protoplasma 126, 62–73. stomata greatly exceeds that for symplastic transport, and Fahlén J. & Salmén L. (2004) Pore and matrix distribution in the fiber wall revealed by atomic force microscopy and image analysis. Biomacro- that the gas phase conductance of the intercellular air spaces molecules 6, 433–438. can be a substantial fraction of total conductance for vertical Farquhar G.D. & Raschke K. (1978) On the resistance to transpiration of the water transport if significant temperature gradients exist sites of evaporation within the leaf. Plant Physiology 61, 1000–1005. between the leaf centre and the epidermis. Further clarity on Fleischer A., O’Neill M.A. & Ehwald R. (1999) The pore size of non- graminaceous plant cell walls is rapidly decreased by borate ester cross- these matters awaits better knowledge of the hydraulic prop- linking of the pectic polysaccharide rhamnogalacturonan II. Plant erties of cell membranes and cell walls in tissues outside the Physiology 121, 829–838. © 2014 John Wiley & Sons Ltd, Plant, Cell and Environment, 38, 7–22 18 T. N. Buckley

Hanba Y.T., Miyazawa S.-I., Kogami H. & Terashima I. (2001) Effects of leaf Shatil-Cohen A., Attia Z. & Moshelion M. (2011) Bundle-sheath cell regula-

age on internal CO2 transfer conductance and in tree species tion of xylem-mesophyll water transport via aquaporins under drought having different types of shoot phenology. Functional Plant Biology 28, stress: a target of xylem-borne ABA? The Plant Journal 67, 72–80. 1075–1084. Strugger S. (1943) Der aufsteigende Saftstrom in der Pflanze. Naturwis- Hanba Y.T., Kogami H. & Terashima I. (2002) The effect of growth irradiance senschaft 31, 181–191. on leaf anatomy and photosynthesis in Acer species differing in light Tanton T. & Crowdy S. (1972) Water pathways in higher plants III. The demand. Plant, Cell & Environment 25, 1021–1030. transpiration stream within leaves. Journal of Experimental Botany 23, Holz M., Heil S.R. & Sacco A. (2000) Temperature-dependent self-diffusion 619–625. coefficients of water and six selected molecular liquids for calibration in Terashima I., Hanba Y.T., Tazoe Y., Vyas P. & Yano S. (2006) Irradiance and accurate 1H NMR PFG measurements. Physical Chemistry Chemical phenotype: comparative eco-development of sun and shade leaves in rela- Physics 2, 4740–4742. tion to photosynthetic CO2 diffusion. Journal of Experimental Botany 57, John G.P., Scoffoni C. & Sack L. (2013) Allometry of cells and tissues within 343–354. leaves. American Journal of Botany 100, 1936–1948. Thomas L.H., Forsyth V.T., Šturcová A., Kennedy C.J., May R.P.,Altaner C.M., Johnson D.L., Koplik J. & Dashen R. (1987) Theory of dynamic permeability . . . Jarvis M.C. (2013) Structure of cellulose microfibrils in primary cell walls and tortuosity in fluid-saturated porous media. Journal of Fluid Mechanics from collenchyma. Plant Physiology 161, 465–476. 176, 379–402. Tyree M., Cruiziat P., Benis M., LoGullo M. & Salleo S. (1981) The kinetics of Kaiser H. & Legner N. (2007) Localization of mechanisms involved in rehydration of detached sunflower leaves from different initial water defi- hydropassive and hydroactive stomatal responses of Sambucus nigra to dry cits. Plant, Cell & Environment 4, 309–317. air. Plant Physiology 143, 1068–1077. Tyree M., Sobrado M.A., Stratton L. & Becker P. (1999) Diversity of hydraulic Kennedy C., Šturcová A., Jarvis M. & Wess T. (2007) Hydration effects on conductance in leaves of temperate and tropical species: possible causes and spacing of primary-wall cellulose microfibrils: a small angle X-ray scattering consequences. Journal of Tropical Forest Science 11, 47–60. study. Cellulose 14, 401–408. Tyree M.T. & Yianoulis P. (1980) The site of water evaporation from sub- Koponen A., Kataja M. & Timonen J. (1996) Tortuous flow in porous media. stomatal cavities, liquid path resistances and hydroactive stomatal closure. Physical Review E 54, 406–410. Annals of Botany 45, 175–193. Lee S.H., Chung G.C. & Steudle E. (2005) Gating of aquaporins by low World Meteorological Organization (2008) Guide to Meteorological Instru- temperature in roots of chilling-sensitive cucumber and chilling-tolerant ments and Methods of Observation, WMO, Geneva, Switzerland. figleaf gourd. Journal of Experimental Botany 56, 985–995. Zwieniecki M.A., Brodribb T.J. & Holbrook N.M. (2007) Hydraulic design of McCann M.C., Wells B. & Roberts K. (1990) Direct visualization of cross-links leaves: insights from rehydration kinetics. Plant, Cell and Environment 30, in the primary plant-cell wall. Journal of Cell Science 96, 323–334. 910–921.

