the kinetics and energetics of transitions between water repellent and wettable soil conditions: a...

7
HYDROLOGICAL PROCESSES Hydrol. Process. 21, 2248–2254 (2007) Published online in Wiley InterScience (www.interscience.wiley.com) DOI: 10.1002/hyp.6747 The kinetics and energetics of transitions between water repellent and wettable soil conditions: a linear free energy analysis of the relationship between WDPT and MED/CST Peter Douglas, 1 * Kathryn A. Mainwaring, 1 Christopher P. Morley 1,2 and Stefan H. Doerr 3 1 Chemistry Department, University of Wales Swansea, Singleton Park, Swansea, SA2 8PP, UK 2 School of Chemistry, Cardiff University, PO Box 912, Cardiff, CF10 3AT, UK 3 Department of Geography, University of Wales Swansea, Singleton Park, Swansea, SA2 8PP, UK Abstract: Soil water repellency is commonly assessed using the Water Droplet Penetration Time (WDPT) and Molarity of Ethanol (MED) or Critical Surface Tension (CST) tests. The former is a kinetic measurement indicating the persistence of water repellency; the latter, a thermodynamic measurement, gives the initial severity of water repellency. This study aims to provide a theoretical framework to understand (i) what determines differences in persistence and initial severity of soil water repellency and (ii) correlations between such data. A linear free energy relationship between Solution Droplet Penetration Time (SDPT) and the difference between the surface tension of the droplet solution, sol , and the Critical Surface Tension of the soil, c x , has been derived. ln1/SDPT D lnA [f sol c x /Nk T] Here: sol is the surface tension of water (for WDPT) or the solution being used (SDPT); N is the number of adsorbed molecules per unit area; k the Boltzmann constant; T the temperature; A is 1/SDPT when c x D sol ; and f is an experimentally determined parameter. WDPT, SDPT (ethanol/water and propan-1-ol/water), and CST measurements for a group of sandy soils support this analysis. The value of f obtained (4Ð2(š0Ð4)), suggests that the surface free energy contribution to the free energy of activation of wetting is given by the difference between the cohesive energy of the molecular film adsorbed on the soil grains and that of the wetting solution. In this interpretation: c x is determined by the cohesive energy of the organic film adsorbed on the soil; WDPT is determined by the difference in cohesive energies between this adsorbed film and the droplet solution; and soil-to-soil variations in both WDPT and c x are due to organic films of different cohesive energies present on the soil particles. For aqueous solutions of simple linear alcohols and ketones, SDPT depends on sol but is independent of the compound used to control sol . Copyright 2007 John Wiley & Sons, Ltd. KEY WORDS water repellence; hydrophobic; MED; WDPT; CST Received 20 January 2006; Accepted 18 January 2007 INTRODUCTION Water Droplet Penetration Time (WDPT; Watson and Letey, 1970) and Molarity of Ethanol (MED; King, 1981) tests are widely used in studies of soil water repellency. WDPT is a kinetic measurement and gives the persistence of water repellency (Dekker et al., 1999). MED or, more fundamentally, the Critical Surface Tension, CST, is a thermodynamic measurement and gives the initial severity of water repellency (Dekker and Ritsema, 1994; Roy and McGill, 2002). There have been a number of empirical studies of correlations between the data obtained from these two tests. Crockford et al. (1991) found a close relationship between the WDPT and MED test results for most, but not all, samples in a study of Australian soils. Harper and Gilkes (1994) found good correlation for WDPTs ranging from 10 to 2000 s and * Correspondence to: Peter Douglas, Chemistry Department, University of Wales Swansea, Singleton Park, Swansea, SA2 8PP, UK. E-mail: [email protected] computed a conversion formula MED D1Ð3 C 1Ð1 (log WDPT) in a study of Australian clayey soils. In a study of medium textured Portuguese soils with WDPTs ranging from <5 s to >5 h Doerr (1998) reported that MED corresponded reasonably well to WDPT for highly water repellent, but less well for moderately water repellent, soils. However, Dekker and Ritsema (1994), with a much larger sample size of Dutch dune sands (WDPTs <5 s to >6 h), found only a limited correlation between the two methods. There is currently a lack of a theoretical framework to understand (i) what determines differences in persis- tence and initial severity of soil water repellency and (ii) correlations between such data where they are found. Here we explore the relationship between the WDPT and CST tests using a combination of kinetic and thermo- dynamic arguments to provide such a framework. The approach used is evaluated using data from a range of sandy soils examined in a series of previous studies (Llewellyn et al., 2004; Mainwaring, et al., 2004, Doerr et al., 2005a; Morley et al., 2005). Copyright 2007 John Wiley & Sons, Ltd.

Upload: peter-douglas

Post on 12-Jun-2016

212 views

Category:

Documents


0 download

TRANSCRIPT

HYDROLOGICAL PROCESSESHydrol. Process. 21, 2248–2254 (2007)Published online in Wiley InterScience(www.interscience.wiley.com) DOI: 10.1002/hyp.6747

The kinetics and energetics of transitions between waterrepellent and wettable soil conditions: a linear free energyanalysis of the relationship between WDPT and MED/CST

Peter Douglas,1* Kathryn A. Mainwaring,1 Christopher P. Morley1,2 and Stefan H. Doerr3

1 Chemistry Department, University of Wales Swansea, Singleton Park, Swansea, SA2 8PP, UK2 School of Chemistry, Cardiff University, PO Box 912, Cardiff, CF10 3AT, UK

3 Department of Geography, University of Wales Swansea, Singleton Park, Swansea, SA2 8PP, UK

Abstract:

