• Aucun résultat trouvé

Model limitations 521

Dans le document en fr (Page 27-51)

Our approach has some limitations that need to be understood in order to perform a 522

correct interpretation of the results. One important aspect that the SiSPAT model does not 523

represent is fluid flow due to solute concentration effects (Barbour and Fredlund, 1989;

524

Nachshon et al., 2011). Indeed, as shown by Nassar et al. (1989), the osmotic gradient can be an 525

important driving force for water movement in unsaturated soils with high clay content. We used 526

the approach proposed by Barbour and Fredlund (1989) to determine the order of magnitude of 527

the liquid and water vapor fluxes due to solute concentration gradients. In this analysis, we 528

assumed that sodium chloride (NaCl) was the main component that contributes to the osmotic 529

27 pressure and that the maximum solute concentration occurred at saturated conditions (~25 wt%

530

for NaCl). Also, the observed experimental conditions were utilized to evaluate the temperature 531

in the Van’t Hoff equation (Barbour and Fredlund, 1989), and the experimental soil water content 532

was used to determine the osmotic flow of water within the soil. This analysis resulted in osmotic 533

fluid flows on the order of 10-11 m/s, which are at least one order of magnitude smaller than those 534

obtained using SiSPAT. Thus, in our experiments we expect a small contribution of the osmotic 535

flow of water in the liquid and the gaseous phase.

536

It is also important to discuss that discrepancy between experimental data and model 537

predictions has led to the development of vapor-flux enhancement factors, such as  (Philip and 538

de Vries, 1957; Cass et al., 1984). The inclusion of these enhancement factors in numerical 539

models, such as that used in this work, is due to a lack of sufficiently accurate description of 540

vapor dynamics to represent correctly the soil physics processes that occur during evaporation 541

(Assouline et al., 2013). There are many processes that explain this vapor-flux enhancement 542

(Philip and de Vries, 1957; Cass et al., 1984; Bachmann et al., 2001; Grifoll et al., 2005; Shokri 543

et al., 2009; Shahraeeni and Or, 2012; Assouline et al., 2013; Trautz et al., 2015). For instance, 544

investigations have shown that with proper account of capillary flow, continuity and pathways, 545

no vapor-flux enhancement factors are required (Grifoll et al., 2005; Shokri et al., 2009).

546

Shahraeeni and Or (2012) demonstrated that water transport can be enhanced by ~10% when 547

isolated liquid-phase bridges are present due to a reduction in the gaseous diffusion path length, 548

and that thermal gradients can enhance water vapor diffusion. Also, vapor flux from within the 549

soil profile can be enhanced by thermally driven convective transport mechanisms (Bachmann et 550

28 al., 2001). Recently, Trautz et al. (2015) argued that non-equilibrium phase change is also 551

responsible for vapor-flux enhancement.

552

From the vapor-flux enhancement mechanisms that have been reported in the scientific 553

literature, we hypothesize that a combination between non-equilibrium phase change and cyclic 554

thermal conditions typically found between day and night, could be responsible for the vapor-flux 555

enhancement observed in our experiments. When cyclic thermal conditions are present, two 556

processes can drive soil evaporation: evaporation from the soil surface into the atmosphere during 557

early morning and subsurface evaporation limited by Fickian Diffusion until late afternoon. As 558

explained by Assouline et al. (2013), evaporation from the soil surface depletes the water 559

condensed and redistributed during nighttime. After the depletion of this water, Fickian diffusion 560

becomes the governing process and increases the thickness of the dry layer of soil observed at the 561

soil surface. In addition, as discussed by Trautz et al. (2015), non-equilibrium processes are 562

relevant when cyclic thermal conditions occur.

563

Many studies have used laboratory columns, but very few of them have focused on 564

understanding natural highly saline dry soils. Moreover, typical column experiments do not deal 565

with natural highly salty soils and low water content. Our results show that evaporation in 566

initially dry soils produce moisture content and conductivity distributions that are atypical. This 567

behavior has not been reported in the literature, which typically reports experiments in which an 568

initially saturated soil column is subsequently dried out.

