numerical study of evaporation‐induced salt accumulation ... · the liquid water flow to cause...

23
RESEARCH ARTICLE 10.1002/2013WR015127 Numerical study of evaporation-induced salt accumulation and precipitation in bare saline soils: Mechanism and feedback Chenming Zhang 1 , Ling Li 1,2 , and David Lockington 1 1 National Centre for Groundwater Research and Training, School of Civil Engineering, University of Queensland, Brisbane, Queensland, Australia, 2 State Key Laboratory of Hydrology-Water Resources and Hydraulic Engineering, Hohai University, Nanjing, China Abstract Evaporation from bare saline soils in coastal wetlands causes salt precipitation in the form of efflorescence and subflorescence. However, it is not clear how much the precipitated salt in turn affects the water transport in the soil and hence the evaporation rate. We hypothesized that efflorescence exerts a mulching resistance to evaporation, while subflorescence reduces the pore space for water vapor to move through the soil. A numerical model is developed to simulate the transport of water, solute, and heat in the soil, and resulting evaporation and salt precipitation with the hypothesized feedback mechanism incorpo- rated. The model was applied to simulate four evaporation experiments in soil columns with and without a fixed shallow water table, and was found to replicate well the experimental observations. The simulated results indicated that as long as the hydraulic connection between the near surface soil layer and the water source in the interior soil layer exists, vaporization occurs near the surface, and salt precipitates exclusively as efflorescence. When such hydraulic connection is absent, the vaporization plane develops downward and salt precipitates as subflorescence. Being more substantial in quantity, efflorescent affects more signifi- cantly evaporation than subflorescence during the soil-drying process. Different evaporation stages based on the location of the vaporization plane and the state of salt accumulation can be identified for character- izing the process of evaporation from bare saline soils with or without a fixed shallow water table. 1. Introduction Evaporation from terrestrial surface is a fundamental component of the global hydrologic cycle and a key element of energy exchange between the earth surface and atmosphere. Better understandings on the pro- cess of evaporation from bare soils and related heat exchange at the soil-air interface would assist in improving the efficiency of irrigation in agricultural applications [Ritchie, 1972] and the measures to slow down the process of degradation and desertification in some arid and semiarid areas [Xue, 1996]. In many eco-hydrological systems, evaporation provides an important drive for transport of water (liquid and vapor), solute, and heat, which underlie the behavior and functions of these systems [Hughes et al., 2001; Carter et al., 2008; Moffett et al., 2010]. Mainly driven by the gradient of water vapor density between the near sur- face soil layer (defined as the soil layer closest to the soil surface where the soil properties are practically measurable. It is usually on the milimeter scale and is abbreviated as NSL, hereafter) and atmosphere, the evaporation process involves phase change of liquid water to water vapor and their movement in the soil prior to vapor leaving the soil surface. When the NSL is fully saturated with freshwater, the evaporation rate is largely determined by the solar radiation level, wind velocity, humidity, and temperature in the air above the soil [Penman, 1948; Lehmann and Or, 2009, Ren et al., 2012]. When the NSL is not fully saturated, the evaporation rate becomes dependent on local soil conditions, including soil water content (saturation), sol- ute (salt) concentration, and temperature [e.g., Wilson et al., 1994; Fujimaki et al., 2006; Kollet et al., 2009]. Previous studies have identified that the evaporation processes from bare soils in the absence of salt are rather dependent on the condition of water table below the evaporating soil surface [e.g., Shimojima et al., 1990; Rose et al., 2005]. Various classifications on evaporation stages have been defined to describe the behaviors of evaporation resulted from different water table conditions. For the cases with a fixed shallow water table, a stable hydraulic connection is maintained between NSL and the water source in the interior soil layer (defined as the soil layer beneath the NSL, and abbreviated as ISL hereafter); and so relatively high water saturation is maintained in the NSL. The water loss due to evaporation gets compensated by liquid Key Points: The interactions between evaporation and salt precipitation Numerical model simulating the transport of water, solute, and heat Efflorescence affects more significantly evaporation than subflorescence Supporting Information: Readme Sensitivity analysis on the developed numerical model Correspondence to: C. Zhang, [email protected] Citation: Zhang, C., L. Li, and D. Lockington (2014), Numerical study of evaporation-induced salt accumulation and precipitation in bare saline soils: Mechanism and feedback, Water Resour. Res., 50, 8084–8106, doi:10.1002/2013WR015127. Received 8 DEC 2013 Accepted 25 SEP 2014 Accepted article online 30 SEP 2014 Published online 18 OCT 2014 ZHANG ET AL. V C 2014. American Geophysical Union. All Rights Reserved. 8084 Water Resources Research PUBLICATIONS

Upload: others

Post on 13-May-2020

4 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Numerical study of evaporation‐induced salt accumulation ... · the liquid water flow to cause changes in the evaporation rate, but reduce the soil porosity and lead to obstruction

RESEARCH ARTICLE10.1002/2013WR015127

Numerical study of evaporation-induced salt accumulation andprecipitation in bare saline soils: Mechanism and feedbackChenming Zhang1, Ling Li1,2, and David Lockington1

1National Centre for Groundwater Research and Training, School of Civil Engineering, University of Queensland, Brisbane,Queensland, Australia, 2State Key Laboratory of Hydrology-Water Resources and Hydraulic Engineering, Hohai University,Nanjing, China

Abstract Evaporation from bare saline soils in coastal wetlands causes salt precipitation in the form ofefflorescence and subflorescence. However, it is not clear how much the precipitated salt in turn affects thewater transport in the soil and hence the evaporation rate. We hypothesized that efflorescence exerts amulching resistance to evaporation, while subflorescence reduces the pore space for water vapor to movethrough the soil. A numerical model is developed to simulate the transport of water, solute, and heat in thesoil, and resulting evaporation and salt precipitation with the hypothesized feedback mechanism incorpo-rated. The model was applied to simulate four evaporation experiments in soil columns with and without afixed shallow water table, and was found to replicate well the experimental observations. The simulatedresults indicated that as long as the hydraulic connection between the near surface soil layer and the watersource in the interior soil layer exists, vaporization occurs near the surface, and salt precipitates exclusivelyas efflorescence. When such hydraulic connection is absent, the vaporization plane develops downwardand salt precipitates as subflorescence. Being more substantial in quantity, efflorescent affects more signifi-cantly evaporation than subflorescence during the soil-drying process. Different evaporation stages basedon the location of the vaporization plane and the state of salt accumulation can be identified for character-izing the process of evaporation from bare saline soils with or without a fixed shallow water table.

1. Introduction

Evaporation from terrestrial surface is a fundamental component of the global hydrologic cycle and a keyelement of energy exchange between the earth surface and atmosphere. Better understandings on the pro-cess of evaporation from bare soils and related heat exchange at the soil-air interface would assist inimproving the efficiency of irrigation in agricultural applications [Ritchie, 1972] and the measures to slowdown the process of degradation and desertification in some arid and semiarid areas [Xue, 1996]. In manyeco-hydrological systems, evaporation provides an important drive for transport of water (liquid and vapor),solute, and heat, which underlie the behavior and functions of these systems [Hughes et al., 2001; Carteret al., 2008; Moffett et al., 2010]. Mainly driven by the gradient of water vapor density between the near sur-face soil layer (defined as the soil layer closest to the soil surface where the soil properties are practicallymeasurable. It is usually on the milimeter scale and is abbreviated as NSL, hereafter) and atmosphere, theevaporation process involves phase change of liquid water to water vapor and their movement in the soilprior to vapor leaving the soil surface. When the NSL is fully saturated with freshwater, the evaporation rateis largely determined by the solar radiation level, wind velocity, humidity, and temperature in the air abovethe soil [Penman, 1948; Lehmann and Or, 2009, Ren et al., 2012]. When the NSL is not fully saturated, theevaporation rate becomes dependent on local soil conditions, including soil water content (saturation), sol-ute (salt) concentration, and temperature [e.g., Wilson et al., 1994; Fujimaki et al., 2006; Kollet et al., 2009].

Previous studies have identified that the evaporation processes from bare soils in the absence of salt arerather dependent on the condition of water table below the evaporating soil surface [e.g., Shimojima et al.,1990; Rose et al., 2005]. Various classifications on evaporation stages have been defined to describe thebehaviors of evaporation resulted from different water table conditions. For the cases with a fixed shallowwater table, a stable hydraulic connection is maintained between NSL and the water source in the interiorsoil layer (defined as the soil layer beneath the NSL, and abbreviated as ISL hereafter); and so relatively highwater saturation is maintained in the NSL. The water loss due to evaporation gets compensated by liquid

Key Points:� The interactions between

evaporation and salt precipitation� Numerical model simulating the

transport of water, solute, and heat� Efflorescence affects more

significantly evaporation thansubflorescence

Supporting Information:� Readme� Sensitivity analysis on the developed

numerical model

Correspondence to:C. Zhang,[email protected]

Citation:Zhang, C., L. Li, and D. Lockington(2014), Numerical study ofevaporation-induced salt accumulationand precipitation in bare saline soils:Mechanism and feedback, WaterResour. Res., 50, 8084–8106,doi:10.1002/2013WR015127.

Received 8 DEC 2013

Accepted 25 SEP 2014

Accepted article online 30 SEP 2014

Published online 18 OCT 2014

ZHANG ET AL. VC 2014. American Geophysical Union. All Rights Reserved. 8084

Water Resources Research

PUBLICATIONS

Page 2: Numerical study of evaporation‐induced salt accumulation ... · the liquid water flow to cause changes in the evaporation rate, but reduce the soil porosity and lead to obstruction

water transported from the water table, so that the evaporation process keeps at a stable stage with evapo-ration rate close to the potential rate [Nachshon et al., 2011b]. In the cases without a fixed shallow watertable, evaporation starts from an initially water-saturated soil with no external water supply. The soil-dryingprocess occurs in three stages, corresponding with the evaporation intensity [e.g., Idso et al., 1974; Wilsonet al., 1994; Rose et al., 2005; Smits et al., 2011]. The first stage has relatively high and constant evaporationrate, as evaporation is only limited by the atmospheric demand. The second stage is featured by a sharplyfalling evaporation rate since soil water becomes a limiting factor for evaporation. During both the first andsecond stages, the vaporization plane, defined as the interface where intensive water vaporization takesplace, is located in the NSL. During the third stage, the vaporization plane moves into the ISL. A layer of drysoil with liquid saturation nearly zero and liquid water flow completely ceased (defined as dry soil layerhereafter), developed above the vaporization plane. The evaporation at the third stage has constantly lowrates as it is sustained by water vapor diffusing through the dry soil layer.

When evaporation occurs from soils with saline water, the behavior of evaporation becomes more complex asit is not only affected by water saturation, but also salt concentration/precipitate in soils [e.g., Nachshon et al.,2011a]. When the solute concentration is below the solubility, solute accumulates in the vaporization planewith increased salt concentrations in the soil water. Once the increased salt concentration reaches the solubil-ity, salt starts to precipitate in the vaporization plane. Many studies have proposed classifications to describethe evaporation process subject to both water and salt change. For example, Fujimaki et al. [2006] and Nach-shon et al. [2011b] investigated the evaporation process from a saline soil column with a fixed shallow watertable. They defined two stages to describe the entire evaporation process, i.e., a fast evaporation decline stagedue to falling osmotic potentials followed by a slow evaporation decrease stage with solid salt formed abovethe soil surface as efflorescence. Based on their experiments in a saline soil column without a fixed shallowwater table, Nachshon et al. [2011a] described the evaporation process in three stages: the first stage with agradual decrease of the evaporation rate due to both reduced osmotic potential and liquid water saturation inthe NSL, the second stage with a sharp evaporation rate decline due to both low liquid water saturation in theNSL and the formation of a salt crust on the soil surface, and the third stage with a constantly low evaporationrate as it is hampered by the thickening of the dry soil layer and the continuous salt precipitation.