Soil water repellency is commonly assessed using the Water Droplet Penetration Time (WDPT) and Molarity of Ethanol(MED) or Critical Surface Tension (CST) tests. The former is a kinetic measurement indicating the persistence of waterrepellency; the latter, a thermodynamic measurement, gives the initial severity of water repellency. This study aims to providea theoretical framework to understand (i) what determines differences in persistence and initial severity of soil water repellencyand (ii) correlations between such data. A linear free energy relationship between Solution Droplet Penetration Time (SDPT)and the difference between the surface tension of the droplet solution, �sol, and the Critical Surface Tension of the soil, �cx,has been derived.

ln�1/SDPT� D lnA � [f��sol � �cx�/NkT]

Here: �sol is the surface tension of water (for WDPT) or the solution being used (SDPT); N is the number of adsorbed moleculesper unit area; k the Boltzmann constant; T the temperature; A is 1/SDPT when �cx D �sol; and f is an experimentally determinedparameter. WDPT, SDPT (ethanol/water and propan-1-ol/water), and CST measurements for a group of sandy soils supportthis analysis. The value of f obtained (4Ð2 (š0Ð4)), suggests that the surface free energy contribution to the free energy ofactivation of wetting is given by the difference between the cohesive energy of the molecular film adsorbed on the soil grainsand that of the wetting solution.

In this interpretation: �cx is determined by the cohesive energy of the organic film adsorbed on the soil; WDPT is determinedby the difference in cohesive energies between this adsorbed film and the droplet solution; and soil-to-soil variations in bothWDPT and �cx are due to organic films of different cohesive energies present on the soil particles. For aqueous solutions ofsimple linear alcohols and ketones, SDPT depends on �sol but is independent of the compound used to control �sol. Copyright 2007 John Wiley & Sons, Ltd.

KEY WORDS water repellence; hydrophobic; MED; WDPT; CST

Received 20 January 2006; Accepted 18 January 2007

INTRODUCTION

Water Droplet Penetration Time (WDPT; Watson andLetey, 1970) and Molarity of Ethanol (MED; King, 1981)tests are widely used in studies of soil water repellency.WDPT is a kinetic measurement and gives the persistenceof water repellency (Dekker et al., 1999). MED or,more fundamentally, the Critical Surface Tension, CST,is a thermodynamic measurement and gives the initialseverity of water repellency (Dekker and Ritsema, 1994;Roy and McGill, 2002). There have been a numberof empirical studies of correlations between the dataobtained from these two tests. Crockford et al. (1991)found a close relationship between the WDPT and MEDtest results for most, but not all, samples in a study ofAustralian soils. Harper and Gilkes (1994) found goodcorrelation for WDPTs ranging from 10 to 2000 s and

* Correspondence to: Peter Douglas, Chemistry Department, Universityof Wales Swansea, Singleton Park, Swansea, SA2 8PP, UK.E-mail: [email protected]

computed a conversion formula MED D �1Ð3 C 1Ð1 (logWDPT) in a study of Australian clayey soils. In a study ofmedium textured Portuguese soils with WDPTs rangingfrom <5 s to >5 h Doerr (1998) reported that MEDcorresponded reasonably well to WDPT for highly waterrepellent, but less well for moderately water repellent,soils. However, Dekker and Ritsema (1994), with a muchlarger sample size of Dutch dune sands (WDPTs <5 s to>6 h), found only a limited correlation between the twomethods.

There is currently a lack of a theoretical frameworkto understand (i) what determines differences in persis-tence and initial severity of soil water repellency and(ii) correlations between such data where they are found.Here we explore the relationship between the WDPT andCST tests using a combination of kinetic and thermo-dynamic arguments to provide such a framework. Theapproach used is evaluated using data from a range ofsandy soils examined in a series of previous studies(Llewellyn et al., 2004; Mainwaring, et al., 2004, Doerret al., 2005a; Morley et al., 2005).

Copyright 2007 John Wiley & Sons, Ltd.

KINETICS AND ENERGETICS OF SOIL WATER REPELLENCY 2249

MATERIALS AND METHODS

Soil sampling and sample preparation

Samples of sandy soil were taken from locationswith differing climates and vegetation cover from theUK, Netherlands (NL), Portugal (PT), Greece (GK), andAustralia (AU). The sample sites in the Netherlandsand UK experience an oceanic humid-temperate climatewith rainfall occurring throughout the year. The sitesin Greece are also temperate, but with a summer dryseason. The sites in Portugal and Australia have a warmerMediterranean-type climate with prolonged dry periodsduring the summer. All samples were surface material(<20 cm depth) except at the Netherlands site where theywere taken from successive 10 cm depths up to 40 cm.Sample locations, particle size data, water repellencycharacteristics and total organic carbon are given inTable I.

Samples were dried at 20 °C and passed through a2 mm sieve prior to analysis. Apart from some largerorganic debris all soil components passed the 2 mm sieve.All soils were of medium sand texture (0Ð22–0Ð70 mmdiameter) with a clay content of <0Ð1%. Sub-sampleswere obtained by coning and quartering. For more detailsregarding sampling sites and sample preparation seeDoerr et al. (2005a).