569

5. Conclusions

570

29 To improve the understanding of bare soil evaporation, which is typically the main source 571

of aquifer depletion in zones such as those located in the Altiplano basins of northern Chile, we 572

conducted a modeling study to investigate evaporation processes under non-isothermal 573

conditions. A model that couples liquid water, water vapor, and heat transport was developed and 574

calibrated using laboratory observations performed in a homogeneous natural soil column that 575

was filled with soil from the Huasco salt flat, Chile. Modeled and experimental results only 576

agreed when the soil hydrodynamic properties (water retention and hydraulic conductivity curve) 577

were calibrated. The change in the soil hydrodynamic properties can be explained by 578

precipitation/dissolution reactions that are driven by evaporation and that yield a stratified soil 579

profile when quasi steady state conditions were achieved.

580

Model results showed a good agreement with the experimental observations of the soil 581

water content and thermal profiles, and also reproduced the experimental cumulative evaporation 582

with differences of 0.01 and 0.67 mm for the 0.75 and 0.40 m water table depths, respectively.

583

Model results permitted to distinguish three characteristic zones in the soil profile. The first zone 584

(region 1) is located near the surface, where the total flux is directed upwards and the water 585

movement is mainly due to vapor fluxes driven by pressure gradients. The second zone (region 2) 586

is located below region 1 and is where the liquid water flux dominates. The third zone (region 3) 587

is where the liquid water flux is very small. Model results indicate that evaporation occurs in the 588

upper soil profile and that the position of the evaporation front depends on the water table depth.

589

A sensitivity analysis allowed understanding the impact of the enhancement factor and the 590

tortuosity on the cumulative evaporation. The enhancement factor had the largest influence on 591

30 cumulative evaporation that can even result in condensation at the soil surface. This analysis also 592

allowed discarding different set of calibrated parameters that yield non-physical conditions, 593

which minimize the problem of equifinality. The results presented in this study are important as 594

they allow understanding the main evaporation processes that occur in bare soils from Altiplano 595

basins where these processes are not well understood.

596

Acknowledgments

597

The authors acknowledge funding from the Chilean National Commission for Scientific 598

and Technological Research (CONICYT/FONDECYT/1130522). F. Suárez and J. Gironás thank 599

the Centro de Desarrollo Urbano Sustentable for additional support (CEDEUS - 600

CONICYT/FONDAP/15110020), and J. Gironás also acknowledges the Centro de Investigación 601

para la Gestión Integrada de Desastres Naturales (CIGIDEN - CONICYT/FONDAP/15110017).

602

The modeling study presented in this paper was possible thanks to a grant for a stay of M.F.

603

Hernández-López at Irstea, funded by Pontificia Universidad Católica de Chile and Irstea. We 604

also thank the anonymous reviewers for the positive comments that improved this paper.

605

References

606

Abu-El-Sha’r WY, Abriola LM. 1997. Experimental assessment of gas transport mechanisms in 607

natural porous media: parameter evaluation. Water Resource Research, 33(4): 505-516.

608

Alazard M, Leduc C, Travi Y, Boulet G, Ben Salem A. 2015. Estimating evaporation in semi-arid 609

areas facing data scarcity: example of the El Haouareb dam (Merguellil catchment, Central 610

Tunisia). Journal of Hydrology: Regional Studies 3:265-284.

611

31 Assouline S, Tyler SW, Selker JS, Lunati I, Higgins CW, Parlange MB. 2013. Evaporation from 612

a shallow water table: diurnal dynamics of water and heat at the surface of drying sand.

613

Water Resources Research, 49:4022-4034.

614

Bachmann J, Horton R, van del Ploeg RR. 2001. Isothermal and nonisothermal evaporation from 615

four Sandy soils of different water repelency. Soil Sci. Soc. Am. J., 65, 1599-1607.

616

Barbour SL, Fredlund DG. 1989. Mechanisms of osmotic flow and volume change in clay soils.

617

Can. Geotech. J., 26, 551-562.