As the vaporization plane moves into the ISL, salt precipitates in between the soil grains as subflorescence.The salt crystal developed in between soil grains changes the local soil porosity, and hence the soil hydrau-lic and thermodynamic properties, including permeability of liquid water, porosity, dispersivity, water reten-tion curve, and heat conductivity. Many studies have been carried out to quantify the effects of changes insoil hydraulic properties due to evaporation and dissolution, such as the statistical grain size analysis [Pandaand Lake, 1996], discrete description of mineral geometries [Steefel and Lasaga, 1994], and change of pore-size distribution estimation [Baveye and Valocchi, 1989; Clement et al., 1996]. However, most of these analy-ses and methods apply only to saturated conditions. Recently Wissmeier and Barry [2009] developed a radiusshift model to correlate changes in mineral volume due to precipitation and dissolution to changes inhydraulic properties, including porosity, volumetric water saturation, and permeability of liquid water. Thismethod involves a curve-fitting procedure to resolve the discontinuity of the soil water retention curve.

The above literature has exposed two major knowledge gaps. First, the classifications of evaporation stagesare found to be rather dependent on the understandings of the solute concentration and groundwatertable conditions in the evaporating soil, which may be unknown beforehand. Therefore, it is essential topropose a unified classification of evaporation stages regardless of soil water salinity and the shallow watertable conditions. In this study, we systematically defined the evaporation stages according to the locationsof the vaporization plane as well as the state of salt accumulation as shown in Figure 1. Water stage 1 (W1)is defined for the time period when the vaporization plane is in the NSL. During W1, if the salt solute con-centration in the NSL is below the solubility, it continues to accumulate in that layer (defined as salt changephase 1, and abbreviated as S1, hereafter). Once the solute concentration reaches the solubility, salt precipi-tates above the soil surface as efflorescence, inducing an extra resistance to evaporation (defined as saltphase 2, and abbreviated as S2, hereafter). Water stage 2 (W2) and salt phase 3 (S3) generally occurs simul-taneously, when the vaporization plane moves to the ISL and salt precipitation occurs below the NSL as sub-florescence (Figure 1b).

Another knowledge gap lies on the quantifications of the efflorescence and subflorescence during the soil-drying process, and the understanding of the influence of solid salt on the evaporation rate. Although the

Water Resources Research 10.1002/2013WR015127

ZHANG ET AL. VC 2014. American Geophysical Union. All Rights Reserved. 8085

Page 3: Numerical study of evaporation‐induced salt accumulation ... · the liquid water flow to cause changes in the evaporation rate, but reduce the soil porosity and lead to obstruction

studies of Fujimaki et al., [2006] and Nachshon et al. [2011b] have found the efflorescence could reduce theevaporation rate, their studies were only limited to the cases with a fixed shallow water table and did notdiscuss how subflorescence would affect the evaporation. Although subflorescence changes the soil proper-ties, it is not clear how significantly such disturbance in turn affects the evaporation process. However, inthe case without a fixed shallow water table, evaporation rate is high only when the vaporization planeremains in the NSL, while becomes significantly low once the vaporization plane moves underneath theNSL, even if there is no salt in the soil [Novak, 2010; Gran et al., 2011a]. Therefore, it is inferable that the saltprecipitaed near the soil surface would have more substantial effect on evaporation than the salt precipi-tated in the soil. Moreover, the location where salt precipitates as subflorescence is in the dry soil layer,where the liquid water flow is completely ceased. This implies that although subflorescence may have thepotential to disturb the soil properties and so to the liquid water flow and evaporation, such disturbancemay be trivial during evaporation. If this is true, the mathematical descriptions on feedback effect of subflor-escence to evaporation could be significantly simplified. To further investigate these issues, in this study,following Fujimaki et al. [2006] and Nachshon et al. [2011b], we first hypothesize that salt precipitated in theNSL does not affect the soil properties there, but develops above the soil surface as efflorescence, exertinga mulching resistance to evaporation. We further hypothesized that the subflorescence would not affectthe liquid water flow to cause changes in the evaporation rate, but reduce the soil porosity and lead toobstruction to the water vapor flow. To test the proposed classifications on the evaporation stages, and thehypotheses on the feedback of efflorescence and subflorescence on the evaporation process, a numericalmodel was developed to simulate the transport of water, solute, and heat in the soil, and resulting evapora-tion and salt precipitation. The model was applied to and validated by results from four previous laboratoryexperiments in soil columns: two under an isothermal condition with a fixed shallow water table conductedby Fujimaki et al. [2006] and two under a nonisothermal condition without a fixed shallow water table con-ducted by Gran et al. [2011a]. Combining results from numerical simulations and experiments, we aimed toexplore the evaporation processes and behaviors at the proposed different stages.

2. Methodology and Approach

2.1. Model DevelopmentThe numerical model considers the transport of liquid water, water vapor, solute, and heat in soils andenergy balance at the soil-air interface. The processes associated with evaporation and salt precipitationwith the hypothesized feedback mechanism are also incorporated.

Figure 1. Schematic diagrams of the process of evaporation from bare soils. (a) During W1S2, salt precipitates above the soil surface asefflorescence. Evaporation is sustained by water vapor being converted from liquid water in the vaporization plane below the salt crustand moving to the atmosphere subjected to salt resistance and aerodynamic resistance sequentially at the soil-air interface. (b) DuringW2S3, salt precipitates as subflorescence. Evaporation is maintained by water vapor at vaporization plane in the ISL moving through thedry soil layer with flux qv and further to the atmosphere subjected to surface resistance rs, salt resistance rsc, and aerodynamic resistance ra

at the soil-air interface.

Water Resources Research 10.1002/2013WR015127

ZHANG ET AL. VC 2014. American Geophysical Union. All Rights Reserved. 8086

Page 4: Numerical study of evaporation‐induced salt accumulation ... · the liquid water flow to cause changes in the evaporation rate, but reduce the soil porosity and lead to obstruction

2.1.1. Water TransportOne-dimensional mass balance equations for liquid water and water vapor moving in the soil are, respec-tively [e.g., Bear, 1972]:

@qlhl

@t52

@qlql

@z2Seql0 (1)

052@qv qv

@z1Seql0 (2)

where ql , qv , and ql0 (kg m23) are the densities of liquid water, water vapor (qv5q�v hr , where hr is the rela-tive humidity in the soil; q�v (kg m23) is the saturated water vapor density), and freshwater (freshwater is thewater with zero solute concentration, ql0 5 1000 kg m23), respectively; hl (m3 m23) is the volumetric liquidwater content (hl5hpSl , where hp (m3 m23) is porosity and Sl is liquid water saturation); t (s) is the time;z (m) is the vertical coordinate pointing upward; ql and qv (m s21) are liquid water and water vapor fluxes,respectively; Se (s21) is the local vaporization (positive) or condensation (negative) rate per unit volume ofsoil. Note that no storage change of local water vapor content is considered in equation (2). This approxima-tion is based on the assumption that water vapor content is in equilibrium with the local liquid water con-tent and the fact that water vapor content is negligible smaller than liquid water content. With theapproximation, the storage of liquid water content in equation (1) lumped both water in liquid and vaporphase [e.g., Philip and De Vries, 1957; Fujimaki et al., 2006; Novak, 2010]. Combining equations (1) and (2)leads to:

@qlhl

@t52

@ql ql

@z2@qv qv

@z(3)

The liquid water flux ql is given by Darcy’s law [e.g., Bear, 1972]:

ql52kskrql g

l@Wm

@z11

� �(4)

where ks (m2) is the intrinsic permeability; kr is the relative permeability of liquid water; g (9.8 m s22) is themagnitude of the gravitational acceleration; l (kg m21 s21) is the dynamic viscosity of liquid water, and Wm

(m) is the matric potential. The effect of temperature on the liquid water flow is small and hence neglected[Milly, 1984]. Driven by the gradient of water vapor density, the water vapor flux can be written based onFick’s law [e.g., Campbell, 1985], including an enhanced vapor flux term for intermediate water content [Phi-lip and De Vries, 1957]:

qv52Dvsha@qv

@z52Dvshaq

�v@hr

@z2Dvshahrg

@q�v@z

(5)

where Dv (m2 s21) is the molecular diffusivity of water vapor in the air; ha (m3 m23) is the volumetric air con-tent; s is the tortuosity factor (s5h7=3

a =h2p) [e.g., Millington and Quirk, 1961; Saito et al., 2006; Sakai et al.,

2009]; and g is the vapor flux enhancement factor.

2.1.2. Solute TransportThe one-dimensional nonreactive solute (salt) transport is governed by [e.g., Bear and Cheng, 2010]:

@ qlhlCð Þ@t

52f 2@ql qlC@z

1@

@zqlhl Di1Dmð Þ @C

@z

� �(6)

and

@M@t

5f (7)

where C (kg kg21) is the solute concentration in the liquid water calculated by the mass of solute per massof solution; Di (m2 s21) is the molecular diffusion coefficient; Dm (m2 s21) is the mechanical dispersion coef-ficient (Dm 5ajql=hpSlj, where a (m) is longitudinal dispersivity); M (kg m23) is the mass of solid salt per unitvolume; f (kg m23 s21) is the precipitation (positive) and dissolution rate of salt per unit volume.

Equations (6) and (7) need to be solved iteratively to determine the solute concentration C and solid saltmass per unit volume M, respectively. However, if allowing C to exceed the solubility Cs (kg kg21) and

Water Resources Research 10.1002/2013WR015127

ZHANG ET AL. VC 2014. American Geophysical Union. All Rights Reserved. 8087

Page 5: Numerical study of evaporation‐induced salt accumulation ... · the liquid water flow to cause changes in the evaporation rate, but reduce the soil porosity and lead to obstruction

assuming M as a function of C beyond the solubility to store the precipitated salt mass per unit volume, onecan combine these two equations and solve them simultaneously. For that purpose, M and C are related asfollows:

M50 C � Cs

m C2Csð Þ C > Cs

((8)

where m (kg m23) is a parameter representing the density of a virtual medium that stores precipitated salt.Figure 2 shows the relationship between salt mass per unit volume of soil and solute concentration. WhenC � Cs, salt is completely dissolved in the solute phase within the liquid water. When C > Cs , m is set to alarge value so that the majority of the excess salt is stored in the virtual medium as the precipitated saltwith a small proportion stored in the liquid water (increase of C above Cs). The error associated with theexcess solute salt stored in the liquid water can be minimized by setting m to a very large value. With therelation given by equation (8), equations (6) and (7) can be combined as:

@ qlhlC1mCð Þ@t

52@ql ql C@z

1@

@zqlhl Di1Dmð Þ @C

@z

� �(9)

2.1.3. Heat TransportTo consider the effect of soil temperature on the water vapor flow and evaporation, heat transport in thesoil is simulated in the numerical model. The heat transport equation is written as [e.g., Smits et al., 2011]:

@

@tqscs T2T 0

� �1qlclhl T2T 0

� �� 5@

@zkT@T@z

� �2@

@zL0ql qv1qlcv T2T 0

� �qv

2@

@zqlcl T2T 0� �

ql� (10)

where qs (kg m23) is the density of soil solid grains; cs, cl , and cv (J kg21 K21) are, respectively, the specificheat capacity of soil solid grains, liquid water (4200 J kg21 K21), and vapor (cv 5 2000 J kg21 K21); T is thebulk soil temperature (K); T 0 is the initial soil temperature; kT (J m21 s21 K21) is the bulk thermal conductiv-ity of soil matrix and fluid; and L0 (J kg21) is the latent heat required for vaporization(L052:501310622369:2T ). Note that minor energy generation and consumption due to soil drying andwetting, salt precipitation and dissolution, viscose liquid movement have been neglected [e.g., Nassar et al.,1992].