Massman W.J. (1998) A review of the molecular diffusivities of H2O, CO2, CH4, CO, O3, SO2, NH3, N2O, NO, and NO2 in air, O2 and N2 near STP. Received 6 March 2014; received in revised form 5 May 2014; accepted Atmospheric Environment 32, 1111–1127. for publication 7 May 2014 Meziane D. & Shipley B. (1999) Interacting determinants of specific leaf area in 22 herbaceous species: effects of irradiance and nutrient availability. Plant, Cell & Environment 22, 447–459. SUPPORTING INFORMATION Morillon R. & Chrispeels M.J. (2001) The role of ABA and the transpiration stream in the regulation of the osmotic water permeability of leaf cells. Additional Supporting Information may be found in the Proceedings of the National Academy of Sciences 98, 14138–14143. online version of this article at the publisher’s web-site: Mott K.A. & Peak D. (2010) Stomatal responses to humidity and temperature in darkness. Plant, Cell & Environment 33, 1084–1090. Table S1. Spreadsheet demonstrating the calculations shown Mott K.A., Denne F. & Powell J. (1997) Interactions among stomatal in in this paper. response to perturbations in humidity. Plant, Cell and Environment 20, 1098–1107. Murai-Hatano M., Kuwagata T., Sakurai J., Nonami H., Ahamed A., Nagasuga APPENDIX 1. CONDUCTANCES THROUGH K., . . . Okada M. (2008) Effect of low temperature on hydraulic con- EACH TISSUE TYPE AND FLOW DIRECTION ductivity of rice plants and the possible role of aquaporins. Plant and Cell Physiology 49, 1294–1305. This Appendix will derive estimates for the relative flow Nobel P.S. (1991) Physicochemical and Environmental Plant Physiology, Aca- areas of each pathway, in both horizontal and vertical direc- demic Press, London, UK. Peak D. & Mott K.A. (2011) A new, vapour-phase mechanism for stomatal tions and for each tissue type separately, in order to create responses to humidity and temperature. Plant, Cell & Environment 34, 162– expressions for the conductances of each pathway. These 178. areas will be estimated using the conceptual model shown in Pieruschka R., Huber G. & Berry J.A. (2010) Control of transpiration by radia- tion. Proceedings of the National Academy of Sciences 107, 13372–13377. Fig. 1, in which a leaf consists of epidermal, palisade and Rezvani Moghaddam P. & Wilman D. (1998) Cell wall thickness and cell spongy layers whose thicknesses are fractions zE, zP and zS of dimensions in plant parts of eight forage species. The Journal of Agricultural total leaf thickness. Palisade cells will be modelled as cylin- Science 131, 59–67. ders with spherical ends (capsules) with height eight times Rockwell F., Holbrook N.M. & Stroock A.D. (2014) The competition between liquid and vapor transport in transpiring leaves. Plant Physiology 164, 1741– the radius rP, and spongy cells will be modelled as spheres 1758. doi: http://dx.doi.org/10.1104/pp.1114.236323. with radius rS (e.g. Nobel 1991, Evans et al. 1994). Epidermal Sack L., Streeter C.M. & Holbrook N.M. (2004) Hydraulic analysis of water cells will be modelled as cubes with length xE. In what follows, flow through leaves of sugar maple and red oak. Plant Physiology 134, 1824–1833. the subscripts H and V will be used to denote horizontal and Sack L., Tyree M.T. & Holbrook N.M. (2005) Leaf hydraulic architecture vertical transport, respectively. correlates with regeneration irradiance in tropical rainforest trees. The New Phytologist 167, 403–413. Horizontal liquid phase transport (parallel to Scafaro A.P.,Von Caemmerer S., Evans J.R. & Atwell B.J. (2011) Temperature the epidermis) response of mesophyll conductance in cultivated and wild Oryza species with contrasting mesophyll cell wall thickness. Plant, Cell & Environment 34, The area available for horizontal gas phase transport through 1999–2008. the palisade is zP·pP·AH, where pP is the porosity of the pali- Scoffoni C., Pou A., Aasamaa K. & Sack L. (2008) The rapid light response of leaf hydraulic conductance: new evidence from two experimental methods. sade tissue and AH is the total area of a transverse leaf Plant, Cell and Environment 31, 1803–1812. section. For gas transport through the spongy tissue, the area