618

Benavente D, García del Cura MA, Fort R, Ordónez S. 1999. Thermodynamic modelling of 619

changes induced by salt pressure crystallisation in porous media of stone. Journal of Crystal 620

Growth, 204: 168-178.

621

Boulet G, Braud I, Vauclin M. 1997. Study of the mechanisms of evaporation under arid 622

conditions using a detailed model of the soil-atmosphere continuum. Application to the 623

EFEDA I experiment. Journal of Hydrology, 193(1-4): 114-141.

624

Boulet G, Chehbouni A, Braud I, Vauclin M, Haverkamp R, Zammit C. 2000. A simple water 625

and energy balance model designed for regionalization and remote sensing data utilization.

626

Agricultural and Forest Meteorology, 105: 117-132.

627

Braud I, Bariac T, Gaudet JP, Vauclin M. 2005a. SiSPAT-Isotope, a coupled heat, water and 628

stable isotope (HDO and H218O) transport model for bare soil. Part I: Model description 629

and first verification, J. Hydrology, 309(1-4), 277-300.

630

Braud I, Bariac T, Vauclin M, Boujamlaoui Z, Gaudet JP, Biron Ph, Richard P. 2005b. SiSPAT-631

Isotope, a coupled heat, water and stable isotope (HDO and H218O) transport model for 632

32 bare soil. Part II: Evaluation and sensitivity tests using two laboratory data sets, J.

633

Hydrology, 309(1-4), 301-320.

634

Braud I, Biron P, Bariac T, Richard P, Canale L, Gaudet JP, Vauclin M. 2009a. Isotopic 635

composition of bare soil evaporated water vapor. Part I: RUBIC IV experimental setup and 636

results. Journal of Hydrology, 369: 1-16.

637

Braud I, Bariac T, Biron P, Vauclin M. 2009b. Isotopic composition of bare soil evaporated water 638

flow. Part II: Modeling of RUBIC IV experimental results. Journal of Hydrology, 369: 17-639

29.

640

Braud I, Dantas-Antonino AC, Vauclin M, Thony JL, Ruelle P. 1995. A simple soil-plant-641

atmosphere transfer model (SiSPAT) development and field verification. Journal of 642

Hydrology, 166: 213-250.

643

Brooks RH, Corey AT, 1964. Hydraulic properties of porous media. Hydrology paper 3, 644

Colorado State University, Fort Collins, 27 pp.

645

Cahill AT, Parlange MB, 1998. On water vapor transport in field soils. Water Resources 646

Research, 34(4): 731-739.

647

Cass A, Campbell GS, Jones TL. 1984. Enhacement of thermal water vapor difussion in soil. Soil 648

Science Society of America Proceedings, 48: 25-32.

649

de la Fuente A, Niño Y. 2010. Temporal and spatial features of the thermohydrodynamics of 650

shallow salty lagoons in northern Chile. Limnology and Oceanography 55:279-288.

651

Fierro V. 2015. SAR effects on evaporation fluxes from shallow groundwater. M. Sc. Thesis, 652

Pontificia Universidad Católica de Chile, Santiago, Chile.

653

33 Gran M, Carrera J, Massana J, Saaltink MW, Olivella S, Ayora C, Lloret A. 2011. Dynamics of 654

water vapor flux and water separation processes during evaporation from a salty dry soil.

655

Journal of Hydrology, 396: 215-220.

656

Grifoll J, Gastó JM, Cohen Y. 2005. Non-isothermal soil water transport and evaporation.

657

Advances in Water Resources, 28: 1254-1266.

658

Hernández-López MF, Gironás J, Braud I, Suárez F, Muñoz JF. 2014. Assessment of evaporation 659

and water fluxes in a column of dry saline soil subject to different water table levels.

660

Hydrological Processes, 28(10): 3655-3669.

661

Johnson E, Yáñez J, Ortiz C, Muñoz J. 2010. Evaporation from shallow groundwater in closed 662

basins in the Chilean Altiplano. Hidrological Sciences Journal, 55(4): 624-635.