Water droplet penetration time method

Water droplet penetration times for soils at 20 °C�WDPT20� were determined as follows. All sub-sampleswere equilibrated in a controlled atmosphere of 20 °C and45–55% relative humidity for 24 h to avoid any influenceof changing atmospheric conditions on test results (Doerret al., 2002). Samples (¾10 g) were placed in plasticdishes (50 mm diameter ð 10 mm depth) and the surfacesmoothed by gently tapping and shaking the dish onthe bench top. Five drops of distilled water (¾80 µl)from a Nalgene droplet dispenser bottle were placed onthe soil surface from a height of less than 5 mm, andWDPT20 obtained as the time recorded for complete dropinfiltration for each drop. WDPT105 values, for samplesexposed to heating at 105 °C for 24 h as suggested byMa’shum and Farmer (1985) and Dekker et al. (1998),were obtained in the same way.

Critical surface tension ��cx)

Using a series of ethanol/water solutions (or, for someexperiments, propan-1-ol/water solutions) of concentra-tions 0–6 M, to give solutions with surface tensions of33–72Ð8 mN m�1 at increments of �2 mN m�1 (Main-waring, 2004), drops (¾80 µl) of increasing ethanol con-centration (and thus decreasing surface tension) wereplaced on the levelled soil surface of each of three sub-samples (¾30 g) from Nalgene droplet dispenser bottles,from a height of less than 5 mm, until a concentrationwas reached which allowed rapid infiltration �<5 s� ofthe applied drop (Roy and McGill, 2002). �cx was takenas the mean surface tension of the most dilute solution,

-12

-10

-8

-6

-4

-2

00 5 10 15 20 25 30 35 40

ln(1

/WD

PT

)

(γLVw - γcx)/mJ m-2

Figure 1. ln(1/WDPT) against ��LVw � �cx� where x is a water/ethanolmix. Filled diamonds—WDPT20; open squares—WDPT105. Equation ofline of best fit is y D �0Ð23�š0Ð023�x � 1Ð01�š0Ð45�; (R2 D 0Ð81). Onlysamples with WDPT values <18 000 s are included, because at longer

times evaporation caused a significant loss in droplet volume

-10-9-8-7-6-5-4-3-2-10

0 5 10 15 20 25 30 35

ln(1

/SD

PT

)

(γsol - γcx)/mJ m-2

Figure 2. ln(1/SDPT) against (�sol � �cx�, for water/alcohol mixtureson four different soils: NL2—circles water/ethanol, open triangleswater/propan-1-ol; NL3—diamonds water/ethanol, open triangles water/propan-1-ol; AU1—filled triangles water/ethanol; AU3—squares water/ethanol. �cx measured using aqueous solutions of either ethanol orpropan-1-ol as appropriate. Equation of line of best fit is

y D �0Ð19�š0Ð012�x � 3Ð07�š0Ð15�; (R2 D 0Ð80)

which penetrated the surface within 5 s. Error estimatesfor Figures 1 and 2 are one standard deviation.

THEORETICAL APPROACH

Following the results of many previous studies we beginwith the assumption that soil water repellency is causedprimarily by a layer of organic material adsorbed onthe surface of the soil particles (Ma’shum et al., 1988;Bisdom et al., 1993; Franco et al., 2000). We neglectthe influence of any other factors such as particle sizeand size distribution (McHale et al., 2005) and variationin the composition and distribution of soil minerals.We expect our experimental dataset of sandy soils ofsimilar particle size and size distribution to match theseassumptions quite well.

For instantaneous immersion wetting of the soil usingliquid x:

�LVix � �SLi

x � �SVix �1�

where � indicates interfacial tension; LV, SL and SVindicate liquid/vapour, solid/liquid and solid/vapour inter-faces respectively; and superscript i indicates the situationat the instant the drop is added (Jaycock and Parfitt,1981). MED or CST studies provide the critical surface

Copyright 2007 John Wiley & Sons, Ltd. Hydrol. Process. 21, 2248–2254 (2007)DOI: 10.1002/hyp

2250 P. DOUGLAS ET AL.

Tabl

eI.

Ori

gins

,cha

ract

eris

tics,

WD

PT,�

c xan

dto

talo

rgan

icca

rbon

(TO

C)

(Doe

rret

al.,

2005

a)fo

rso

ilsa

mpl

esus

edin

this

stud

y.W

DPT

are

the

mea

nof

½fiv

em

easu

rem

ents

,WD

PTD

0s

spec

ifies

imm

edia

tepe

netr

atio

n(i

.e.<

1s)

.�c x

isth

em

ean

ofth

ree

mea

sure

men

tsfo

rw

ater

/eth

anol

mix

ture

s;an

erro

res

tim

ate

indi

cate

sth

atth

etr

ansi

tion

tow

etta

bili

ty,a

sth

esu

rfac

ete

nsio

nof

the

solu

tion

was

decr

ease

d,w

asno

tth

esa

me

for

all

thre

esu

b-sa

mpl

es.

Cou

ntry

ofor

igin

code

s:PT

–Po

rtug

al,

NL

–N

ethe

rlan

ds,

UK

–U

nite

dK

ingd

om,

AU

–A

ustr

alia

,G

K–

Gre

ece,

(aC

code

indi

cate

sa

wet

tabl

e‘c

ontr

ol’

soil

)

Cod

eR

egio

nL

atit

ude,

Lon

gitu

deV

eget

atio

nty

peSa

mpl

ing

dept

h(c

m)

Mea

ndi

am.

dist

ribu

tion

wid

th(m

m)

WD

PT20

(s)

�c x

20

�mN

m�1

�W

DPT

105

(s)