© 2014 John Wiley & Sons Ltd, Plant, Cell and Environment, 38, 7–22 Leaf water pathways 19

is zS·pS·AH. The total cross-sectional area of epidermal cells Finally, the conductance for horizontal apoplastic flow is AH,E = zEAH. The cross-sectional areas of palisade and around a single cell of type j, KaH,j, is the product of apoplastic spongy mesophyll cells (AH,P and AH,S, respectively) are (1 – conductivity and apoplastic flow area, or pP)zP·AH and (1 – pS)zS·AH, respectively. = (A7) The projected areas AH,E, AH,S and AH,P will now be related KAtkraH,, P121., H P a a P to the actual areas relevant to transmembrane, transcellular KAtkraH,, S= 2 H S a a S (A8) and apoplastic flow in the horizontal direction. A large frac- tion of mesophyll surface area is not available for liquid and phase symplastic water flow between cells because it adjoins KAtkx= 4 . (A9) air spaces rather than other cells and is typically appressed aH,, E H E a a E with (e.g. Evans et al. 1994, Terashima et al. Horizontal gas phase transport 2006). The remaining area is a fraction, fc, of mesophyll surface area. The ratio of cell surface area to horizontal proj- The areas for horizontal gas diffusion through the palisade ected cross-sectional area is easily shown to be 3.32 for pali- and spongy mesophyll were derived earlier as zP·pP·AH and sade cells and 4 for spongy cells with the geometry assumed zS·pS·AH, respectively, so the horizontal gas phase conduc- here and shown in Fig. 1. Because water moving away from tances for palisade and spongy tissues are the bundle sheath encounters only half of this surface area