663

Kampf S, Tyler S, Ortiz C, Muñoz J, Adkins P. 2005. Evaporation and land surface energy 664

budget at the Salar de Atacama, northern Chile. Journal of Hydrology 310:236–252.

665

Konukcu F, Istanbulluoglu A, Kocaman I. 2004. Determination of water content in drying soils:

666

incorporating transition from liquid phase to vapour phase. Australian Journal of Soil 667

Research 42(1): 1-8.

668

Laurent JP, Guerre-Chaley C. 1995. Influence de la teneur en eau et de la température sur la 669

conductivité thermique du béton cellulaire autoclave. Materials and Structures, 28:464-472.

670

Lictevout E, Maass C, Córdoba D, Herrera V, Payano R. 2013. Recursos Hídricos Región de 671

Tarapacá – Diagnóstico y Sistematización de la Información. CIDERH. ISBN:978 956 302 672

081 – 6. Available at: http://www.ciderh.cl/documentos/recursos-hidricos-region-de-673

tarapaca/ (last accessed July 2016).

674

34 Liu S, Lu L, Mao D, Jia L. 2007. Evaluating parametrizations of aerodynamic resistance to heat 675

transfer using field measurements. Hydrology and Earth System Sciences, 11: 769-783.

676

Millington RJ, Quirk JM. 1961. Permeability of porous solids. 57. Transaction of the Faraday 677

Society, 57: 1200-1207.

678

Milly PCD. 1984. A simulation analysis of thermal effects on evaporation from soil. Water 679

Resources Research, 20(8): 1087-1098.

680

Nachshon U, Weisbroad N, Dragila MI, Grader A. 2011. Combined evaporation and salt 681

precipitation in homogeneous and heteregenous porous media. Water Resource Research, 682

47: W03513, doi:10.1029/2010WR009677.

683

Nachshon U, Weisbrod N. 2015. Beyond the Salt Crust: On Combined Evaporation and 684

Subflorescent Salt Precipitation in Porous Media. Transport in Porous Media, 110: 295-685

310. DOI: 10.1007/s11242-015-0514-9.

686

Nassar IN, Horton R. 1989. Water transport in unsaturated nonisothermal salty soil: II.

687

Theoretical development. Soil Science Society of America Proceedings, 53: 1330-1337.

688

Novak MD. 2010. Dynamics of the near-surface evaporation zone and corresponding effects on 689

the surface energy balance of a drying bare soil. Agricultural and Forest Meteorology, 690

1501358-1365.

691

Penman HL. 1940. Gas and vapor movement in the soil. I. The diffusion of vapors through 692

porous solids. The Journal of Agricultural Science, 30(4): 570-581.

693

Philip JR, de Vries DA. 1957. Moisture movement in porous materials under temperature 694

gradient. Transactions American Geophysical Union, 38: 222-232.

695

35 Saito H, Simunek J, Mohanty BP. 2006. Numerical Analysis of Coupled Water, Vapor, and Heat 696

Transport in the Vadose Zone. Vadose Zone Journal, 5: 784-800.

697

Schulz S, Horovitz M, Rausch R, Michelsen N, Mallast U, Köhne M, Siebert C, Shüth C, Al-698

Saud M, Merz R. 2015. Groundwater evaporation from salt pans: examples from the 699

eastern Arabian Peninsula. Journal of Hydrology. doi:10.1016/j.jhydrol.2015.10.048 700

Scotter DR. 1974. Salt and water movement in relatively dry soil. Australian Journal of Soil 701

Research, 12(1): 27-35.

702

Sghaier N, Geoffroy S, Prat M, Eloukabi H, Nasrallah SB. 2014. Evaporation-driven growth of 703

large crystallized salt structures in a porous medium. Phys. Rev. E 90, 042402 704

Shahraeeni E, Or D. 2012. Pore scale mechanisms for enhanced vapor transport through partially 705

saturated porous media. Water Resour. Res., 48, W05511, doi:10.1029/2011WR011036.

706

Shokri NP, Lehmann P, Or D. 2009. Critical evaluation of enhancement factors for vapor 707

transport through unsaturated porous media. Water Resour. Res., 45, W10433, 708

doi:10.1029/2009WR007769.