2.1.4. Liquid Water Transport PropertiesDifferent types of soils were used in the four experimental cases that the model simulates The functions ofsoil water retention and relative permeability were obtained from studies that conducted the experiments(see Table 1). For Masa loamy sand, the following constitutive relation between matric potential and liquidwater saturation was used [Fujimaki and Inoue, 2003]:

0 0.05 0.1 0.15 0.2 0.25 0.30

10

20

30

40

50

60

70

Solute Concentration (kg kg−1)

Sal

t mas

s pe

r un

it vo

lum

n of

soi

l (kg

m−

3 )

Salt dissolved in liquid solvent whensolute concentration is below Cs

Salt dissolved in liquid solvent thatcauses excessive density driven flow

Salt stored in the virtual media as solid phase

Cs

Figure 2. Salt mass (including solid and solute from) per unit volume of soil at different solute concentrations with the liquid water satura-tion Sl , initial porosity hp0, and parameter m, respectively, fixed at 0.2, 0.4, and 1000.

Water Resources Research 10.1002/2013WR015127

ZHANG ET AL. VC 2014. American Geophysical Union. All Rights Reserved. 8088

Page 6: Numerical study of evaporation‐induced salt accumulation ... · the liquid water flow to cause changes in the evaporation rate, but reduce the soil porosity and lead to obstruction

Sl512Slr

11 a1Wmð Þn1½ � 121=n1ð Þ1Slr 12ln 2Wm11ð Þln 2W011ð Þ

� �2( )

(11)

where Slr is the residual liquid water saturation; a1 (m21) and n1 are fitting parameters; W0 (m) is the matricpotential when the liquid water saturation is nearly zero. A bimodal function was applied to describe thesoil water retention characteristic of the Toyoura sand [Fujimaki et al., 2006]:

Sl5w

11 a1Wmð Þn1½ � 121=n1ð Þ112wð Þ

11 a2Wmð Þn2½ � 121=n2ð Þ (12)

where w, a1 (m21), a2 (m21), n1, and n2 are fitting parameters. A modified van Genuchten water retentionequation was applied to describe the behavior of the Silica sand and Aluminous Silt [Gran et al., 2011a]:

Sl512bSlr

11 a1Wmð Þn1½ � 121=n1ð Þ1bSlr (13a)

b5ln 2W0ð Þ2ln 2Wmð Þ

ln 2W0ð Þ (13b)

where a1 (m21) and n1 are fitting parameters.

It should be noted that these three soil water retention equations were modified from the original equationproposed by van Genuchten [1980]. For equations (11) and (13a), the modifications followed the approaches

Table 1. Parameters Values Applied in the Experiment and Numerical Simulation for Each Evaporation Casea

PropertiesCase ML MasaLoamy Sand

Case TSToyoura Sand

Case SSSilica Sand

Case ASAluminous Silt

Soil Hydraulic PropertiesDensity of soil solid grains qs (kg m23) 2620a 2640a 2650b 3960b

Initial porosity (no salt present) hp0 (m3 m23) 0.43a 0.44a 0.4c 0.4c

Intrinsic permeability ks (m2) 5.38 3 10213a 2.04 3 10211a 2.8 3 10211c 7.09 3 10213d

Dispersivity a (m) 1.4 3 1023a 2.3 3 1023a 0.01c 0.01Ionic diffusion coefficient Di (m2 s21) 3.89 3 10210a 7.78 3 10211a 5 3 10211c 5 3 10211c

Water retention curve equation (11)a (12)a (13a)c (13a)e

a1 (m21) 24.21a 22.65a 24c 21.63d

n1 1.37a 33.6a 8.5c 1.37d

a2 (m21) 22.14a

n2 1.84a

Slr 0.031a 0.1c 0.074d

W0 (m) 21 3 105a 25 3 104c 25 3 104e

w 0.748a

Relative permeability equation (18)a (19)a (20)c (20)c

x1 7.2a

ak 1.57a

bk 0.023a

Heat Conditions Isothermala Isothermala Nonisothermalb Nonisothermalb

Thermodynamic Propertiescs (J m23 s21) 875c 875R (W m2) 750c 300Solute PropertiesInitial salt concentration C0 (kg kg21) 0.003a 0.003a 0.007b 0.020b

Solubility Cs (kg kg21) 0.231a 0.231a 0.264b 0.264b

Aerodynamic ConditionsTemperature in the air TRef (K) 298.15a 298.15a 297.15b 297.15b

Relative humidity in the air hrRef 0.25a 0.4a 0.5b 0.5b

Initial temperature in soil column T0 (K) 298.15a 298.15a 297.15b 297.15b

Matric potential at the bottom of the column WmBot (m) 20.75a 20.4a

ra5 rH (s m21) 90a 64a 60c 60c

rHSide (s m21)O 50c 50c

Numerical ParametersTime step Dt (s) 10 10 5 5Nodal spacing below the NSL Dz (m) 0.001 0.001 0.002 0.002m (kg m23) 1000 1000 1000 1000Depth of the NSL DzNSL (m) 0.0005 0.0005 0.001 0.001

aThe data with superscript a, b, c, d, and e are from Fujimaki et al. [2006], Gran et al. [2011a, 2011b], Carsel and Parrish [1988], andFayer and Simmons [1995], respectively.

Water Resources Research 10.1002/2013WR015127

ZHANG ET AL. VC 2014. American Geophysical Union. All Rights Reserved. 8089

Page 7: Numerical study of evaporation‐induced salt accumulation ... · the liquid water flow to cause changes in the evaporation rate, but reduce the soil porosity and lead to obstruction

introduce by Mehta et al. [1994] and Fayer and Simmons [1995], respectively, in order to extend the lowersaturation limit from the residual in the original equation to nearly zero (i.e., oven dry). This extension isneeded particularly for the dry soil layer, where the liquid water saturation drops below the residual. Thebimodal equation (12) is useful for describing the soil water retention relationship of a mixed-type soil[�Simunek et al., 2003].

Based on the studies that conducted the laboratory experiment, three equations were applied to expressthe relative permeability-liquid water saturation relationship for the four types of soil. The equation givenby Campbell [1974] was applied to the Masa loamy sand [Fujimaki et al., 2006]:

kr5Sx1l (14)

where x1 is a fitting parameter. For the Toyoura sand, the relative permeability data were fitted with the fol-lowing equation [Fujimaki et al., 2006]:

kr5ak

11bk1

ak

Sl1bk(15)

where ak and bk are fitting parameters. According to Gran et al., [2011b], the relative permeability of theSilica sand and Aluminous silt was calculated according to the pore-size distribution model of Mualem[1976]:

kr5Sl2Slr

12Slr

� �0:5

12 12Sl2Slr

12Slr

� � nn21ð Þ

" # n21nð Þ8<

:9=;

2

(16)

The water retention curves and the liquid water permeabilies (i.e., kskr ) of the four soils are illustrated in Fig-ure 3. The parameters involved in equations (11–16) are listed in Table 1.

The liquid water density ql varies with the concentration of the solute (salt) as follows [e.g., Hughes and San-ford, 2004; Voss and Provost, 2010]:

ql5ql01@ql

@CC (17)

for sodium chloride, @ql@C equals 702.24 based on the units used in density and salt concentrations. The effect

of temperature on the liquid water density is negligible. The liquid water viscosity is affected by the temper-ature, solute concentration, and air pressure [e.g., Kestin et al., 1981]. Assuming a constant air pressure andlinear solute concentration-viscosity relationship, the viscosity can be calculated according to Hughes andSanford [2004]:

0 0.1 0.2 0.3 0.4 0.510

−2

10−1

100

101

102

103

104

105 (a)

Volumetric Water Content (m3 m−3)

−Mat

ric

Po

ten

tial

(m

)

Legend for (a) and (b)

Case MLCase TSCase SSCase AS

0 0.1 0.2 0.3 0.410

−18

10−16

10−14

10−12

10−10 (b)

Volumetric Water Content (m3 m−3)

Per

mea

bili

ty (

m2 )

Figure 3. Water retention curve and liquid water permeability of soils applied in the study.

Water Resources Research 10.1002/2013WR015127

ZHANG ET AL. VC 2014. American Geophysical Union. All Rights Reserved. 8090

Page 8: Numerical study of evaporation‐induced salt accumulation ... · the liquid water flow to cause changes in the evaporation rate, but reduce the soil porosity and lead to obstruction

l5239:4310248:37T2140ð Þ27

1@l@C

C (18)

where @l@C is equal to 3.25 3 1023 [Kestin et al., 1981].

2.1.5. Water Vapor Transport PropertiesThe saturated water vapor density q�v in equation (5) is calculated by the following equation based on theideal gas law:

q�v5p�v Mw

RT(19)

where Mw (0.018 kg mol21) is the molecular weight of water; R (8.314 J mol21 K21) is the ideal gas constant.The saturated water vapor pressure p�v is a function of temperature alone [e.g., Murray, 1967], i.e.,

p�v5611 exp17:27 T2273:15ð Þ

T235:85

� �(20)

The relative humidity hr in the pore space can be evaluated based on the assumption of a thermodynamicequilibrium between the liquid water and water vapor (i.e., Kelvin’s equation) [e.g., Philip and De Vries, 1957;Bresler, 1981]:

hr5expWw gMw

RT

� �(21)

where Ww is the water potential, which can be calculated according to:

Ww5Wm1Wo (22)

where Wo (m) is the osmotic potential, dependent on the solute concentration [e.g., Campbell, 1977]:

Wo52vvCMsRT

g(23)

where v is the number of ions per molecule (2 for sodium chloride) and Ms (0.0585 kg mol21) is the molecu-lar weight of sodium chloride. v (mol J21) is the osmotic coefficient and can be calculated as follows [Fuji-maki et al., 2006]:

v50:91axqlC12C

1bxqlC12C

� �px

(24)

where ax , bx , px are fitting parameters. For sodium chloride, they equal 0.0026, 0.0026, and 0.053,respectively.

The coefficient of water vapor diffusion in the air is temperature-dependent as follows [e.g., Kimball et al.,1976]:

Dv52:2931025 T273:15

� �1:75

(25)

The empirical equation developed by Cass et al. [1984] was used to describe the enhancement factor g inequation (5), which is possibly induced by the existence of liquid islands and increased temperature gra-dients in the gas phase [Philip and De Vries, 1957]:

g59:5 1 3Sl 2 8:5 exp 2 112:6ffiffiffiffi

fcp

� �Sl

� �4( )

(26)

where fc (kg kg21) is the mass fraction of clay in the soil.

The volumetric air content in the soil where water vapor moves through is subjected to changes due to theliquid water saturation as well as the salt precipitation in the soil:

ha5 12Slð Þhp5 12Slð Þ hp02hsalt� �

(27)

Water Resources Research 10.1002/2013WR015127

ZHANG ET AL. VC 2014. American Geophysical Union. All Rights Reserved. 8091

Page 9: Numerical study of evaporation‐induced salt accumulation ... · the liquid water flow to cause changes in the evaporation rate, but reduce the soil porosity and lead to obstruction

where hsalt (m3 m23) is the volumetric solid salt content (hsalt5M=qsalt , where qsalt (2165 kg m23) is the den-sity of solid salt); hp0 (m3 m23) is the initial porosity of the soil without the presence of solid salt. Note thatthe change of volumetric solid salt content and the subsequent change of volumetric air content only occurbeneath the NSL, where salt precipitates as subflorescence.