(the other half is at the far side of the cell), the area for KzpAkgasH, P= P P H g (A10) transmembrane flow in the horizontal direction is 1.66f A c H,P and for palisade cells and 2fcAH,S for spongy cells. For epidermal cells, this area is simply AH,E. The areas for transcellular flow KzpAkgasH, S= S S H g. (A11) are simply the projected cross-sectional areas (AH,P, AH,S and AH,E); in practice, cellular contents will impede diffusion to Vertical liquid phase transport some degree, so these values most likely overestimate the effective transcellular flow areas. Epidermal cells do not participate in vertical transport, so it The area for apoplastic flow is approximately equal to the is only necessary to compute areas for vertical transport product of cell wall thickness, ta, and cell circumference, and through palisade and spongy cells. The calculations below the latter, in turn, can easily be expressed in relation to proj- assume that the fraction of area occupied by airspace in a ected cross-sectional area. It is easily shown that the ratios of paradermal plane (normal to the vertical flow axis) is equal circumference to cross-sectional area are 4/xE, 1.21/rP and 2/rS to the tissue porosity, which implies that the projected cross- for epidermal, palisade and spongy cells, respectively, using sectional areas of palisade and spongy cells in this plane are the cell geometries assumed here and shown in Fig. 1. Thus, AV,P =(1–pP)AV and AV,S =(1–pS)AV, where AV is the total the apoplastic flow areas for epidermal, palisade and spongy area of a paradermal section. Using the cell geometries cells are 4AH,E·ta/xE, 1.21AH,P·ta/rP and 2AH,S·ta/rS, respectively. shown in Fig. 1, and following the logic presented earlier in These areas will now be applied to the conductivities the Horizontal liquid phase transport section, the area for derived in the main text. The conductance for horizontal transmembrane flow vertically into a single cell is 2fc AV,P for entry into a single cell of tissue type j (j = P, S and E palisade cells and 2fcAV,S for spongy cells. The transcellular for palisade, spongy and epidermal cells, respectively), KmH,j, flow areas are AV,P and AV,S, respectively.The apoplastic areas is the product of membrane conductivity km and trans- are 2AV,P·ta/rP and 2AV,S·ta/rS for palisade and spongy cells, membrane flow area, or respectively. Thus, the conductances for vertical transmem-

brane flow into a single cell of type j, KmV,j, are = KfAkmH,, P166., c H P m (A1) KfAkmV,, P= 2 c V P m (A12) KfAkmH,, S= 2 c H S m (A2) and and

KfAkmV,, S= 2 c V S m. (A13) KAkmH,, E= H E m. (A3)

The conductance for horizontal flow across a single cell of The conductances for vertical transcellular flow across a type j, KcH,j, is the product of transcellular conductivity kc and single cell of type j, KcV,j, are transcellular flow area, or KAkcV,, P= V P c (A14) = KAkcH,, P H P c, (A4) and

KAkcH,, S= H S c (A5) KAkcV,, S= V S c. (A15) and The conductances for vertical apoplastic flow around a = (A6) KAkcH,, E H E c. single cell of type j, KaV,j, are © 2014 John Wiley & Sons Ltd, Plant, Cell and Environment, 38, 7–22 20 T. N. Buckley

KAtkraV,, P= 2 V P a a P (A16) described in Appendix 1. Values and ranges for parameters that varied in sensitivity analyses are summarized in Table 2. and Physical parameters

The gas constant R is 8.3144621 Pa m3 mol−1 K−1. The molar KAtkraV,, S= 2 V S a a S. (A17) −5 3 −1 volume of liquid water, Vw, is 1.8·10 m mol . Saturation

water vapour pressure, psat, is a function of temperature:

psat = 611.2·exp(17.62·Tc/(243.12 + Tc)) Pa, where Tc is in Vertical gas phase transport degrees Celsius (World Meteorological Organization 2008).The The areas available for vertical gas transport through the self-diffusion coefficient of liquid water in water, Dww,is 1.635·10−8·(T/215.05–1)2.063 m2 s−1 (Holz et al. 2000), and the dif- palisade and spongy tissues are pP·AV and pS·AV, respectively, −5 so the conductances for vertical gas transport through these fusion coefficient of water vapour in air, Dwa, is 2.178·10 ·(T/ 1.81 2 −1 tissues are 273.15) m s (Massman 1998), where T is in kelvins. A leaf temperature of 25 °C was used unless otherwise noted.