709

Shuttleworth WJ, Wallace JS. 1985. Evaporation from sparse crops an energy combination 710

theory. Quarterly Journal of the Royal Meteorological Society, 111: 839-855.

711

Tang J, Zhuang Q. 2008. Equifinality in parameterization of process-based biogeochemistry 712

models: A significant uncertainty source to the estimation of regional carbon dynamics, 713

Journal of Geophysical Research, 113, G04010, doi:10.1029/2008JG000757.

714

Tóth B, Makó A, Guadagnini A, Tóth G. 2012. Water retention of salt-affected soils: quantitative 715

estimation using soil survey information. Arid Land Research and Management 26:103-716

121.

717

36 Trautz AC, Smits KM, Cihan A. 2015. Continuum-scale investigation of evaporation from bare 718

soil under different boundary and initial conditions: An evaluation of nonequilibrium phase 719

change, Water Resour. Res., 51, 7630–7648, doi:10.1002/2014WR016504.

720

van Genuchten MT. 1980. A closed-form equation for predicting the hydraulic conductivity on 721

unsatured soils. Soil Science Society of America Proceedings, 44(5): 892-898.

722

Vásquez C, Ortiz C, Suárez F, Muñoz JF. 2013. Modeling flow and reactive transport to explain 723

mineral zoning in the Atacama salt flat aquifer, Chile. Journal of Hydrology, 490:114-125.

724

Weisbrod N, Nachshon U, Dragila M, Grader A. 2014. Micro-CT analysis to explore salt 725

precipitation impact on porous media permeability. Transport and Reactivity of Solutions in 726

Confined Hydrosystems. Part of the series NATO Science for Peace and Security Series C:

727

Environmental Security, 231-241.

728

Wissmeier L, Barry DA. 2008. Reactive transport in unsaturated soil: Comprehensive modeling 729

of the dynamic spatial and temporal mass balance of water and chemical components, 730

Advances in Water Resources, 31, 858–875.

731 732

37 Table captions

733

Table 1: Coefficients and variables used in the SiSPAT model to estimate liquid water and water 734

vapor flows.

735

Table 2: Model parameters, and initial and boundary conditions.

736

Table 3: Model parameters used for the three horizons.

737

Table 4: Model calibration efficiency for soil water content () and temperature (T), and 738

comparison of the experimental cumulative evaporation and the data obtained from the calibrated 739

model.

740 741

38 Figure captions

742

Figure 1: Observed and modeled soil water content () and temperature profiles, and observed 743

electrical conductivity () profile for different water table levels (WTL), after 20 days of fixing 744

the water table level. Model results assume a vertically homogeneous soil profile. (a) and (b) 745

show the soil water content profiles for WTL of 0.75 and 0.40 m, respectively; (c) and (d) show 746

the temperature profiles for WTL of 0.75 and 0.40 m, respectively; and (e) and (f) show the 747

electrical conductivity profiles for WTL of 0.75 and 0.40 m, respectively.

748

Figure 2: Observed and modeled soil water content () and temperature profiles for different 749

water table levels (WTL), after 20 days of fixing the water table level and assuming the soil 750

stratifies in three horizons (H1, H2, and H3) due to salt transport. (a) and (b) show the soil water 751

content profiles for WTL of 0.75 and 0.40 m, respectively; (c) and (d) show the temperature 752

profiles for WTL of 0.75 and 0.40 m, respectively.

753

Figure 3: Variation of liquid and water vapor flux along the soil profile for different water table 754

levels (WTL). (a) and (b) show the liquid flux (qL), the water vapor flux (qv) and the total water 755

flux (qtotal) for WTL of 0.75 and 0.40 m, respectively; (c) and (d) show the thermal vapor flux 756

(qvT), the isothermal (due to pressure) vapor flux (qvh), and the total vapor flux (qv) for WTL of 757

0.75 and 0.40 m, respectively. Negative (positive) fluxes correspond to upward (downward) 758

movement.