2.1.6. Soil Heat Transport PropertiesThe model of Chung and Horton [1987] is applied to describe the bulk thermal conductivity, i.e.,

kT 5 b1 1 b2hl 1 b3h0:5l (28)

where b1, b2, and b3 are fitting parameters and equal to 0.228, 22.406, and 4.909, respectively, according toChung and Horton [1987]. The assignments of these parameters by the literature are due to the fact that thesoil thermal properties are much more affected by water content than by the textures and that the thermo-properties of mineral components in the solid phase are almost in the same order of magnitude [Jury andHorton, 2003; Saito et al., 2006].

2.1.7. Boundary ConditionsAssuming no heat storage, the energy exchange in the infinitesimal air layer just above the soil surface canbe expressed by the following surface energy balance equation [e.g., Saito et al., 2006]:

G 1 L0Eql0 1 H 5 Rn (29)

where Rn (W m22) is net radiation; G (W m22) is the heat flux density to the soil (positive downward); H (W m22)is the specific sensible heat exchange between soil and atmosphere (positive upward); E (m s21) is the evapora-tion rate. With the assumption that the temperature of the infinitesimal air layer equals the temperature in theNSL, the specific sensible heat flux from the infinitesimal air layer to the atmosphere, H, can be calculated fromthe following equation [e.g., Saito et al., 2006]:

H 5caqa

rHTNSL2TRefð Þ (30)

where ca (1013 J kg21 K21) is the specific heat capacity of the dry air; qa (kg m23) is the density of dry air(qa5paMa=RTa, where Ma (0.02897 kg mol21) is the molecular weight of dry air; pa (100,325 Pa) is theatmospheric pressure; and Ta (K) is the dry air temperature); rH (s m21) is the aerodynamic resistance to theheat transfer, TNSL (K) is the temperature in the NSL, TRef (K) is the temperature at the reference point in theatmosphere. Note that equation (30) does not only apply at the soil surface at the nonisothermal cases, butalso on the side of the soil column to describe the sensible heat exchange between soil column and ambi-ent environment.

With the effects of atmospheric and soil conditions included, the evaporation rate can be calculated by[Camillo and Gurney, 1986; Fujimaki et al., 2006]:

E 5qv NSL2qvRef

ql0 ra1rs1rscð Þ (31)

where the subscript NSL and Ref refer to, respectively, values in the NSL and at the reference point; ra 5rH (s m21)is the aerodynamic resistance to water vapor transfer; rs (s m21) is the surface resistance; and rsc (s m21) is theresistance due to salt precipitation as efflorescence.

Describing the variation of evaporation rate due to water availability in the evaporating soil, the surfaceresistance is found to be one of the key parameters to determine the temporal change of evaporation rate,particularly for the cases without the presences of a fixed shallow water table. The selection of surfaceresistance equation is conducted by comparing the root mean square error of the predicted evaporationrates by the models using different surface resistance equations (e.g., equations proposed by Sun [1982],Camillo and Gurney [1986], Kondo and Saigusa [1990], and van de Griend and Owe [1994]) with regard to themeasured rates. The linear equation proposed by Camillo and Gurney [1986] was used in this study:

rs52as1bs hp02hlNSL� �

(32)

where as (s m21) and bs (s m21) are fitting parameters, and were found to be 2805 and 4140, respectively[Camillo and Gurney, 1986].

Water Resources Research 10.1002/2013WR015127

ZHANG ET AL. VC 2014. American Geophysical Union. All Rights Reserved. 8092

Page 10: Numerical study of evaporation‐induced salt accumulation ... · the liquid water flow to cause changes in the evaporation rate, but reduce the soil porosity and lead to obstruction

For the salt resistance, the empirical equation developed by Fujimaki et al. [2006] was adopted in the pres-ent model. Assuming that solid salt precipitated in the NSL is formed as effloresce above the soil surface,we can calculate rsc by:

rsc5

0 C < 0

ar ln 100C1exp 2br

ar

� �� �1br C � 0

8<: (33)

where C (kg m22) is the specific mass accumulation (per unit area) of efflorescence (C5MNSLDzNSL, whereDzNSL (m) is the depth of the NSL and MNSL (kg m23) is the mass of solid salt precipitated in the NSL);ar (s m21) and br (s m21) are fitting parameters, and equal 69 and 2104, respectively, for sodium chloride.

2.1.8. Model Integration and SolutionThe mathematical model was integrated and solved numerically based on SUTRA [Voss and Provost, 2010],an existing three-dimensional numerical model simulating coupled variably saturated density-dependentgroundwater flow and solute transport. The original SUTRA model solves the liquid water transport equa-tion (1) and solute transport equation (6). To extend SUTRA’s capability for simulating the evaporation pro-cess, the water vapor transport equation (2) and the change of porosity due to salt precipitation wasincorporated into the liquid water transport equation (1) to form water transport equation (3), which wassolved by the finite element solver provided in the original model. The salt precipitation as modeled byequation (9) can be readily simulated by SUTRA with the apparent absorption term (m) included in the sol-ute transport equation. The heat transport equation (10) was solved by a fully implicit, central finite differ-ence method [Zheng and Bennett, 2002; Bear and Cheng, 2010]. The energy balance equation (29) wasintegrated as the boundary conditions for both heat and water transport equations.

2.2. Laboratory Experiments and Application of the Model to the ExperimentsThe developed model was tested through applications to simulate laboratory experiments conducted previ-ously under various conditions (Table 1). The laboratory data for the cases with a fixed shallow water table(cases ML and TS) were from Fujimaki et al. [2006], who conducted the evaporation study based on a cylin-drical column under an isothermal condition. Masa loamy (ML) sand and Toyoura sand (TS) were used inthe two cases, respectively. The column had a size of 0.038 m in diameter and 0.052 m in depth. A water-saturated ceramic plate was placed at the bottom of the column. The ceramic plate was connected to aMariotte bottle for supplying saline water while maintaining certain negative matric potential at the bottomof the column. The matric potentials at the bottom of the soil column were 20.75 and 20.4 m for cases MLand TS, respectively. These conditions were equivalent to fixing the water table at 0.802 and 0.452 m belowsoil surface. Before the experiment started, equilibrium between the matric potential and liquid water satu-ration was achieved in the soil column. Such equilibrium was maintained by the constant matric potentialapplied at the bottom of the column. Evaporation was then intensified by blowing air over the surface ofthe column. An isothermal condition was maintained at 298 K by regulating the upper heat source andwrapping the side of the column with polystyrene foil. The evaporation rate was manually measured byweighing the column using an electronic scale at a time interval of 12 h. Both the water content and soluteconcentrations were measured over the depth during the experiments. Further details can be found in Fuji-maki et al. [2006].

The cases without a fixed shallow water table (cases SS and AS) were from Gran et al. [2011a], who con-ducted the evaporation study under nonisothermal conditions. Silica sand (SS) and Aluminous silt (AS) wereused in the two cases, respectively (Table 1). The size of the cylindrical column used in the experiments was0.144 m in diameter and 0.24 m in depth. The soil medium was initially saturated by saline solution of a dif-ferent initial concentration in each case. As the heat was introduced at the top of the column through aninfrared lamp to accelerate the evaporation rate, the temperature profile in the column varied over time,which in turn affected the evaporation rate. The evaporation rates were measured by weighing the columnusing electrical scale. The liquid water saturation and solute concentration were obtained by setting up anddismantling identical columns for measuring the profiles of both parameters at different stages of theexperiments. The temperature profiles were also measured before the dismantlement.

The numerical model has been configured to simulate the evaporation processes during the four experi-mental cases. The lengths of the calculation domain in all cases were based on the length of the soil

Water Resources Research 10.1002/2013WR015127

ZHANG ET AL. VC 2014. American Geophysical Union. All Rights Reserved. 8093

Page 11: Numerical study of evaporation‐induced salt accumulation ... · the liquid water flow to cause changes in the evaporation rate, but reduce the soil porosity and lead to obstruction

column. For the isothermal cases ML and TS, only transport equations of water and salt were solved. Thefixed matric potential at the bottom of the column was simulated using Dirichlet boundary condition. Theevaporation rate was calculated using equation (31). At the beginning of the simulation, the matric poten-tial in the column was hydrostatic maintained by the fixed matric potential at the bottom of the column;the temperature and solute concentration were uniform. For the nonisothermal cases SS and AS, the trans-port equations of water, salt, and heat are solved iteratively. No water fluxes were set on both the bottomand the side of the column. Equation (30) was applied to calculate the exchange of heat on the side of thesoil column. Surface energy balance equation (29) was adopted on the top of the soil column, which calcu-lates both the sensible and latent heat exchange of heat. Simulations were started from the hydrostatic con-dition where the soil column was water-saturated with uniform temperature and solute concentration. Therelevant parameters and their sources are listed in Table 1. The results were tested to be independent fromthe numerical parameters.

3. Results and Analysis

Generally, the model predicted reasonably well the measurements made in the experiments for all fourcases. The results are presented below for each case with the focus on the depth profiles of liquid water sat-uration, temperature and salt concentrations, and the evaporation rates at times corresponding with themeasurements. Additional simulations results, including the depth profiles of the liquid water and watervapor flux, thickness of effloresce, and change of porosity, are also analyzed to better elaborate the trans-port processes of water, solute, and heat underlying the evaporation behavior.

3.1. Evaporation From a Soil Column With a Fixed Shallow Water Table Under an IsothermalConditionFigure 4 shows the measured and simulated results for case ML. With a fixed shallow water table, the evapo-ration process remained at state W1 over the whole period of the experiment, with both salt accumulation(S1) and precipitation (S2) occurred (Figure 4d). At the beginning of the process, the water saturation andsolute concentration over the column depth largely remained at the initial levels (solid lines, Figures 4a and4b). The liquid water in the upper soil layer started to move upward to compensate the evaporating water

0.4 0.5 0.6 0.7

0

10

20

30

40

50

(a)

Saturation (−)

Dep

th (

mm

)

0 0.05 0.1 0.15 0.2Concentration (kg kg−1)

(b)

Legend for Figure (a) (b) and (c)Measurements at 11 daysCalc. right after startedCalculations at 3.5 daysCalculations at 11 days

0 5 10 15

0

10

20

30

40

50

(c)

Liquid water flux (mm day−1)

Dep

th (

mm

)

0 2 4 6 8 10 120

4

8

12

16

20

Eva

pora

tion

and

grou

ndw

ater

up

take

rat

e (m

m d

ay−

1 )

Time (day)

(d)

syad11syad5.3

W12S1S

0 2 4 6 8 10 120

0.02

0.04

0.06

0.08

0.1

Sol

id s

alt d

epth

(m

m)

Measured evaporationCalculated evaporation rate (RMSE=1.0 mm/day)Calculated groundwater uptakeCalculated solid salt depth on soil surface

Figure 4. Predicted and measured results for case ML: (a) liquid water saturation, (b) solute concentration, (c) liquid water flux across thesoil column, and (d) temporal changes of evaporation, groundwater uptake and solid salt depth on the soil surface. The evaporation stagesbased on the predicted evaporation rate are indicated in the top of Figure 4d.

Water Resources Research 10.1002/2013WR015127

ZHANG ET AL. VC 2014. American Geophysical Union. All Rights Reserved. 8094

Page 12: Numerical study of evaporation‐induced salt accumulation ... · the liquid water flow to cause changes in the evaporation rate, but reduce the soil porosity and lead to obstruction

loss with an intensified flux closer to the soil surface (solid line, Figure 4c). The highest evaporation rateoccurred at the beginning of the evaporation process due to the lowest solute concentration and theabsence of salt precipitation in the NSL (Figure 4d). As evaporation continued, the liquid water flow(upward) developed with a relatively uniform rate over the whole column. Essentially this uniform flow ratesupplied the amount of water for evaporation and decreased gradually over time (Figure 4c) as the evapo-ration weakened slowly (Figure 4d). The presences of a shallow water table allowed the loss of water due toevaporation to be readily compensated by water moving up from the source underneath. Therefore, therewas no significant change of the water saturation profile (Figure 4a).