KpAkgasV, P= P V g (A18) Ratio of gas phase pathlength to liquid and phase pathlength

This ratio cannot be measured directly from anatomical KpAkgasV, S= S V g. (A19) dimensions, because the actual paths taken by vapour and liquid water cannot be deduced solely from geometrical con- siderations. One approach to estimating this ratio is to esti- Scaling from single cells to whole paths mate the ratio of the tortuosities of each pathway (tortuosity The liquid phase conductances derived earlier apply to single is the ratio of actual mean pathlength to the minimum cells, and for symplastic flow they apply to different components straight-line pathlength). Studies on the tortuosity of flow of the flow pathway within each cell: that is through a mem- paths in porous materials have generally concluded that tor- brane and across the cell. This section will scale these conduct- tuosity increases strongly as porosity decreases (e.g. Johnson ances up to whole-path values for each cell type and flow et al. 1987, Koponen et al. 1996), so a reasonable range for the direction. The total symplastic conductance across a single cell ratio of gas and liquid phase pathlengths can be estimated −1 using models of the tortuosity–porosity relationship applied of type j is (2/KmH,j +1/KcH,j) for horizontal flow or (2/KmV,j +1/ −1 to measured tissue porosities. One widely cited model KcV,j) for vertical flow.This is because two membranes and one cellular interior must be crossed in series. The total symplastic (Koponen et al. 1996) suggests tortuosity=1+0.8·(1 – poros- conductance of the entire liquid phase pathway for cell type j ity); applied to the porosity ranges given earlier, this suggests −1 −1 a range of pathlength ratios (gas/liquid) from 1.12 to 1.65 for is KsymH,j = nH,j ·(2/KmH,j +1/KcH,j) for horizontal flow or −1 −1 palisade and from 0.86 to 1.30 for spongy tissues, with ratios KsymV,j = nV,j ·(2/KmV,j +1/KcV,j) for vertical flow, where nH,j and of 1.43 and 1.13, respectively, corresponding to the mean nV,j are the numbers of cells of type j encountered along the horizontal and vertical flow paths, respectively. The apoplastic values of pP and pS given earlier. Nominal values for the ratio of gas to liquid phase pathlength in palisade and spongy conductance along a single cell of type j is simply KaH,j (horizon- tissues were thus taken as 1.43 and 1.13, respectively. To tal) or KaV,j (vertical), and the total apoplastic conductance of accommodate uncertainty in the tortuosity/porosity relation- the entire liquid phase pathway for cell type j is KaH,j/nH,j (hori- ship, the ranges given above were doubled (i.e. 0.90–1.96 for zontal) or KaV,j/nV,j (vertical). The number of cells of each type encountered in the horizon- palisade and 0.69–1.57 for spongy mesophyll). Gas phase pathlengths in both horizontal and vertical directions for tal flow path, nH,j, can be calculated from cell dimensions and the spongy mesophyll, and for the horizontal direction in pali- total horizontal distance along the liquid phase flow axis, lLH: sade mesophyll, were computed in this manner. However, nH,E·xE = nH,P·2rP = nH,S·2rS = lLH,sonH,E = lLH/xE, nH,P = lLH/2rP because vertical gas pathways in palisade are not tortuous and nH,S = lLH/2rS. Similarly, nV,P is the vertical liquid phase flow due to the parallel nature of palisade cells, the ratio of liquid pathlength in palisade (lLV,P) divided by the height of a palisade to gas phase pathlengths was assumed to be 1.0 for vertical cell (8rP), or nV,P = lLV,P/8rP; for spongy cells, nV,S is the vertical pathways in the palisade. liquid phase flow pathlength in spongy mesophyll (lLV,S) divided by the height of a spongy cell (2rS), or nV,S = lLV,S/2rS. Cell water permeability, Pm