759 760

39 Figure 4: Comparison of cumulative evaporation (mm) for different tortuosity values (a) and 761

enhancement factors () for the 0.75 and 0.40-m water table levels. a* was calculated based on 762

equation (15) (Millington and Quirk, 1961), and equation (16) was used when  = variable (Cass 763

et al., 1984).

764

Figure 5: Water retention curves fitted for each horizon (H1, H2, and H3) for the 0.75 m water 765

table level (a), and the 0.40 m water table level (b).

766

Figure 6. Comparison of the soil water content and temperature profiles for the 0.75-m (a and c) 767

and 0.40-m (b and d) water table levels (WTL), for the simulation presented in Figure 2 (black) 768

and after steady state was reached (red) when using the bottom boundary conditions equal to the 769

top evaporation flux.

770

Figure 7: Comparison of the water vapor fluxes profiles for the 0.75 and 0.40 m water table levels 771

(WTL). Results are shown for different tortuosity values, a, (a) and (c), and for different 772

enhancement factor values,  (b) and (d). a* was calculated based on equation (15) (Millington 773

and Quirk, 1961), and equation (16) was used when  = variable (Cass et al., 1984). Negative 774

(positive) fluxes correspond to upward (downward) movement.

775 776

40

Table 1: Coefficients and variables used in the SiSPAT model to estimate liquid water and water vapor flows.

777 778

Parameter Formula

Saturated vapor pressure

Diffusivity of water vapor in the air (Philip and De Vries, 1957)

88

Isothermal diffusivity of water vapor (Philip and De Vries, 1957)

h

Thermal diffusivity of water vapor (Philip and De Vries, 1957)

T

Apparent thermal conductivity (Laurent and Guerre-Chaley, 1995) 

are empirical fitting parameters of the water retention curve; Ks: saturated hydraulic conductivity (m s-1);

781

41

Table 2: Model parameters and initial and boundary conditions.

787

Parameter Value Observations

Saturated moisture content s (m3 m-3) 0.3 Measured

Residual moisture content r (m3 m-3) 0.0 Parameter fitted from the water retention data Inverse of the air-entry pressure (m-1) 0.7160 Parameter fitted from the water retention data

n (-) 1.1859 Parameter fitted from the water retention data

Saturated hydraulic conductivity Ks (m s-1) 1.06 x 10-5 Measured

Porosity (-) 0.3 Assumed equal to the saturated moisture content

aLG 0.3

Parameters of the apparent thermal conductivity from the Laurent and Guerre-Chaley model (1995).

bLG 0.5

cLG 1.0

dLG 4.0

0 0.12

Initial soil temperature conditions T(z)

Linear interpolation of temperature measurements in the soil at 0.05, 0.10, 0.15, 0.20, 0.25, 0.30, 0.35, 0.40 and 0.45 m. The initial time corresponds to the 1st day after fixing the water table depth.

Initial soil pressure conditions h(z)

Determined from moisture content measurements at 0.05, 0.10, 0.15, 0.20, 0.25, 0.30, 0.35, 0.40 and 0.45 m depth using the water retention curve. The initial time corresponds to the 1st day after fixing the water table depth.

Boundary conditions at the bottom

Known flux A known water flux equal to the evaporation flux This boundary condition was then evaluated in a sensitivity analysis.

T(t) Extrapolated from the measurements of the thermal profile within the soil column.

Boundary condition on the surface T(t) Temperature measured at the soil column surface.

788 789

42

Table 3: Model parameters used for the three horizons.