Evaporation resulted in rising solute concentrations in the NSL during W1S1 (water stage W1 and salt stageS1), which led to reduced osmotic potential and decline of the evaporation rate. The rise of the solute con-centration also led to increase in the liquid density. This effect resulted in lower liquid water flux in theupper soil layer of high solute concentrations than that in the area below (dashed line and dash-dotted line,Figure 4c). The solute concentration reached the solubility in the middle of the second day with salt startingto precipitate. Subsequently, the evaporation process entered W1S2 where the evaporation rate declinedgradually as the salt resistance increased slowly with the salt precipitation. The upward liquid water fluxprofile remained uniform over the depth with the flux rate decreased slowly with the decline of the evapo-ration rate (Figure 4c). Since the diffusion and dispersion processes were slow, little salt was transporteddownward despite high solute concentrations near the soil surface and low solute concentrations under-neath (dashed line, Figure 4b). Over the stage of W1S2, the depth of efflorescence increased almost linearly(Figure 4d) but with no salt precipitated as subflorescence.

The measured and simulated results for case TS are shown in Figure 5. The major difference between caseML and TS lay on (1) the soil used in case TS had coarser soil grains, thus yielding a higher intrinsic perme-ability but lower liquid water saturation under the same matric potential; (2) the aerodynamic resistance incase TS (64 m s21) was lower than that in case ML (90 m s21); (3) the ionic diffusion coefficient in case TS(3.89 3 10210 m2 s21) was about a fifth of that in case ML (7.78 3 10211 m2 s21). Similar to case ML, withthe presence of a fixed shallow water table, the evaporation process remained at stage W1 throughout theexperiment/simulation, with initial accumulation of solute (S1) followed by salt precipitation (S2) in the NSL.

0 0.2 0.4 0.6

0

10

20

30

40

50

(a)

Saturation (−)

Dep

th (

mm

)

0 0.05 0.1 0.15 0.2Concentration (kg kg−1)

(b)

Legend for Figure (a) (b) and (c)Measurements at 11 daysCalc. right after startedCalculations at 3.5 daysCalculations at 11 days

0 5 10 15 20

0

10

20

30

40

50

(c)

Liquid water flux (mm day−1)

Dep

th (

mm

)

0 2 4 6 8 10 120

4

8

12

16

20

Eva

pora

tion

and

grou

ndw

ater

up

take

rat

e (m

m d

ay−

1 )

Time (day)

(d)

syad11syad5.3

W12S1S

0 2 4 6 8 10 120

0.02

0.04

0.06

0.08

0.1

Sol

id s

alt d

epth

(m

m)

Measured evaporationCalculated evaporation rate (RMSE=0.8 mm/day)Calculated groundwater uptakeCalculated solid salt depth on soil surface

Figure 5. Predicted and measured results for case TS: (a) liquid water saturation, (b) solute concentration, (c) liquid water flux across thesoil column at three time stages, and (d) temporal changes of evaporation, groundwater uptake and solid salt precipitation depth on thesoil surface. The evaporation stages based on the predicted evaporation rate are indicated in the top of Figure 5d.

Water Resources Research 10.1002/2013WR015127

ZHANG ET AL. VC 2014. American Geophysical Union. All Rights Reserved. 8095

Page 13: Numerical study of evaporation‐induced salt accumulation ... · the liquid water flow to cause changes in the evaporation rate, but reduce the soil porosity and lead to obstruction

Although the liquid water saturation in the NSL was lower than that in case ML due to a weaker capillaryforce (Figure 5a), the liquid water flow (Figure 5c) was capable of delivering sufficient liquid water requiredfor evaporation. Therefore, there was little change of saturation profile during the evaporation process (Fig-ure 5a). Due to relatively lower aerodynamic resistance in case TS, the initial evaporation rate was greaterthan that in case ML (Figure 5d). This higher initial evaporation rate, together with lower diffusivity, resultedin faster solute accumulation in the NSL and hence earlier commencement of W1S2 in case TS (day 1.2)than case ML (day 1.7).

3.2. Evaporation From a Soil Column Without a Fixed Shallow Water Table Under a NonisothermalConditionFigure 6 depicts the measured and simulated results for case SS. The evaporation process in this case cov-ered all the stages shown in Figure 1: W1S1 from the beginning to middle day 3; W1S2 from middle day 3to middle day 6; and W2S3 from middle day 6 to the end of the simulation/experiment (Figure 6h). Differentfrom the cases shown above, there was a short period of evaporation rise at the beginning of the process(solid line, Figure 6h). Being consistent with experimental observations by Norouzi Rad and Shokri [2012]and Nachshon et al. [2011a], this predicted evaporation rise was caused by the rise of soil temperature: atthe beginning of the experiment, the vaporization plane in the NSL was subjected to not only liquid waterdesaturation (Figure 6a) and solute accumulation (Figure 6b) which hampered evaporation, but also risingtemperature (Figure 6c) which enhanced evaporation. The enhancing effect overwhelmed the hamperingeffect in the early stage and led to the rise of the evaporation rate.

Also different from the cases with a fixed shallow water table, the loss of water due to evaporation in thevaporization plane in this case was not adequately compensated by upward liquid and vapor flows from ISL(Figures 6d and 6e), resulting in an immediate water table drawdown (Figure 6a). As the liquid water satura-tion decreased, the upward water flow became weakened, which in turn led to further reduction of the liq-uid water saturation as well as intensified solute accumulation (i.e., increased solute concentration). As aresult, the rising evaporation trend was reversed approximately 2 h after the start (solid line, Figure 6h).Along with the decline of the evaporation rate, the soil temperature kept on rising (solid line, Figure 6c).The reduced evaporation rate was still greater than the liquid water flow rate and thus the liquid water satu-ration in the NSL continued to decrease (solid line, Figure 6a). The liquid water desaturation vacated porespace, which allowed the development of water vapor movement (solid line and dotted line, Figure 6e).

The solute concentration reached the solubility on middle day 3, marking the commencement of stageW1S2. At this stage, the vaporization plane still remained in the NSL; thereby salt precipitated in the form ofefflorescence. Similar to the case with a fixed shallow water table, the depth of efflorescence increased line-arly with time (dashed line, Figure 6h) and no salt precipitation occurred as subflorescence (dashed line, Fig-ure 6f). Stage W1S2 ended on day 6 with the stop of efflorescence and an abrupt drop of the evaporationrate (solid line, Figure 6h). This signified the start of stage W2S3 with the dry soil layer formed and vaporiza-tion plane moved into the ISL. Evaporation was then sustained by only water vapor converted from liquidwater in the moist vaporization plane beneath the dry soil layer and then moving upward to theatmosphere.

At the beginning of stage W2S3, since the depth of the dry soil layer was thinnest, the water vapor fluxthrough the dry soil layer reached the highest rate (dashed line, Figure 6e). After that, the rate of the watervapor flux decreased due to the thickening of the dry soil layer (dash-dotted line, Figure 6e). In contrast, theliquid water flux in the dry soil layer ceased as the liquid water saturation there was nearly zero (dash-dot-ted line, Figures 6a and 6d). As the evaporation process continued, the liquid water saturation in the vapori-zation plane beneath the dry soil layer dropped below the residual, leading to the thickening (downwardexpansion) of the dry soil layer. Since the evaporation rate during stage W2S3 was nearly constant, the drysoil layer thickening process followed a linear trend over time (Figure 6g). By day 20, the dry soil layerexpanded to reach a thickness of 3 cm. During the dry soil layer thickening process, no salt precipitated asefflorescence (dashed line, Figure 6h). Instead, the solute concentration in the vaporization plane beneaththe dry soil layer increased rapidly and reached the solubility, resulting in salt precipitation in the form ofsubflorescence. The solid salt occupied the pore space in between the soil solid grains, reducing the poros-ity at the dry soil layer (Figure 6f). The extent of porosity reduction exhibited a decreasing trend with thedepth, corresponding with less salt precipitated locally as the vaporization plane moved downward. The

Water Resources Research 10.1002/2013WR015127

ZHANG ET AL. VC 2014. American Geophysical Union. All Rights Reserved. 8096

Page 14: Numerical study of evaporation‐induced salt accumulation ... · the liquid water flow to cause changes in the evaporation rate, but reduce the soil porosity and lead to obstruction

temperature profile over stage W2S3 did not change significantly as the equilibrium developed among thenet radiation, heat flux into the soil, sensible heat exchange from both the surface and side of the columnand latent heat released during evaporation.

There are three notable discrepancies between the simulated and measured results: (1) the over predictionof the temperature profile on day 6 (Figure 6c); (2) the under prediction of dry soil layer and evaporation

0 0.5 1

0

50

100

150

200

(a)

Saturation (−)

Dep

th (

mm

)

0 0.1 0.2 0.3

Concentration (kg kg−1)

(b)

Legend for Figure (a) (b) and (c)

Measurements at 74% sat.Measurements at 50% sat.Measurements at 32% sat.Calculations at 74% sat.Calculations at 50% sat.Calculations at 32% sat.

30 35 40 45 50

0

50

100

150

200

(c)

Temperature ( ° C)

Dep

th (

mm

)

0 5 10

0

50

100

150

200

(d)

Dep

th (

mm

)

Liquid water flux (mm day−1)

0 1 2 3

Water vapor flux (mm day−1)

(e)

Legend for Figure (d) (e) and (f)Calculations at 74% sat.Calculations at 55% sat.Calculations at 50% sat.Calculation at 32% sat.

0.39 0.395 0.4

0

50

100

150

200

(f)

Porosity (m3 m−3)

Dep

th (

mm

)

0 2 4 6 8 10 12 14 16 18 20 22

0

20

40

(g)

Time (day)

Dep

th (

mm

)

Legend for Figure (g)

Calculated location of the vaporization plane

0 2 4 6 8 10 12 14 16 18 20 220

8

16

24

32

Eva

pora

tion

(mm

day

−1 )

Time (day)

(h)

74% sat. 55% sat. 50% sat. 32% sat.

W1 W2S1 S2 S3

Legend for Figure (h)

0 2 4 6 8 10 12 14 16 18 20 220

0.02

0.04

0.06

0.08S

olid

sal

t dep

th (

mm

)

Measured evaporation rate

Calculated evaporation rate (RMSE=2.7 mm/day)

Calculated solid salt depth on surface

Figure 6. Predicted and measured results for case SS: (a) liquid water saturation, (b) solute concentration, (c) temperature, (d) liquid waterflux, (e) water vapor flux across the soil column, (f) porosity, (g) temporal change of vaporization plane location in the NSL, and (h) tempo-ral changes of evaporation rate and solid salt depth on the soil surface. The evaporation stages based on the predicted evaporation rateare indicated in the top of Figure 6h.

Water Resources Research 10.1002/2013WR015127

ZHANG ET AL. VC 2014. American Geophysical Union. All Rights Reserved. 8097

Page 15: Numerical study of evaporation‐induced salt accumulation ... · the liquid water flow to cause changes in the evaporation rate, but reduce the soil porosity and lead to obstruction

(Figure 6a); and (3) the under prediction of evaporation rate during stage W1 (Figure 6h). These discrepan-cies could be caused by the model not properly considering some key processes in the experiment. Forexample, (1) the variation of axial temperature may exists [Smits et al., 2011]; (2) the neglect of the albedochange on the soil surface due to efflorescence; (3) the model adopts a simple consideration of sensibleheat exchange on the side of the soil column (i.e., fixed ambient temperature and aerodynamic resistanceon the column wall); (4) the correlation between evaporation and liquid water saturation may vary amongsoils [Shahraeeni et al., 2012], while the empirically based surface resistance (i.e., equation (32)) lacks suchphysical insight. Albeit these discrepancies, the proposed model can still produce the temporal and spatialchanges of the liquid water saturation, solute concentration and temperature, and the temporal changes ofthe evaporation rate with general agreement with the measured results. Solving these issues would beessential to further improve the accuracy of the model.