�c x

105

�mN

m�1

�T

OC

�gkg

�1�

PT1

Ave

iro

40°

320 N

,8°

460 W

Cab

bage

0–

100Ð5

7;0Ð2

316

Ð1š

2Ð061

Ð048

Ð3š

7Ð161

Ð06Ð3

š0Ð1

PT2

Ave

iro

40°

190 N

,8°

460 W

Euc

alyp

tus

glob

ulus

0–

100Ð4

6;0Ð1

672

105

53Ð0

1800

045

Ð010

Ð3š

1Ð5PT

3A

veir

o40

°20

0 N,

8°47

0 WP

inus

pina

ster

0–

100Ð4

7;0Ð1

62Ð5

š0Ð2

72Ð8

15Ð7

š4Ð2

65Ð7

š4Ð6

0Ð6š

0Ð4PT

CA

veir

o40

°20

0 N,

8°47

0 WP

inus

pina

ster

0–

100Ð5

0;0Ð1

70

72Ð8

2Ð6š

1Ð372

Ð80Ð1

š0Ð2

NL

1H

olla

nd51

°48

0 N,

3°54

0 EG

rass

&m

oss

0–

100Ð2

7;0Ð2

231

47š

496

44Ð0

1800

040

Ð036

Ð2š

2Ð9N

L2

Hol

land

51°

480 N

,3°

540 E

Gra

ss&

mos

s10

–20

0Ð23;

0Ð10

2063

š96

48Ð7

š1Ð5

1800

044

Ð7š

0Ð65Ð9

š0Ð3

NL

3H

olla

nd51

°48

0 N,

3°54

0 EG

rass

&m

oss

20–

300Ð2

2;0Ð0

844

67š

1319

53Ð0

4063

š55

452

Ð7š

0Ð60Ð8

š0Ð2

NL

CH

olla

nd51

°48

0 N,

3°54

0 EG

rass

&m

oss

30–

400Ð2

2;0Ð0

70

72Ð8

072

Ð8no

nede

tect

edU

K1

Gow

er51

°35

0 N,

4°06

0 WD

une

herb

s&

gras

ses

0–

50Ð3

3;0Ð0

875

109

41Ð0

š1Ð2

2572

š23

641

Ð3š

1Ð211

Ð4š

2Ð7U

K2

Gow

er51

°35

0 N,

4°06

0 WT

urf

gras

s0

–5

0Ð30;

0Ð08

76Ð8

š5Ð8

61Ð0

1800

052

Ð7š

0Ð67Ð8

š1Ð0

UK

CG

ower

51°

350 N

,4°

060 W

Un-

vege

tate

d0

–5

0Ð39;

0Ð12

072

Ð80

72Ð8

3Ð1š

0Ð6A

U1

Nar

acoo

rte

36° 2

60 S,14

0°40

0 EPa

ddoc

k0

–10

0Ð25;

0Ð16

608

š80

49Ð0

1525

š14

945

Ð011

Ð7š

1Ð0A

U2

Nar

acoo

rte

36° 2

60 S,14

0°40

0 EPa

ddoc

k0

–10

0Ð29;

0Ð23

10Ð7

š1Ð2

67Ð7

š5Ð8

21Ð7

š0Ð6

61Ð0

14Ð4

š0Ð5

AU

3N

arac

oort

e36

° 260 S,

140°

410 E

Padd

ock

0–

100Ð2

3;0Ð1

150

7845

4421

š25

744

Ð3š

0Ð66Ð0

š1Ð1

AU

CN

arac

oort

e36

° 300 S,

140°

420 E

Padd

ock

0–

100Ð2

4;0Ð1

40

72Ð8

072

Ð82Ð2

š0Ð4

GK

1T

hrac

e41

°07

0 N,

25°

070 W

Perm

anen

tpa

stur

e0

–12

0Ð45;

0Ð32

43Ð1

š2Ð4

56Ð0

196

š46

51Ð3

š1Ð2

10Ð4

š0Ð2

GK

2T

hrac

e40

°57

0 N,

20°

190 W

Perm

anen

tpa

stur

e0

–5

0Ð47;

0Ð24

136

š19

50Ð0

1526

š13

947

Ð021

Ð2š

0Ð8G

K3

Thr

ace

40°

560 N

,24

°59

0 WD

une

herb

s&

gras

ses

0–

190Ð3

5;0Ð1

48Ð8

š1Ð4

68Ð3

š4Ð6

12Ð4

š1Ð0

63Ð0

1Ð6š

0Ð5G

KC

Thr

ace

40°

550 N

,24

°53

0 WW

inte

rw

heat

0–

190Ð7

0;0Ð3

62Ð3

š0Ð7

72Ð8

11Ð0

š3Ð6

63Ð0

6Ð2š

0Ð7

Copyright 2007 John Wiley & Sons, Ltd. Hydrol. Process. 21, 2248–2254 (2007)DOI: 10.1002/hyp

KINETICS AND ENERGETICS OF SOIL WATER REPELLENCY 2251

tension, �cx, i.e. the maximum liquid surface tension atwhich instantaneous wetting occurs; and for which

�cx D �SLix � �SVi

x �2�

where x is normally a mixture of ethanol and water. Onlyif �SLw � �SVw ½ �LVw (72Ð8 mN m�1 (or mJ m�2))(where superscript w signifies water) will wet soil instan-taneously. In all other cases instantaneous wetting is notexpected but after a time delay ranging from a few sec-onds to a few hours the soils may be wetted. For this tohappen there must be a convergence such that

�SLw � �SVw D �LVw �3�

This can be achieved by:

(1) Diffusion of surfactant material from soil to water toreduce �LVw.

(2) Adsorption of water at the soil organic interfacebecause of the high water vapour pressure next tothe droplet causing an increase in hydration and anincrease in �SLw � �SVw.