APPENDIX 2. ESTIMATION OF Estimates of Pm vary widely, and the distribution of Pm PARAMETER VALUES among mesophyll cells in a given leaf is commonly reported to be bimodal. For example, Morillon & Chrispeels (2001) −6 −1 This Appendix will estimate nominal values and ranges for found modes at Pm ≈ 5 and 160·10 ms with an overall the parameters used to compute the conductivities derived in mean of 69·10−6 and similar cell numbers in each mode. −6 −1 the main text, and the conductances whose calculation is Shatil-Cohen et al. (2011) found modes at Pm = 5.4·10 ms © 2014 John Wiley & Sons Ltd, Plant, Cell and Environment, 38, 7–22 Leaf water pathways 21

−6 −1 and 58·10 ms , with 90% of cells in the low-Pm mode and data).These means were used in the present study as nominal an overall mean of 10.7·10−6 ms−1. In the present study, values, and the ranges were used as upper and lower limits. 10·10−6 and 70·10−6 were used as low and high values, respec- tively, for average Pm among mesophyll cells in a leaf, with a Ratio of apoplastic to transcellular pathlength, y nominal value in the middle of this range, at 40·10−6 ms−1. Water moving around a cell that is circular in cross-section with radius r will have to travel half of the cell’s circumfer- Water-conducting fraction of mesophyll ence, or πr, whereas water moving directly across the cell will surface area, fc only have to travel a distance 2r, which suggests y = π/2 for

The fraction, fc, of mesophyll surface area that is connected to horizontal flow around palisade or spongy mesophyll cells. adjacent cells averaged 0.18 among seven annual species and However, connections between mesophyll cells will reduce 0.37 among 25 tree species in the metadata compiled from the effective cell circumference by a factor 1 − fc (where fc is the literature by Terashima et al. (2006). In the present study, the fraction of mesophyll cell surface area that is connected π 0.15 and 0.40 were used as lower and upper values, respec- to other mesophyll cells), so y =(1−fc) /2 for horizontal tively, and 0.275 was used as the nominal value. flow around mesophyll cells. For vertical flow in palisade cells, the transcellular pathlength is the cell height, or 8rP,

whereas the apoplastic pathlength is (1 − fc)πrP +6rP; thus

Cell wall thickness, ta y = ((1 − fc)π + 6)/8 for vertical flow around palisade cells. For epidermal cells, y =1. Mesophyll and epidermal cell walls are typically between 0.2 and 0.4·10−6 m in thickness (e.g. Nobel 1991, Rezvani

Moghaddam & Wilman 1998, Hanba et al. 2001, Hanba et al. Tissue thickness fractions, zE, zP and zS 2002, Scafaro et al. 2011). This range was used in the present The mean epidermal fraction, z , of total leaf thickness study, with a nominal value of 0.3·10−6 m. However, a recent E among 22 herbaceous species grown at high light and nutri- study by John et al. (2013) of 14 species found much larger ent supplementation in a study by Meziane & Shipley (1999) values for cell wall thickness,averaging 1.4·10−6 m and ranging was 0.19, with lower and upper limits of 0.05 and 0.38; the from 0.5 to 2.5·10−6 m. An alternative set of calculations was present study used these as the nominal value, lower and therefore performed using the latter mean and range. upper limit, respectively. The mean ratio of palisade to

spongy thickness (zP/zS) varied from 0.85 to 0.69 in a survey Cell radii, rP and rS of 14 species by John et al. (2013). The present study used a

nominal value of 0.75 and a range of 0.5 to 2.0 for zP/zS, and Mesophyll cell size varies widely and cannot easily be gener- calculated zS and zP thus: zS =1−zE − zP =1−zE −(zP/zS)zS, alized. John et al. (2013) reported mean transverse areas of which gives zS =(1−zE)/(1 + zP/zS) and zP =1−zE − zS. spongy and mesophyll cells across 14 species, ranging from approximately 100 to 1100 m2 for spongy cells and 200 to 1400 m2 for palisade cells. For the cell models shown in Poiseuille radius of cell wall bulk flow −6 Fig. 1, this implies radius ranges of rS = 5.6 to 18.7·10 m for pathways, Ra spongy cells and r = 3.6 to 9.6·10−6 m for palisade cells. Based P Cellulose microfibrils in primary cell walls are around 3 nm on these values, the present study used 5·10−6–19·10−6 masthe (3·10−9 m) in diameter (Thomas et al. 2013) and are spaced range for spongy cells and 3·10−6–10·10−6 m for palisade cells, approximately 20–40 nm apart (McCann et al. 1990). with nominal values taken as the middle of these ranges However, much of the space between microfibrils is occupied (r = 12·10−6 m and r = 6.5·10−6 m). S P by pectins and by cross-linking molecules such as xyloglucans (Carpita & Gibeaut 1993; Cosgrove 2005). Studies on