Figure 7 displays the measured and simulated results for case AS. There are three major differencesbetween cases AS and SS in terms of input conditions (Table 1): (1) the aluminium silt applied in case AShad finer soil grains, which gave lower intrinsic permeability but higher liquid water saturation at the samematric potential (Figure 3); (2) the initial solute concentration was 0.02 in case AS, higher than 0.007 in caseSS; (3) the net radiation in case AS was 300 W m22, lower than 750 W m22 applied in case SS. The compari-son of the results between case AS and SS can be served as sensitivity analysis of the model in terms of soiltypes, initial solute concentration and heat input. Similar to case SS, the evaporation process of case AS cov-ered all the stages shown in Figure 1, that is, W1S1 from the beginning to day 4 with an early evaporationrise also due to increasing soil temperature; W1S2 from day 4 to day 12; and W2S3 from day 12 to the endof the simulation/experiment. Although the initial solute concentrations differed, the spatial and temporalvariations of the solute concentration (Figure 7b) were similar.

One of the distinct differences in terms of the simulated results between the two cases was that the evapo-ration rate in case AS (Figure 7h) was lower than that in case SS. This is due to the variations of net radiationinput in two cases [Gran et al., 2011a]. With less heat absorbed by the soil, less energy became available forevaporation in case AS, resulting in overall lower evaporation rates. Another major difference between thetwo cases was that case AS had longer duration of W1 in case AS; in particular, the period of W1S2 (Figure7g). Due to the prolonged W1S2 period and higher initial solute concentration, the entire depth of efflores-cence above the soil surface, around 0.45 mm (Figure 7h), was 8 times thicker than that in case SS. One ofthe reasons that case AS had prolonged stage W1S1 is that the evaporation rate is generally lower in caseAS. Another reason for prolonged stage W1S1 in case AS is that the aluminous silt could maintain higher liq-uid water saturation and higher relative permeability than the Silica sand in the unsaturated zone undersame matric potential (Figure 3). This led to water at the bottom of the column be more readily transportedupward for evaporation during stage W1 (Figure 7d) and more uniform liquid water saturation profiles (Fig-ure 7a) in case AS. Comparing with case SS, the variation of soil properties, initial solute concentration andheat input in case AS had not significantly affected the vapor transport regime (Figure 7e). However, even ifthe potential evaporation rate is lower, case AS had faster development of dry soil layer than case SS as indi-cated by the rate that vaporization plane moving downward (the penetration rates of vaporization planeduring stage W2S3 in cases SS and AS were, respectively, 1.9 and 2.6 mm/d depicted in Figures 6g and 7g).This is due to the fact that during the stages W2S3, the liquid water saturation below the vaporization planein case AS was generally lower than that in case SS, making the soil more easily to be dried (dash-dottedlines, Figures 6a and 7a). Case AS also had greater shrinkage of porosity in the dry soil layer than case SS(Figure 7f), while such shrinkage had yet significantly disrupted the vapor transport regime.

The simulated results in case AS agree generally well with the measured values, except about the earlyemergence of the dry soil layer (Figure 7a) and the nonagreement on the temperature profile (Figure 7c) atthe time point with 32% water saturation. These discrepancies are similar to the ones observed in cases SSand so can attribute to the same reasons listed above.

4. Discussions

The developed model was based on the hypothesis that salt precipitates in the NSL as efflorescence, whichdoes not change the local soil properties but exerts mulching resistance to the water vapor flow; and saltprecipitates in the ISL as subflorescence, the effect of which is mainly manifested in reduction of the soil

Water Resources Research 10.1002/2013WR015127

ZHANG ET AL. VC 2014. American Geophysical Union. All Rights Reserved. 8098

Page 16: Numerical study of evaporation‐induced salt accumulation ... · the liquid water flow to cause changes in the evaporation rate, but reduce the soil porosity and lead to obstruction

porosity and hence slowing-down of vapor movement in the soil. Through the comparison between theexperimental measurements and simulated results, it is confirmed that this hypothesis worked well for twoexperimental cases in isothermal environment with the presence of shallow water table, which differ eachother by soil types and atmospheric conditions. The model with hypothesis incorporated also shows

0 0.5 1

0

50

100

150

200

(a)

Saturation (m3 m−3)

Dep

th (

mm

)

0 0.1 0.2 0.3

Concentration (kg kg−1)

(b)

Legend for Figure (a) (b) and (c)Measurements at 32% sat.Calculations at 81% sat.Calculations at 32% sat.Calculations at 21% sat.

30 35 40 45

0

50

100

150

200

(c)

Temperature ( ° C)

Dep

th (

mm

)

0 5 10

0

50

100

150

200

(d)

Dep

th (

mm

)

Liquid water flux (mm day−1)

0 0.5 1 1.5

Water vapour flux (mm day−1)

(e)

Legend for Figure (d) (e) and (f)

Calculations at 81% sat.Calculations at 34% sat.Calculations at 32% sat.Calculations at 21% sat.

0.39 0.395 0.4

0

50

100

150

200

(f)

Porosity (m3 m−3)

Dep

th (

mm

)

0 5 10 15 20 25

0

20

40

60

(g)

Time (day)

Dep

th (

mm

)

Legend for Figure (g)

Calculated location of the vaporization front

0 5 10 15 20 250

5

10

Eva

pora

tion

(mm

day

−1 )

Time (day)

(h)

.tas%12.tas%23.tas%18 34% sat.

2W1W3S2S1S

Legend for Figure (h)

0 5 10 15 20 250

0.2

0.4

Sol

id s

alt d

epth

(m

m)

Measured evaporation rate

Calculated evaporation rate (RMSE=0.8 mm/day)

Calculated solid salt depth on surface

Figure 7. Predicted and measured results for case AS: (a) liquid water saturation, (b) solute concentration, (c) temperature, (d) liquid waterflux, (e) water vapor flux across the soil column, (f) porosity, (g) temporal change of vaporization plane location in the NSL, and (h) tempo-ral changes of evaporation rate and solid salt depth on the soil surface. The evaporation stages based on the predicted evaporation rateare indicated in the top of Figure 7h.

Water Resources Research 10.1002/2013WR015127

ZHANG ET AL. VC 2014. American Geophysical Union. All Rights Reserved. 8099

Page 17: Numerical study of evaporation‐induced salt accumulation ... · the liquid water flow to cause changes in the evaporation rate, but reduce the soil porosity and lead to obstruction

agreement with two experimental cases under nonisothermal condition without the presence of shallowwater table, which are distinct by soil types, initial solute concentration, and atmospheric conditions.

Table 2 lists the mass fractions of water and salt at the beginning and the end of each evaporation stageduring the experiments/simulations. The results show that as long as sufficient liquid water can be suppliedto the NSL, salt precipitates exclusively on the soil surface as efflorescence. For the cases with no fixed shal-low water table, the ratios of efflorescence to the total salt at the end of the simulation/experiment are 52%for case AS with finer soil grains and 21% for case SS with coarser soil grains. These ratios are higher thanratios of subflorescence to the total salt, indicating the dominance of efflorescence in salt precipitation dueto evaporation. The salt resistance formulation to account for the effect of efflorescence on evaporationwas developed by regression analysis from a soil with intermediate saturation and relatively high potentialevaporation rate [Fujimaki et al., 2006], However, the effect of efflorescence on evaporation is not onlydependent on the thickness of the efflorescence, but also potential evaporation rates and water saturationon the soil surface. For example, if water saturation in the NSL is relatively high, evaporation may not be sus-tained by vapor transporting through the mulching efflorescence, but by liquid water transporting throughthe precipitated salt layer—i.e., vaporization occurs above the efflorescence rather than below it [Sghaierand Prat, 2009]. Also it has been found that salt tends to precipitate as large crystals at high potential evap-oration rates while small homogeneous crystals at low potential evaporation rates [Nachshon et al., 2011b].How the different crystal forms affect the behavior of efflorescence and its resistance to evaporationrequires further investigation.

The ratio of subflorescence to the overall salt at the end of the simulation/experiment was 7% for case ASwith finer soil grains and 14% for case SS with coarser soil grains. This amount of subflorescence causedonly up to 2.5% reduction of porosity in the upper dry soil layer beneath the NSL (Figures 6f and 7f). It isworth noting that in the cases without a fixed shallow water table, approximately a third of the waterremained in the column at the end of the simulation/experiment, which dissolved approximately 4–6 timesthe amount of salt that precipitated as subflorescence at the end of the simulation/experiment (65% forcase SS and 41% for case AS, as list in Table 2). The solute would further precipitate as subflorescence if theevaporation continued. Assuming that all of the remaining solute in the column at the end of simulation/experiment precipitates as subflorescence within 5 mm depth of soil, the estimated reduction of porositycan rise up to 10% and 17% relative to the original porosity, respectively, for case SS and AS. Given the factthat the vaporization plane will be moving downward further if the drying process continues, the estimatedfurther reduction of porosity at the dry soil layer is still small and therefore the hypothesis would still hold.

Although the model with proposed hypothesis can properly describe the processes of the four evaporationcases, it is still not clear how much the presence of salt in pore water would disrupt the evaporation

Table 2. Mass Fractions of Water and Salt at the Beginning and End of Each Evaporation Stage for Each Casea

Mass Fraction of

Case ML Case TS

Initial W1S1 W1S2 Initial W1S1 W1S2

Water in the reservoir (%) 88 64 0 94 74 0Water in the column (%) 12 12 12 6 6 6Water evaporated (%) 0 24 88 0 20 94Salt in the reservoir as solute (%) 88 66 0 93 76 0Salt as solute in the column (%) 12 34 50 7 24 28Salt as efflorescence (%) 0 0 50 0 0 72Salt as subflorescence (%) 0 0 0 0 0 0

Mass Fraction of

Case SS Case AS

Initial W1S1 W1S2 W2S2 Initial W1S1 W1S2 W2S2

Water in the column (%) 100 68 53 32 100 66 33 38Water evaporated (%) 0 32 47 68 0 35 68 62Salt as solute in the column (%) 100 100 79 65 100 100 48 41Salt as efflorescence (%) 0 0 21 21 0 0 52 52Salt as subflorescence (%) 0 0 0 14 0 0 0 7

aFor case ML and TS, the cumulative mass of water is calculated as the sum of water evaporated and water remained in the columnat the end of the simulation; the cumulative mass of salt is calculated as the sum of salt (in both solute and solid form) at the end of sim-ulation. For cases SS and AS, the cumulative mass of water (and salt) is the sum of water (solute) in the column at the beginning of thesimulation.