(3) The proximity of water and water vapour causinga change in the surface arrangement of organicmolecules to increase �SLw � �SVw.

(4) Any combination of the above.

Process (1) is unlikely to be important because waterpassed through, and wetting, a column of soil wasshown not to have a shorter WDPT on fresh soil(Llewellyn et al., 2004). It seems likely that the majorchanges with time involve �SLw and �SVw (with �SLw

being the dominant term because of the much lowerconcentration of molecules in a gas). The question is howcan measurement of �cx using ethanol water mixturesbe related to the rate of change of �SLw � �SVw fromthe initial condition to the condition at which the soilbecomes wettable; i.e. how can �cx be related to WDPT?If we make the assumption that the major effect of addingethanol to water is a change in �LV rather than a changein either �SL or �SV; i.e. the major effect of ethanol is todisrupt hydrogen bonding in water rather than changingthe energy of either the surface-liquid or surface-vapourinteractions from those with water alone, then we cansay,

�cx D �SLix � �SVi

x D �SLiw � �SVi

w �4�

and use �cx as a measure of the initial state of the surfacewith respect to water.

From the thermodynamic formulation of transitionstate theory the relationship between the free energy ofactivation, G‡ and the rate is given by Equation (5)(Laidler, 1987; Laidler and Meiser, 1982):

rate / EXP�-G‡/kT� �5�

where T is temperature and k the Boltzmann constant.G‡ can be considered as a sum of terms and for a seriesof reactions in which only one term changes, e.g. for our

case the surface free energy of activation G‡s, we can

separate out the changing term to give Equation (6),

rate D A.EXP�-G‡s/kT� �6�

where A is the rate for the reference process for which,by definition, G‡

s is zero. Since wetting by waterwill occur when �SL � �SV D �LVw then we mightexpect �LVw � �cx, to be related to G‡

s. The simplestapproach is a linear relationship:

G‡s D f��LVw � �cx�/N �7�

where N is the number of molecules per unit area, and f isa constant relating the required change in surface energyto the overall change in energy as the soil shifts fromrepellent to wettable. The rate for the wetting of soil isgiven by 1/WDPT, and with Equations (6) and (7), andtaking logarithms, this gives

ln�1/WDPT� D lnA � [f ��LVw � �cx�/NkT] �8�

Here A is 1/WDPT for our reference process i.e. thatfor which �cx D �LVw which, with our experimental pro-tocols for determining �cx, means complete penetrationwithin 5 s. Relationships between kinetic and thermody-namic parameters are well known in chemistry wherethey are examples of more general Linear Free EnergyRelationships (Gould, 1959; Moore and Pearson, 1981).Such LFERs are usually written as:

log�k/kref� D �� �9�

where k is the kinetic parameter for the system; kref isthe kinetic parameter for a reference system/compound,� is the thermodynamic parameter and � is a constantreflecting the sensitivity of the kinetics to changes inthe thermodynamic parameter. This is of the same formas Equation (8), with k/kref D WDPT0/WDPT (whereWDPT0 is 1/A), � D �LVw � �cx, and � D f/NkT.

RESULTS AND DISCUSSION

Relationship between WDPT and �cx

Figure 1 shows a plot of ln(1/WDPT) against ��LVw ��cx�. The plot shows a good linear correlation (R2 D0Ð81) consistent with Equation (8). There is no systematicvariation in results from soils at 20 °C and those heated to105 °C. The slope of Figure 1 is given by f/NkT. We canestimate N from the number of molecules in a monolayer.Using a molecular area of 2Ð0 ð 10�19 m2 (Moore, 1972;Shaw, 1995) which is a typical value for a close packedlong chain carboxylic acid, then N D 5Ð0 ð 1018 m�2.This gives 1/NkT D 49 J�1 m2 and from the experimen-tal slope of �230 �š23� J�1 m2, f D 4Ð6�š0Ð46�. Thisvalue is interesting because, from a comparison of thenumber of nearest neighbours in bulk and at a surface,the surface free energy is ca 25% the bulk cohesiveenergy (Castellan, 1971), and so it seems that the sur-face free energy contribution to the activation energy for

Copyright 2007 John Wiley & Sons, Ltd. Hydrol. Process. 21, 2248–2254 (2007)DOI: 10.1002/hyp

2252 P. DOUGLAS ET AL.

a given soil is, to a reasonable approximation, the differ-ence between the cohesive energy of the molecular filmadsorbed on the soil and that of water.

Effect of solution composition on penetration time

To establish whether it is �LV or the nature ofthe substance added to reduce the surface tensionwhich determines Solution Droplet Penetration Time(SDPT), we measured SDPT20 for four soils of dif-ferent repellency characteristics (GK3-slightly repellent,GK2-strongly repellent, NL2-extremely repellent, NL3-extremely repellent; see Table I, based on the classifi-cation of Bisdom et al., 1993) using aqueous solutionsof short chain organic compounds all made up to give�LV D 60 mN m�1 (Mainwaring, 2004). For the lin-ear short chain alcohols examined (methanol, ethanol,propan-1-ol, butan-1-ol, pentan-1-ol), increasing chainlength made either no systematic difference to WDPT20

(GK2, NL2, NL3) or possibly decreased it slightly (GK3,GK2). Changing the functional group on a C3 chain froman alcohol (propan-1-ol) to a ketone (acetone) also hadno effect. Furthermore, in studies where the surface ten-sion of the solution droplet was varied by changing theconcentration of the organic component, application ofEquation (8) to data obtained using propan-1-ol gaveresults which are, within experimental error, the sameas those using ethanol (see below and Figure 2). Theseresults suggest that, for simple short chain compoundsat least, the molecular structure of the organic used toreduce �LV is not an important factor. (How far this con-clusion applies to other classes of compounds is uncer-tain. In a previous study (Mainwaring, 2004) it was foundthat a comparable solution of the aromatic compoundphenol gave a much shorter penetration time than thelinear short chain molecules used here.)