Epidermal cell size, xE primary walls in collenchyma suggest much closer spacing on the order of 5 nm (Kennedy et al. 2007), although this may Like mesophyll cells, epidermal cells vary widely in size. John indicate aggregation of microfibrils into larger-order struc- et al. (2013) reported mean transverse areas ranging from tures (Anderson et al. 2010; Thomas et al. 2013), as is known 2 approximately 120 to 1850 m , giving a cell length range of to occur in wood fibres, where microfibrils aggregate into −6 11 to 43·10 m for cubical cells. The present study used this bundles of roughly 15–20 nm in diameter (Fahlén & Salmén −6 range for xE, and the middle of the range (27·10 m) as the 2004). The effective pore diameter for movement of mol- nominal value of xE. ecules across the primary cell wall (into or out of cells – i.e. perpendicular to the main direction of apoplastic flow in the Palisade and spongy mesophyll tissue porosities, transpiration stream) is roughly 3 nm and this is highly dependent on pectin content (Fleischer et al. 1999).Together, pP and pS these results suggest that the effective pore diameter for In the survey of 14 species by John et al. (2013), the authors water flow through the primary wall matrix is between 3 found mean pP = 0.19 with a range of 0.07 to 0.40, and mean and 20 nm, and is more likely near the low end of this range. −9 pS = 0.39 with a range of 0.27 to 0.63 (John et al., unpublished The present study used a nominal value of 6·10 m for the © 2014 John Wiley & Sons Ltd, Plant, Cell and Environment, 38, 7–22 22 T. N. Buckley

−9 diameter (which gives Ra = 3·10 m), and lower and upper of air at a temperature T (kelvins), and that w/ws in turn is −9 −9 limits of 1.5·10 m and 10·10 m for Ra. equal to the ratio of c/cs, where c and cs are water vapour concentration under actual and saturated conditions, respec-

tively. This can be rearranged to give c in terms of ψ and cs: Effective cell wall porosity, pa/τa ψ This parameter is unknown, but is often assumed to be ≈+Vw ccs ()1. (A22) approximately 0.3 (e.g. Nobel 1991, Evans et al. 1994). RT Terashima et al. (2006) argued that a value of 0.1 was more likely, on the basis that wall thickness would have too little The concentration difference across the gas phase pathway impact on mesophyll CO2 conductance to explain observed can be expressed as differences in this conductance with wall thickness if pa/τa ⎛ cVψψcV⎞ were 0.3. Scafaro et al. (2011) suggested a value of 0.2 was Δ =−=swsw11 −22 +−() cc12 c ⎝ ⎠ ccss12, (A23) consistent with their observations of effects of wall thickness RT1 RT2 and other properties on mesophyll conductance. It is pos- where the subscripts 1 and 2 denote the beginning and end- sible, however, that the value of pa/τa relevant to flow along points of the pathway, respectively, and c is the value of c at the walls, rather than across them, is higher, because unlike s saturation. This can be rearranged to give flow across walls, flow along walls will encounter at least some paths that are approximately parallel to microfibrils.To ΔΔ=+cVsw1 ψψV w ⎛ −T1 ⎞ +−() account for this uncertainty, the present study used a nominal c 21⎝ccss 2⎠ cc ss12, (A24) RT11RT T2 value of 0.2 for pa/τa, but a wider range of 0.1 to 0.5 for sensitivity analysis. where Δψ = ψ1–ψ2. Dividing through by Δψ gives