Water Resources Research 10.1002/2013WR015127

ZHANG ET AL. VC 2014. American Geophysical Union. All Rights Reserved. 8100

Page 18: Numerical study of evaporation‐induced salt accumulation ... · the liquid water flow to cause changes in the evaporation rate, but reduce the soil porosity and lead to obstruction

processes with other factors unchanged, and whether the hypothesis still holds when the initial solute con-centration is extremely high. To investigate these two issues, we further simulated case AS with initial soluteconcentrations C0 (kg kg21) being 0 (named as case AS-F) and 0.25 (named as cased AS-S) in longer simula-tion time (i.e., 40 days). The simulated results of the two cases with that from original case AS (C050:02) aredisplayed in Figure 8. The deepening of vaporization plane (Figure 8e) identifies the duration of stage W2S3(or W2 for case without saline soil water), while the thickening of efflorescence (Figure 8f) shows the periodof stage (W1S2). The proposed classification of evaporation stages can also properly describe the proce-dures of evaporation in cases AS-F and AS-S (specifically, for case AS-F, stage W1 was from the beginning today 8 and stage W2 lasted from day 8 to the end of simulation; for case AS-S, stage W1S1 was very short

0 0.5 1

0

50

100

150

200

(a)

Saturation (−)

Dep

th (

mm

)

0 0.1 0.2 0.3

(b)

Concentration (kg kg−1)

30 35 40 45

(c)

Temperature ( ° C)

0.25 0.3 0.35 0.4

0

50

100

150

200

(d)

Porosity (m3 m−3)

Dep

th (

mm

)

0 5 10 15 20 25 30 35 40

0

50

100

(e)

Time (day)Vap

oriz

atio

n pl

ane

loca

tion

(mm

)

0 5 10 15 20 25 30 35 40

0

5

10(f)

Time (day)

Effl

ores

cenc

e de

pth

(mm

)

0 5 10 15 20 25 30 35 400

5

10

Eva

pora

tion

rate

(m

m d

ay−

1 )

Time (day)

(g)

Case AS−F, C0=0 Case AS, C0=0.02 Case AS−S, C0=0.25

Figure 8. Predicted results for case AS-F, AS, and AS-S: (a) the profile of liquid water saturation, (b) solute concentration, (c) temperature,(d) porosity at the end of the simulation, (e) the temporal change of vaporization plane location, (f) depth of efflorescence, and (g) evapo-ration rate. The circles in Figure 8g show the measured evaporation rate in case AS.

Water Resources Research 10.1002/2013WR015127

ZHANG ET AL. VC 2014. American Geophysical Union. All Rights Reserved. 8101

Page 19: Numerical study of evaporation‐induced salt accumulation ... · the liquid water flow to cause changes in the evaporation rate, but reduce the soil porosity and lead to obstruction

due to high initial solute concentration, stage W1S2 lasted from day 1 to day 23, stage W2S3 lasted fromday 23 to the end of simulation).

Although the variation of the initial solute concentration did not affect temperature profile significantly (Fig-ure 8c), less water got vaporized from the more saline soil water cases at the end of the simulation (87%,85%, and 78%, respectively, for cases AS-F, AS, and AS-S). Within the lost water, the majority gets vaporizedduring stage W1 (including W1S1 and W1S2 for saline soil water cases), accounting for 73%, 77%, and 82%,respectively, in cases AS-F, AS, and AS-S. Although the presence of salt in soil water lengthened the durationof stage W1, the evaporation intensity at that stage was significantly reduced, as the average evaporationrates in cases AS-F, AS, and AS-S were, respectively, 7.31, 5.08, and 3.13 mm/d. As a comparison, stage W2(or W2S3 for saline soil water cases) only takes 27%, 23%, and 18% of all lost water, respectively, in casesAS-F, AS, and AS-S, with average evaporation rates of 0.59, 0.54, and 0.63 mm/d. This suggests that the riseof solute concentration and increasing efflorescence has more significant effect to disrupt evaporation thansubflorescence, which reduces the porosity and possibly affects the soil properties.

Similar to case AS, The majority of the precipitated salt in case AS-S is in the form of efflorescence, mulchingabout 8 mm thickness above the soil surface (Figure 8f). With extremely high initial solute concentration,case AS-S also had more salt precipitated as subflorescence than case AS, with the maximum shrinkage ofporosity down to 0.27, occurred in the soil below the NSL (Figure 8a). Such high porosity reduction,appeared when soil water is extremely saline, has yet completely obstructed the water vapor flow that con-tributes to evaporation during stage W2S3. The location where subflorescence presents had liquid watersaturation almost zero, indicating no liquid water movement taking place. This again confirms the hypothe-sis that subflorescence had little effect on the liquid water flow during the soil-drying processes.

5. Concluding Remarks

In summary, a numerical model is developed to quantify the evaporation from soils saturated with salinewater, based on direct simulations of the transport of liquid water, water vapor, salt solute, and heat in soils.The model incorporates the feedback of salt precipitation on the transport processes and evaporation.Through applications to previous laboratory experiments, the model was shown to be capable of simulatingthe effects of salt accumulation and precipitation on evaporation under different groundwater conditionsas controlled by the water table. In particular, the representations of salt precipitation in the forms of efflo-rescence and subflorescence, and resulting effects on the evaporation were validated by the experimentalresults. The evaporation stages proposed in this study are systematically based on the location of the vapor-ization plane and salt accumulation. These stages mechanistically express the evaporation process fromsaline soil under conditions with and without a fixed shallow water table. Apart from the potential improve-ment to the model discussed above, it is yet to be tested whether the model assumption still stands wherelarge porosity reduction occurs (e.g., intensive salt precipitation as subflorescence may result in blockage ofwater vapor flow path). New experiments should be designed and conducted to further test the model withpotential needs to further expand its capacities. Further work can also focus on incorporating the shape ofsolid salt and the mineral hygroscopic into the model, which is essential for describing the mixed geometryof soil grains and solid salt, and associated effects on the transport of liquid water and water vapor in thesoil. The model should be further tested under conditions involving salt dissolution, which occurs in soilssubjected to flooding and infiltration of fresher water. Ultimately this model, upon further developmentand validation, may become a useful tool for studying the transport and distribution of water and salt incoastal wetlands, which underpins the services provided by these ecosystems.

Notation

C specific mass accumulation (per unit area) of efflorescence, kg m22.W0 matric potential when liquid water saturation is nearly zero, m.Wm matric potential, m.WmBot matric potential applied at the bottom of the soil column for the cases with a fixed shallow water

table, m.Wo osmotic potential, m.

Water Resources Research 10.1002/2013WR015127

ZHANG ET AL. VC 2014. American Geophysical Union. All Rights Reserved. 8102

Page 20: Numerical study of evaporation‐induced salt accumulation ... · the liquid water flow to cause changes in the evaporation rate, but reduce the soil porosity and lead to obstruction

Ww water potential, m.a dispersivity, m.a1 fitting parameter in the water retention curve, m21.a2 fitting parameter in the water retention curve, m21.b a parameter in the water retention to establish a linear relationship between ln 2Wmð Þ and Sl

below Slr .ha volumetric air content, m3 m23.hl volumetric water content, m3 m23.hl NSL volumetric water content in the NSL, m3 m23.hsalt volumetric solid salt content, m3 m23.hp porosity, m3 m23.hp0 initial porosity (no solid salt present), m3 m23.g enhancement factor of water vapor.kT bulk thermal conductivity of soil matrix and fluid, J m21 s21 K21.l dynamic viscosity of liquid pore water solution, kg m21 s21.v number of ions per molecule.qa density of dry air, kg m23.ql density of liquid water, kg m23.ql0 density of fresh liquid water, 1000 kg m23.qs density of soil solid grains, kg m23.qsalt density of solid salt, 2165 kg m23.qv density of water vapor, kg m23.qvRef density of water vapor at the reference point in the atmosphere, kg m23.qvNSL density of water vapor in the NSL, kg m23.q�v saturated water vapor density, kg m23.s tortuosity factor.v osmotic coefficient, mol J21.x1 a fitting parameter in the relative liquid water permeability.C solute concentration in liquid phase, kg kg21.C0 initial solute concentration in liquid phase, kg kg21.Cs solubility, kg kg21.Dv molecular diffusivity of water vapor in the air, m2 s21.Di ionic diffusion coefficient, m2 s21.Dm mechanical dispersion coefficient, m2 s21.E evaporation rate, m s21.G surface heat flux density into the soil, W m22.H sensible heat flux density positive upward, W m22.L0 latent heat of vaporization, J kg21.M mass of solid salt per unit volume, kg m23.MNSL mass of solid salt precipitated in the NSL, kg m23.Ma molecular weight of dry air, 0.02897 kg mol21.Ms molecular weight of sodium chloride, 0.0585 kg mol21.Mw molecular weight of water, 0.018 kg mol21.R ideal gas constant, 8.314 J mol21 K21.Rn net radiation, W m22.Sl liquid water saturation.Slr residual liquid water saturation.Se evaporation (positive) or condensation per unit volume of soil, s21.T bulk soil temperature, K.T 0 initial bulk soil temperature, K.TNSL temperature in the NSL, K.TRef temperature at the reference point in the atmosphere, K.Ta dry air temperature, K.ak a fitting parameter in the relative liquid water permeability.

Water Resources Research 10.1002/2013WR015127

ZHANG ET AL. VC 2014. American Geophysical Union. All Rights Reserved. 8103

Page 21: Numerical study of evaporation‐induced salt accumulation ... · the liquid water flow to cause changes in the evaporation rate, but reduce the soil porosity and lead to obstruction

ar a fitting parameter in the salt resistance formulation, s m21.as a fitting parameter in the surface resistance formulation, s m21.ax a fitting parameter in the osmotic coefficient.b1 a fitting parameter in the bulk thermoconductivity.b2 a fitting parameter in the bulk thermoconductivity.b3 a fitting parameter in the bulk thermoconductivity.bk a fitting parameter in the relative liquid water permeability.br a fitting parameter in the salt resistance formulation, s m21.bs a fitting parameter in the surface resistance formulation, s m21.bx a fitting parameter in the osmotic coefficient.ca specific heat capacity of the dry air, 1013 J kg21 K21.cl specific heat capacity of freshwater, 4200 J kg21 K21.cs specific heat capacity of soil solid grains, J kg21 K21.cv specific heat capacity of vapor, 2000 J kg21 K21.f precipitation (positive) and dissolution rate of salt per unit volume, kg m23 s21.fc mass fraction of clay in the soil.g gravitational acceleration, 9.8 m s22.hr relative humidity in the soil.hrRef relative humidity at the reference point.ks intrinsic permeability, m2.kr relative permeability.m a parameter representing the density of a virtual medium that stores the precipitated salt, m3

kg21.n1 fitting parameter in the water retention curve.n2 fitting parameter in the water retention curve.pa atmospheric pressure, 100,325 Pa.p�v saturated water vapor pressure, Pa.px a fitting parameter in the osmotic coefficient.ql liquid water flux, m s21.qv water vapor flux, m s21.ra aerodynamic resistance to water vapor transfer, s m21.rH aerodynamic resistance to the heat transfer, s m21.rHSide aerodynamic resistance on the side of the soil column, s m21.rs surface resistance, s m21.rsc resistance due to salt precipitation as efflorescence, s m21.t time, s.w a fitting parameter in the water retention curve representing the mass fraction of one soil.z vertical space coordinate positive upward, m.Dt time step, s.Dz nodal spacing below the NSL, m.DzNSL the depth of the NSL, m.

ReferencesBaveye, P., and A. Valocchi (1989), An evaluation of mathematical models of the transport of biologically reacting solutes in saturated soils

and aquifers, Water Resour. Res., 25(6), 1413–1421.Bear, J. (1972), Dynamics of Fluids in Porous Media, Dover, Mineola, N. Y.Bear, J., and A. H.-D. Cheng (2010), Modeling Groundwater Flow and Contaminant Transport, Springer, N. Y.Bresler, E. (1981), Transport of salts in soils and subsoils, Agric. Water Manage., 4(1981), 35–62.Camillo, P. J., and R. J. Gurney (1986), A resistance parameter for bare-soil evaporation models, Soil Sci., 141(2), 95–105, doi:10.1097/

00010694-198602000-00001.Campbell, G. S. (1974), A simple method for determining unsaturated conductivity from moisture retention data, Soil Sci., 117, 311–314.Campbell, G. S. (1977), An Introduction to Environmental Biophysics, Springer, N. Y.Campbell, G. S. (1985), Soil Physics With BASIC: Transport Models for Soil-Plant Systems, Elsevier, N. Y.Carsel, R., and R. Parrish (1988), Developing joint probability distributions of soil water retention characteristics, Water Resour. Res., 24(5),

755–769.Carter, E. S., S. M. White, and A. M. Wilson (2008), Variation in groundwater salinity in a tidal salt marsh basin, North Inlet Estuary, South Car-

olina, Estuarine Coastal Shelf Sci., 76(3), 543–552, doi:10.1016/j.ecss.2007.07.049.