Increasing �LV from 72Ð8 (water) to 74Ð4 mN m�1

by addition of 1 mol dm�3 NaCl causes an increase inWDPT for these soils consistent with the increase in�LV. Use of an alkaline solution, 0Ð01 mol dm�3 NaOH,pH 12, �LV D 72Ð8 mN m�1, gave no change in WDPTwhich suggests that either ionisation of long chain acidsdoes not occur at the SL interface, or, if it does, it is notimportant in determining WDPT.

Effect of solution � on penetration time

In our interpretation, Figure 1 reflects the soil-to-soil variation in cohesive energy of the molecular filmadsorbed on the soil compared to that of water. Wecan use an alternative approach to introduce a changein this energy difference by working with a single soilbut varying the solution cohesive energy, i.e. use a rangeof solutions of differing ethanol, or propan-1-ol, contentand hence different LV surface tensions (�sol).

Following the previous argument we expect the rela-tionship between SDPT and �sol to be:

ln�1/SDPT� D ln A � [f��sol � �cx�/NkT] �10�

where �cx is taken as the interpolated solution surfacetension which gives a penetration time of 5 s when usingaqueous solutions of either ethanol or propan-1-ol asappropriate.

Figure 2 gives a plot of ln(1/SDPT) against (�sol ��cx), for penetration times using water/ethanol andwater/propan-1-ol mixtures of varying �sol for four dif-ferent soils: NL2, NL3, AU1 and AU3. The data show areasonably linear relationship with R2 D 0Ð80 for all datapoints, slope of 191 (š12) J�1 m2, and from this f D 3Ð9(š0Ð24). There are, however, differences in linearity fordifferent soils with data for AU1 showing excellent lin-earity while those for AU3 show some curvature, and,when considering the residuals as a whole, the over-all data show some curvature with increasing slope as(�sol � �cx) decreases, particularly at low (�sol � �cx).The reasons for this are not known, but it may be that, atthe relatively high concentrations of organics necessaryfor low �sol (for a solution with �sol D 40 mJ m�2 theethanol concentration is 17% by weight), there is disso-lution of adsorbed organics from the soil by the solutiondrop, or transfer of alcohol to the soil from the dropletvapour.

Interpretation and implications

In this analysis, �cx is determined by the cohesiveenergy of the organic film adsorbed on the soil, with a lowcohesive energy giving a low �cx; WDPT is determinedby the difference in cohesive energies between thisadsorbed film and water via Equation (8), with a largeenergy difference giving a long WDPT; and the soil-to-soil variations in both WDPT and �cx arise from thepresence of organic films of differing cohesive energieson the different soils. Three questions arise from this.

1. What chemical features are important in determiningthe cohesive energy of this film?The total amount of organic material itself is unlikely

to be of major significance unless this brings aboutchanges in intermolecular interactions. We have previ-ously shown that there is little correlation between eitherorganic mass by extraction or organic CH content byFTIR and WDPT for this soil set (Doerr et al., 2005a).

The cohesive energy will depend upon the natureof the compounds present, i.e. the functional groupsthey contain, their orientation, and the nature of theintermolecular forces between them. A low cohesiveenergy would be associated with non-polar materialswhere dispersion forces are the dominant intermolecularinteraction, while a high cohesive energy would beexpected for polar materials which have the possibility ofhydrogen bonding and dipole/dipole and dipole/induced-dipole interactions (Jaycock and Parfitt, 1981). Theintermolecular forces in this organic layer may showa high degree of orientation due to the presence ofthe soil/organic and organic/air interfaces, and somemolecular orientation with respect to these interfaceswould be expected. For example, a compound such as

Copyright 2007 John Wiley & Sons, Ltd. Hydrol. Process. 21, 2248–2254 (2007)DOI: 10.1002/hyp

KINETICS AND ENERGETICS OF SOIL WATER REPELLENCY 2253

a long chain acid, which has both polar and non-polarcharacter, would be expected to be orientated with thenon-polar chain extending out into the air interface.In this case, the potential high-energy interactions are‘locked in’ within the organic layer and the hydrocarbonchain is exposed at the surface so the film would beexpected to exhibit a low �cx suggesting a low cohesiveenergy in the interfacial region because what is initiallyexposed to the water droplet is a non-polar surface. Thereal situation is complicated by the presence of manypolar and non-polar compounds. How tightly these packtogether may also be important in determining how easythe layer as a whole can undergo the molecular motions,or hydration processes, necessary to shift from waterrepellent to wettable. It is known from emulsion sciencethat combinations of polar and non-polar compounds cangive strong well packed interfacial films (Shaw, 1995).