790

Depth of the water table (m) 0.75 0.40

Horizon H1 H2 H3 H1 H2 H3

Horizon depth (m) 0.20 0.10 0.90 0.15 0.15 0.90

Saturated moisture content, s (m3 m-3) 0.300 0.300 0.300 0.300 0.300 0.300 Residual moisture content, r (m3 m-3) 0.033 0.000 0.000 0.033 0.000 0.000 Inverse of the air entry pressure, (cm-1) 0.6677 0.7163 0.6667 0.40 0.555 0.6667 Shape parameter of the water retention curve, n (-) 1.2370 1.16 1.2470 1.2370 1.1859 1.2470 Saturated hydraulic conductivity, Ks (m s-1) 1.5x10-5 5.0x10-5 1.0x10-5 1.0x10-8 1.5x10-8 1.0x10-5 Shape factor of the hydraulic conductivity curve,

(-) 18 18 18 25 30 30

aLG 0.300 0.734

bLG 0.9 0.3

cLG 35.0 35.0

dLG 5.00 3.82

0 0.12 0.50

791 792 793 794 795 796

43

Table 4: Model calibration efficiency for soil water content () and temperature (T), and comparison of the

797

experimental cumulative evaporation and the data obtained from the calibrated model.

798 799

Depth to groundwater table (m)

0.75 0.40

RMSE (m3 m-3) 0.0095 0.020

B (m3 m-3) 0.002 -0.0003

E (-) 0.985 0.50

RMSET (°C) 2.08 2.61

BT (°C) 0.40 -2.41

ET (-) 0.67 0.46

Experimental cumulative evaporation (mm) 0.85 5.78

Modeled cumulative evaporation (mm) 0.86 5.11

RMSE: root mean square error; B: bias; E: Nash-Sutcliffe efficiency

800 801 802

44 803

Figure 1: Observed and modeled soil water content () and temperature profiles, and observed electrical conductivity

804

() profile for different water table levels (WTL), after 20 days of fixing the water table level. Model results assume

805

a vertically homogeneous soil profile. (a) and (b) show the soil water content profiles for WTL of 0.75 and 0.40 m,

806

respectively; (c) and (d) show the temperature profiles for WTL of 0.75 and 0.40 m, respectively; and (e) and (f)

807

show the electrical conductivity profiles for WTL of 0.75 and 0.40 m, respectively.

808

45 809

Figure 2: Observed and modeled soil water content () and temperature profiles for different water table levels

810

(WTL), after 20 days of fixing the water table level and assuming the soil stratifies in three horizons (H1, H2, and

811

H3) due to salt transport. (a) and (b) show the soil water content profiles for WTL of 0.75 and 0.40 m, respectively;

812

(c) and (d) show the temperature profiles for WTL of 0.75 and 0.40 m, respectively.

813

46 814

Figure 3: Variation of liquid and water vapor flux along the soil profile for different water table levels (WTL). (a)

815

and (b) show the liquid flux (qL), the water vapor flux (qv) and the total water flux (qtotal) for WTL of 0.75 and 0.40

816

m, respectively; (c) and (d) show the thermal vapor flux (qvT), the isothermal (due to pressure) vapor flux (qvh), and

817

the total vapor flux (qv) for WTL of 0.75 and 0.40 m, respectively. Negative (positive) fluxes correspond to upward

818

(downward) movement.

819 820 821

47 822

Figure 4: Comparison of cumulative evaporation (mm) for different tortuosity values (a) and enhancement factors

823

() for the 0.75 and 0.40-m water table levels. a* was calculated based on equation (15) (Millington and Quirk,

824

1961), and equation (16) was used when = variable (Cass et al., 1984).

825 826 827

48 828

Figure 5: Water retention curves fitted for each horizon (H1, H2, and H3) for the 0.75 m water table level (a), and the

829

0.40 m water table level (b).

830 831 832

49 833

834

Figure 6. Comparison of the soil water content and temperature profiles for the 0.75-m (a and c) and 0.40-m (b and

835

d) water table levels (WTL), for the simulation presented in Figure 2 (black) and after steady state was reached (red)

836

when using the bottom boundary conditions equal to the top evaporation flux.

837 838

50 839

Figure 7: Comparison of the water vapor fluxes profiles for the 0.75 and 0.40 m water table levels (WTL). Results

840

are shown for different tortuosity values, a, (a) and (c), and for different enhancement factor values, (b) and (d).

are shown for different tortuosity values, a, (a) and (c), and for different enhancement factor values, (b) and (d).

Dans le document en fr (Page 27-51)

Documents relatifs