AcknowledgmentsThis work was supported by theNational Basic Research Program ofChina (973 Program, 2012CB417005)and National Centre for GroundwaterResearch and Training (NCGRT). Theauthors express their gratitude to theeditors and the anonymous reviewersfor their valuable suggestions. Thedata related to this research areavailable to all interested researchersupon request.

Water Resources Research 10.1002/2013WR015127

ZHANG ET AL. VC 2014. American Geophysical Union. All Rights Reserved. 8104

Page 22: Numerical study of evaporation‐induced salt accumulation ... · the liquid water flow to cause changes in the evaporation rate, but reduce the soil porosity and lead to obstruction

Cass, A., G. Campbell, and T. Jones (1984), Enhancement of thermal water vapor diffusion in soil, Soil Sci. Soc. Am. J., 48(1), 25–32.Chung, S., and R. Horton (1987), Soil heat and water flow with a partial surface mulch, Water Resour. Res., 23(12), 2175–2186.Clement, T., B. Hooker, and R. Skeen (1996), Macroscopic models for predicting changes in saturated porous media properties caused by

microbial growth, Ground Water, 34(5), 934–942.Fayer, M. J., and C. S. Simmons (1995), Modified soil water retention functions for all matric suctions, Water Resour. Res., 31(5), 1233–1238.Fujimaki, H., and M. Inoue (2003), A flux-controlled steady-state evaporation method for determining unsaturated hydraulic conductivity

at low matric pressure head values, Soil Sci., 168(6), 385–395, doi:10.1097/01.ss.0000075284.87447.cf.Fujimaki, H., T. Shimano, M. Inoue, and K. Nakane (2006), Effect of a salt crust on evaporation from a bare saline soil, Vadose Zone J.,

5(1994), 1246–1256, doi:10.2136/vzj2005.0144.Gran, M., J. Carrera, J. Massana, M. W. Saaltink, S. Olivella, C. Ayora, and A. Lloret (2011a), Dynamics of water vapor flux and water separation

processes during evaporation from a salty dry soil, J. Hydrol., 396(3–4), 215–220, doi:10.1016/j.jhydrol.2010.11.011.Gran, M., J. Carrera, S. Olivella, M. W. Saaltink, and U. P. De Catalunya (2011b), Modeling evaporation processes in a saline soil from satura-

tion to oven dry conditions, Hydrol. Earth Syst. Sci., 15, 2077–2089, doi:10.5194/hess-15-2077-2011.Hughes, B. J. D., and W. E. Sanford (2004), SUTRA-MS: A version of SUTRA modified to simulate heat and multiple-solute transport, U.S.

Geol. Surv. Open File Rep., 2004-1207, 141 pp.Hughes, C. E., J. D. Kalma, P. Binning, G. R. Willgoose, and M. Vertzonis (2001), Estimating evapotranspiration for a temperate salt marsh,

Newcastle, Australia, Hydrol. Processes, 15(6), 957–975, doi:10.1002/hyp.189.Idso, S., R. Reginato, R. Jackon, B. Kimball, and F. Nakayama (1974), The three stages of drying of a field soil, Soil Sci. Soc. Am. J., 38(5),

831–837, doi:10.2136/sssaj1974.03615995003800050037x.Jury, W. A., and R. Horton (2003), Soil Physics, 6th ed., John Wiley, N. Y.Kestin, J., H. Khalifa, and R. Correia (1981), Tables of the dynamic and kinematic viscosity of aqueous NaCl solutions in the temperature

range 20–150 C and the pressure range 0.1–35 MPa, J. Phys. Chem. Ref. Data, 10(1), 71–87.Kimball, B. A., R. D. Jackson, R. J. Reginato, F. S. Nakayama, and S. B. Idso (1976), Comparison of field-measured and calculated soil-heat

fluxes, Soil Sci. Soc. Am. J., 40(1), 18–25.Kollet, S. J., I. Cvijanovic, D. Sch€uttemeyer, R. M. Maxwell, A. F. Moene, and P. Bayer (2009), The influence of rain sensible heat and subsur-

face energy transport on the energy balance at the land surface, Vadose Zone J., 8(4), 846, doi:10.2136/vzj2009.0005.Kondo, J., and N. Saigusa (1990), A parameterization of evaporation from bare soil surfaces, J. Appl. Meteorol., 29(4), 386–390.Lehmann, P., and D. Or (2009), Evaporation and capillary coupling across vertical textural contrasts in porous media, Phys. Rev. E, 80(4),

046318, doi:10.1103/PhysRevE.80.046318.Mehta, B. K., S. Shiozawa, and M. Nakano (1994), Hydraulic properties of a sandy soil at low water contents, Soil Sci., 157(4), 208–214.Millington, R., and J. Quirk (1961), Permeability of porous solids, Trans. Faraday Soc., 57, 1200–1207.Milly, P. C. D. (1984), A simulation analysis of thermal effects on evaporation, Water Resour. Res., 20(8), 1087–1098.Moffett, K. B., D. A. Robinson, and S. M. Gorelick (2010), Relationship of salt marsh vegetation zonation to spatial patterns in soil moisture,

salinity, and topography, Ecosystems, 13, 1287–1302, doi:10.1007/s10021-010-9385-7.Mualem, Y. (1976), A new model for predicting the hydraulic conductivity of unsaturated pore meida, Water Resour. Res., 12(3), 513–522.Murray, F. (1967), On the computation of saturation vapor pressure, J. Appl. Meteorol., 6(1), 203–204.Nachshon, U., N. Weisbrod, M. I. Dragila, and A. Grader (2011a), Combined evaporation and salt precipitation in homogeneous and hetero-

geneous porous media, Water Resour. Res., 47, W03513, doi:10.1029/2010WR009677.Nachshon, U., E. Shahraeeni, D. Or, M. Dragila, and N. Weisbrod (2011b), Infrared thermography of evaporative fluxes and dynamics of salt

deposition on heterogeneous porous surfaces, Water Resour. Res., 47, W12519, doi:10.1029/2011WR010776.Nassar, I., R. Horton, and A. Globus (1992), Simultaneous transfer of heat, water, and solute in porous media. II: Experiment and analysis,

Soil Sci. Soc. Am. J., 56(5), 1357–1365.Norouzi Rad, M., and N. Shokri (2012), Nonlinear effects of salt concentrations on evaporation from porous media, Geophys. Res. Lett., 39,

L04403, doi:10.1029/2011GL050763.Novak, M. D. (2010), Dynamics of the near-surface evaporation zone and corresponding effects on the surface energy balance of a drying

bare soil, Agric. For. Meteorol., 150(10), 1358–1365, doi:10.1016/j.agrformet.2010.06.005.Panda, M., and L. Lake (1996), A physical model of cementation and its effects on single-phase permeability, AAPG Bull., 79(3), 431–443.Penman, H. L. (1948), Natural evaporation from open water, bare soil and grass, Proc. R. Soc. London, Ser. A, 193(1032), 120–145.Philip, J., and D. De Vries (1957), Moisture movement in porous materials under temperature gradients, Eos Trans. AGU, 38(2), 222–232.Ren, J. L., Q. F. Li, M. X. Yu, and H. Y. Li (2012), Variation trends of meteorological variables and their impacts on potential evaporation in

Hailar region, Water Sci. Eng., 5(2), 137–144.Ritchie, J. (1972), Model for predicting evaporation from a row crop with incomplete cover, Water Resour. Res., 8(5), 1204–1213.Rose, D. A. A., F. B. Konukcu, and J. W. A. Gowing (2005), Effect of watertable depth on evaporation and salt accumulation from saline

groundwater, Soil Res., 43(5), 565–573.Saito, H., J. �Simunek, and B. Mohanty (2006), Numerical analysis of coupled water, vapor, and heat transport in the vadose zone, Vadose

Zone J., 5(2), 784–800, doi:10.2136/vzj2006.0007.Sakai, M., N. Toride, and J. �Simunek (2009), Water and vapor movement with condensation and evaporation in a sandy column, Soil Sci.

Soc. Am. J., 73(3), 707–717, doi:10.2136/sssaj2008.0094.Sghaier, N., and M. Prat (2009), Effect of efflorescence formation on drying kinetics of porous media, Transp. Porous Media, 80(3), 441–454,

doi:10.1007/s11242-009-9373-6.Shahraeeni, E., P. Lehmann, and D. Or (2012), Coupling of evaporative fluxes from drying porous surfaces with air boundary layer: Charac-

teristics of evaporation from discrete pores, Water Resour. Res., 48, W09525, doi:10.1029/2012WR011857.Shimojima, E., A. Curtis, and J. Turner (1990), The mechanism of evaporation from sand columns with restricted and unrestricted water

tables using deuterium under turbulent airflow conditions, J. Hydrol., 117, 15–54, doi:10.1016/0022-1694(90)90085-c.�Simunek, J., N. J. Jarvis, M. T. van Genuchten, and A. G€arden€as (2003), Review and comparison of models for describing non-equilibrium

and preferential flow and transport in the vadose zone, J. Hydrol., 272, 14–35.Smits, K., A. Cihan, T. Sakaki, and T. H. Illangasekare (2011), Evaporation from soils under thermal boundary conditions: Experimental and

modeling investigation to compare equilibrium- and nonequilibrium-based approaches, Water Resour. Res., 47, W05540, doi:10.1029/2010WR009533.

Steefel, C. I., and A. C. Lasaga (1994), Transport of multiple chemical species and kinetic precipitation/dissolution reactions with applicationto reactive flow in single phase hydrothermal systems, Am. J. Sci., 294(5), 529–592.

Water Resources Research 10.1002/2013WR015127

ZHANG ET AL. VC 2014. American Geophysical Union. All Rights Reserved. 8105

Page 23: Numerical study of evaporation‐induced salt accumulation ... · the liquid water flow to cause changes in the evaporation rate, but reduce the soil porosity and lead to obstruction

Sun, S. F. (1982), Moisture and heat transport in a soil layer forced by atmospheric conditions, M.Sc. Thesis, Dep of Civ. Eng., Univ. of Conn.,Storrs.

van de Griend, A., and M. Owe (1994), Bare soil surface resistance to evaporation by vapor diffusion under semiarid conditions, WaterResour. Res., 30(2), 181–188.

van Genuchten, M. (1980), A closed-form equation for predicting the hydraulic conductivity of unsaturated soils, Soil Sci. Soc. Am. J., 8,892–898.

Voss, C., and A. Provost (2010), UTRA: A model for saturated-unsaturated, variable-density ground-water flow with solute or energy trans-port, U.S. Geol. Surv. Water Resour. Invest. Rep. 02-4231, 291 pp.

Wilson, G. W., D. G. Fredlund, and S. L. Barbour (1994), Coupled soil-atmosphere modelling for soil evaporation, Can. Geotech. J., 31(2),151–161.

Wissmeier, L., and D. A. Barry (2009), Effect of mineral reactions on the hydraulic properties of unsaturated soils: Model development andapplication, Adv. Water Resour., 32(8), 1241–1254, doi:10.1016/j.advwatres.2009.05.004.

Xue, Y. (1996), The impact of desertification in the Mongolian and the inner Mongolian Grassland on the regional climate, J. Clim., 9(9),2173–2189.

Zheng, C., and G. D. Bennett (2002), Applied Contaminant Transport Modeling, 2nd ed., John Wiley, N. Y.

Water Resources Research 10.1002/2013WR015127

ZHANG ET AL. VC 2014. American Geophysical Union. All Rights Reserved. 8106