CST measures the initial �cx and the history of thesoil may be significant in influencing this. Exposureto organic solvents, water, dry air, or moist air, willall influence the orientation of intermolecular forcesin the organic film because they expose the layer toenvironments of different polarity. Elevated temperaturesand the presence of fluids will influence the rate at whichmolecular reorganisation within the film can occur. Thedegree and duration of exposure to these environmentalinfluences will determine both the equilibrium positionto which the system is moving, and how far the systemcan move towards this equilibrium position (e.g. Doerret al., 2005b) Thus we would expect a soil with anadsorbed organic layer not to have an inherent waterrepellency but rather the potential to display a rangeof repellencies, of narrower or wider range dependingupon composition, with the value of �cx at any one timereflecting the particular kinetically metastable state intowhich the organic materials adsorbed on its surface havebeen locked by the environmental history of the soil.2. What is the mechanism by which the interfacial film

is altered to allow wetting?It seems probable that this involves a change in struc-

ture of the organic film (Franco et al., 2000; Doerr et al.,2005a) to allow exposure of polar, and in particular,hydrogen bonding groups, to water at the surface, andpossibly hydration of polar groups within the film by pen-etrating water molecules. With this possible mechanismin mind the cohesive energy of the film, as determinedhere from �cx, may be regarded as some indicator ofthe combined magnitude and accessibility of polarity andhydrogen bonding potential of the organics present on thesoil. Figure 3 shows the free energy changes involved inwetting by water, together with a tentative reaction coor-dinate for the process.3. How applicable is this work to other soils of different

mineralogical characteristics?Wetting will depend upon the particle size, shape

and packing (McHale et al., 2005; 2007), since thiscontrols the contact area between solution drop and soiland the interstitial dimensions of the soil. It is alsoreasonable to assume that the cohesive energy of an

Figure 3. Reaction free energy profile for wetting by water of a mineralgrain with a coating of organics which include both non-polar andpolar groups, and schematic of suggested mechanism. In this tentativeinterpretation, the initial state, A, has some polar groups bound to thesoil grain surface, some polar groups bound together within the interfacialfilm, and non-polar groups at the organic/air (or water) interface. The stateof highest free energy, the kinetic transition state, B, has some of theinterfacial material in a relatively high energy molecular reorganisationinvolving the break-up of internally bound polar groups and the exposureof polar groups at the surface, and possibly some penetration of watermolecules. By state C, this process has progressed far enough for thechange in �SL to be sufficient to allow wetting of the organic surface.This is followed by penetration of water into both the interfacial materialand the interstices between mineral grains. To what extent the processis controlled by rearrangement of the interfacial material or penetrationof water into the interfacial material is unknown, nor is it known towhat extent water displaces bound polar organics at the soil grain/organicinterface, although wetting with water does not wash organic materialoff these soils. Note that the ‘movement’ of polar groups implied by thisschematic may well be as much an ‘unfolding’ of hydrogen bonding sitesto penetrating water as diffusion of molecular material through the organiclayer. State D shows the final situation in which water has penetrated theorganic later (for simplicity, polar groups other than those adsorbed atthe soil surface are not identified in this diagram since we have no reasonto suggest a specific location for them). The energy difference between

A and D is not known

adsorbed film will depend upon the mineral surface towhich it is adsorbed. However, provided these factorsare kept reasonably constant for a set of soils wewould expect the theory proposed here to be applicable.What is more difficult to assess is the behaviour ofsoils with a heterogeneous mineral composition and/ora wide distribution of particle size and/or adsorbedfilm compositions and/or thicknesses. Here, the wettingcharacteristic of each individual soil grain may differgreatly, and it is not certain that this would lead to someeasily understandable averaged kinetic behaviour.

CONCLUSIONS

A linear free energy relationship between SDPT andthe difference between the surface tension of the solu-tion droplet and CST has been derived. WDPT, SDPT(ethanol/water and propan-1-ol/water), and CST measure-ments for a group of sandy soils support this analysis.The data suggest that the surface free energy contributionto the free energy of activation is given by the differ-ence between the cohesive energy of the molecular filmadsorbed on the soil and that of the wetting solution.

Copyright 2007 John Wiley & Sons, Ltd. Hydrol. Process. 21, 2248–2254 (2007)DOI: 10.1002/hyp

2254 P. DOUGLAS ET AL.

In this interpretation: CST ��cx� is determined by thecohesive energy of the organic film adsorbed on thesoil; WDPT is determined by the difference in cohesiveenergies between this adsorbed film and water, with alarge energy difference giving a long WDPT; and soil-to-soil variations in both WDPT and �cx are due to organicfilms of different cohesive energies on the different soils.For aqueous solutions of short chain linear alcohols andketones, SDPT depends on the surface tension of thesolution used but is independent of the compound usedto control the surface tension.

An important implication arising from this work is thatsoil with an adsorbed organic layer is unlikely to have aninherent water repellency but rather the potential to dis-play a range of repellencies, of narrower or wider rangedepending upon composition, with the value of �cx at anyone time reflecting the particular kinetically metastablestate into which the organic materials adsorbed on itsparticle surfaces have been locked by the environmentalhistory of the soil.

ACKNOWLEDGMENTS

KAM would like to thank the University of WalesSwansea for a Postgraduate Studentship. SHD acknowl-edges funding from the NERC (Advanced FellowshipNER/J/S/2002/00662 and Grant NE/C003985/1). Wewould like to thank Aquatrols Corp. for financial support.

REFERENCES

Bisdom EBA, Dekker LW, Schoute JFTH. 1993. Water repellency ofsieve fractions from sandy soils and relationships with organic materialand soil structure. Geoderma 56: 105–118 DOI:10.1016/0016-7061(93)90103-R.

Castellan GW. 1971. Physical Chemistry . Addison-Wesley: Mas-sachusetts.

Cockcroft S, Topalidis S, Richardson DP. 1991. Water repellency in adry sclerophyll forest - measurements and processes. HydrologicalProcesses 5: 405–420.