ΔccV V ψ ⎛ T ⎞ 1 Total horizontal and vertical liquid phase =+sw1 w 2 cc −1 +−()cc. ΔΔΔψ ψψ⎝ ss12⎠ ss12 (A25) pathlengths (lLH, lLV,P, lLV,S) and total horizontal and RT11RT T2 vertical flow areas (AH, AV) Note that T1/T2 =(T2 + ΔT)/T2 =1+ΔT/T2. Simulations by These parameters do not influence the proportions of con- Rockwell et al. (2014) indicated that the vertical temperature ductance in each pathway for a given tissue and flow direc- gradient between the vascular plane and epidermis is likely tion, so they are arbitrary for the purposes of the present on the order of 0.1 K, so ΔT << T2 and thus T1/T2 ≈ 1inthe study and are included only for the sake of completeness.The term in parentheses in Eqn A25. Applying this to Eqn A25 −4 present study used values of 1·10 m for each of these and rearranging terms gives pathlengths and 5·10−8 m2 for each area. ΔccV⎛ cc− ⎞ ⎛ ψ V ⎞ =+sw1 ⎜ s12 s⎟ 2 w+ 1. ΔΔψψ⎝ ⎠ ⎝ ⎠ (A26) APPENDIX 3. MOLAR CONDUCTIVITY OF GAS RT1 RT1 PHASE TRANSPORT WITHIN LEAVES UNDER ANISOTHERMAL CONDITIONS Using the gas law (c = p/RT) to write cs1 and cs2 in terms of the corresponding saturation vapour pressures, psat1 and psat2, In the main text, the molar conductivity for gas phase trans- respectively, we have port within the leaf was derived by beginning with Fick’s first Δ ψ law of diffusion: cpVsat1 w1 ⎛ 2V w⎞ ⎛ p sat1 p sat2 ⎞ = ++1.− (A27) ΔΔψψ()2 ⎝ ⎠ ⎝ ⎠ RT1 R RT1 T1 T2 =≈Δdc Dwa JDg c, (A20) dx lg Finally, applying Eqn A27 to Eqn A21 gives

Dp V D ⎛ ψ V ⎞ ⎛ p p ⎞ where Jg is the water vapour flux (subscript g for gas phase) ′ = wa sat1 w ++wa 2 w sat1 − sat2 kg 1 . (A28) −2 −1 ()2 Δψ ⎝ ⎠ ⎝ ⎠ (mol m s ), Dwa is the molecular diffusivity of water vapour lRTg 1 lRg RT1 T1 T2 2 −1 in air (m s ), lg is the length of the gas phase flow path through the intercellular spaces (m) and Δc is the difference The first term on the right hand side of Eqn A28 is the in water vapour concentration (mol m−3) across that pathway. isothermal molar conductivity to gas diffusion in the intercel-

The molar conductivity of this pathway, kg′,isJg divided by lular air spaces, kg (Eqn 14 in the main text).Thus, Eqn A28 is the water potential difference across the pathway, Δψ,or equivalent to Eqn 15 in the main text.The present study only considered vertical, not horizontal temperature gradients. Δ Dwa c This is because there is strong reason to suppose illuminated kg′ = , (A21) lg Δψ leaves will be warmer near the leaf centre (where light is absorbed by mesophyll tissue) than near the epidermis; by The main text noted that water potential ψ is related to contrast, any assumptions regarding horizontal temperature −1 water vapour mole fraction (w, mol mol )byψ ≈ (RT/Vw)(w/ gradients would be conflated with assumptions regarding the ws − 1), where ws is the saturation water vapour mole fraction distribution of evaporating sites, which are poorly known. © 2014 John Wiley & Sons Ltd, Plant, Cell and Environment, 38, 7–22