Dekker LW, Ritsema CJ. 1994. How water moves in a water repellentsandy soil. 1. Potential and actual water-repellency. Water ResourcesResearch 30: 2507–2517.

Dekker LW, Ritsema CJ, Oostindie K, Boersma OH. 1998. Effect ofdrying temperature on the severity of soil water repellency. Soil Science163: 780–796.

Dekker LW, Ritsema CJ, Wendroth O, Jarvis N, Oostindie K, Pohl W,Larsson M, Gaudet JP. 1999. Moisture distributions and wetting ratesof soils at experimental fields in the Netherlands, France, Sweden andGermany. Journal of Hydrology 215: 4–22.

Doerr SH. 1998. On standardising the “Water Drop Penetration Time”and the “Molarity of an Ethanol Droplet” techniques to classify soilhydrophobicity: a case study using medium textured soils. EarthSurface Processes and Landforms 23: 663–668.

Doerr SH, Dekker LW, Ritsema CJ, Shakesby RA, Bryant R. 2002.Water repellency of soils: the influence of ambient relative humidity.Soil Science Society of America Journal 66: 401–405.

Doerr SH, Llewellyn CT, Douglas P, Morley CP, Mainwaring KA,Haskins C, Johnsey L, Ritsema CJ, Stagnitti F, Allinson G, Fer-reira AJD, Keizer JJ, Ziogas AK, Diamantis J. 2005a. Extraction ofcompounds associated with water repellency in sandy soils of dif-ferent origin. Australian Journal of Soil Research 43: 225–237DOI:101.071/SR04091.

Doerr SH, Douglas P, Evans R, Morley CP, Mullinger N, Bryant R,Shakesby RA. 2005b. Effects of heating and post-heating equilibrationtimes on soil water repellency. Australian Journal of Soil Research 43:261–267 DOI:10.1071/SR04092.

Franco CMM, Clarke PJ, Tate ME, Oades JM. 2000. Hydropho-bic and chemical characterisation of natural water repellentmaterials in Australian sands. Journal of Hydrology 231–232:47–58.DOI:10.1016/S0022-1694(00)00182-7.

Gould ES. 1959. Mechanism and Structure in Organic Chemistry . HoltRinehart Winston: New York.

Harper RJ, Gilkes RJ. 1994. Soil attributes related to water-repellencyand the utility of soil survey for predicting its occurrence. AustralianJournal of Soil Research 32: 1109–1124.

Jaycock MJ, Parfitt GD. 1981. Chemistry of Interfaces . Ellis Horwood:Chichester.

King PM. 1981. Comparison of methods for measuring severity of waterrepellence of sandy soils and assessment of some factors that affect itsmeasurement. Australian Journal of Soil Research 19: 275–285.

Laidler KJ. 1987. Chemical Kinetics (3rd edn). Harper and Row: NewYork.

Laidler KJ, Meiser JH. 1982. Physical Chemistry . Benjamin/CummingsPublishing Company, Inc: Menlo Park.

Llewellyn CT, Doerr SH, Douglas P, Morley CP, Mainwaring KA.2004. Soxhlet extraction of organic compounds associated withsoil water repellency. Environmental Chemistry Letters 2: 41–44DOI:10.1007/s10311-004-0069-4.

Mainwaring KA. 2004. Chemical characterisation and repellency-inducing effects of organic compounds isolated from sandy soils . PhDThesis, University of Wales, Swansea, UK.

Mainwaring KA, Morley CP, Doerr SH, Douglas P, Llewellyn CT,Llewellyn G, Matthews I, Stein BK. 2004. Water repellency of sandysoils from the Netherlands and the UK: the role of high molecular masspolar organic compounds. Environmental Chemistry Letters 2: 35–39DOI:10.1007/s10311-004-0064-9.

Ma’shum M. 1988. A molecular basis for the alleviation of waterrepellency in soils . PhD Thesis, University of Adelaide: Australia.

Ma’shum M, Farmer VC. 1985. Origin and assessment of waterrepellency of a sandy south Australian soil. Australian Journal of SoilResearch 23: 623–626.

McHale G, Newton MI, Shirtcliffe NJ. 2005. Water-repellent soil and itsrelationship to granularity, surface roughness and hydrophobicity: amaterials science view. European Journal Soil Science 56: 445–452DOI:10.1111/j.1365-2389.2004.00683.x.

McHale G, Shirtcliffe NJ, Newton MI, Pyatt FB. 2007. Implications ofideas on super-hydrophobicity for water repellent soil. HydrologicalProcesses 21(17): 2229–2238.

Moore MJ. 1972. Physical Chemistry . Longman: London.Moore JW, Pearson RG. 1981. Kinetics and Mechanism . Wiley-Inter-

science: New York.Morley CP, Mainwaring KA, Doerr SH, Douglas P, Llewellyn CT,

Dekker LW. 2005. Identification of organic compounds at differentdepths in a water repellent soil. Australian Journal of Soil Research43: 239–249 DOI:10.1071/SR04094.

Roy JL, McGill WB. 2002. Assessing water repellency using theMolarity of Ethanol Droplet (MED) test. Soil Science 167: 83–97.

Shaw DJ. 1995. Introduction to Colloid and Surface Science.Butterworth-Heinemann: Oxford.

Watson CL, Letey J. 1970. Indices for characterizing soil-waterrepellency based upon contact angle-surface tension relationships. SoilScience 108: 58–63.

Copyright 2007 John Wiley & Sons, Ltd. Hydrol. Process. 21, 2248–2254 (2007)DOI: 10.1002/hyp