Você está na página 1de 10

HYDROLOGICAL PROCESSES Hydrol. Process. 21, 2266 2275 (2007) Published online in Wiley InterScience (www.interscience.wiley.com) DOI: 10.1002/hyp.

6759

Development of soil water repellency in the course of isothermal drying and upon pH changes in two urban soils
Julia V. Bayer1 and Gabriele E. Schaumann2 *
2 1 School of Engineering, University of Wales Swansea, Singleton Park, Swansea, SA2 8PP, UK Institute of Sciences, Department of Chemistry/Section Organic and Environmental Chemistry, University of Koblenz, Universit tsstr. 1, D-56072 a Koblenz, Germany

Abstract:
The potential inuence of pH on water repellency in soils has already been mentioned in some studies, but no clear correlation between these parameters has been found to date. In addition, although correlations of water content and water repellency have been found in numerous studies, the inuence of drying and subsequent storage conditions on water repellency are still unclear. In this study, a series of samples showing water repellent and wettable conditions respectively at eld moist states from two urban locations were compared regarding two main aspects: (i) the inuence of articial pH-changes on their wetting behaviour and (ii) sample wettability changes during drying for different drying temperatures. The assumption made in a study by Hurra and Schaumann (2006) on the same sample site, that differences in wettability of closely neighboured samples from one sample location can be interrelated with differences in pH, was conrmed. However, in contrast to the assumption that rising pH will improve the wettability and decreasing pH will intensify water repellency, we found a maximum in water repellency at a pH above the initial pH. The results from drying samples at different temperatures conrmed the dependence of water repellency development on the drying temperature. Additionally, we could conrm the conclusion from previous studies that the water content alone cannot explain the water contentWDPT (water drop penetration time) relationship. The results from drying and pH changes show that some location specic factors, like the number of pH active functional groups, may be relevant for sample wettability. As additional mechanism, which may be partly antagonistic to the inuence of the water content, we assume temperature-dependent conformational changes in SOM (Soil Organic Matter). It seems that aspects of these conformational changes may additionally be governed by pH in one of our two locations. Copyright 2007 John Wiley & Sons, Ltd.
KEY WORDS

soil water repellency; pH; drying temperature; WDPT

Received 15 December 2005; Accepted 7 July 2006

INTRODUCTION Soil water repellency is a widespread phenomenon with considerable environmental impact. It can lead to preferential ow and, therefore, to irregular moisture patterns with high spatial variability (Ritsema and Dekker, 1994; T umer et al., 2005) or to the reduction in inltration a rate, which in turn can lead to increased overland ow and soil erosion (Doerr et al., 2000). The modied water uptake, arising from water repellency, can further induce poor plant growth (Wallis and Horne, 1992; Doerr et al., 2000). Numerous studies have been conducted to date on various aspects of soil water repellency, focussing, e.g. on its spatial and geographical distribution or the inuence of water content (e.g. de Jonge et al., 1999) and soil organic matter (SOM) characteristics (e.g. Capriel et al., 1995) on water repellency expression. Despite these numerous studies, the mechanisms controlling water repellency are not yet fully understood. Water content has been identied as one of the main factors inuencing water repellency. Water repellent soil
* Correspondence to: Gabriele E. Schaumann, Institute of Sciences, Department of Chemistry/Section Organic and Environmental Chemistry, University of Koblenz, Universit tsstr. 1, D-56072 Koblenz, Germany. a E-mail: schaumann@uni-koblenz.de Copyright 2007 John Wiley & Sons, Ltd.

conditions are usually associated with low water contents, whereas at higher soil water contents, wettable or only slightly water repellent conditions dominate (e.g. Ziogas et al., 2003; Leighton-Boyce et al., 2005; Keizer et al., 2007). It is widely accepted that water repellency may be present below, but is absent above a critical water content (e.g. Dekker et al., 2001). Therefore, it is assumed that bringing the samples to similar water contents can compensate the original differences in their wettability (Dekker et al., 2001; de Jonge et al., 1999). However, some samples keep their initial differences in wettability when changing the moisture status or drying at different conditions (Hurra and Schaumann, 2006). In contrast to the term actual water repellency for eld conditions, the term potential water repellency is used to describe the wettability of dried samples (Dekker et al., 2001). However, differing temperatures have been applied to determine potential water repellency in numerous studies. Dekker et al. (2001) observed that drying as well as increasing drying temperatures, especially in top layer sandy soils, raised the severity of water repellency. In contrast, e.g. Ziogas et al. (2003) showed that for some agricultural elds in Greece, increased drying temperatures resulted in a complete wettability of the initially water repellent samples. Complementary to this

DRYING AND PH EFFECTS ON SOIL WATER REPELLENCY

2267

observation, de Jonge et al. (1999) and Goebel et al. (2004) found that samples rst become water repellent when dried, but their wettability re-increases at very low water contents after oven drying. The tendency that water repellency decreases upon drying was also found by Doerr et al. (2006) following air drying of samples. This was explained by the absence of adsorbed water molecules on the particle surface; these particles then have a high surface free energy resulting in a strong attraction of the solid surface and the water (Goebel et al., 2004). With increasing water content, the water sorbed on the particle surface evolves to a water lm. This water lm reduces the surface free energy and, therefore, the wettability of the soil (Derjaguin and Churaev, 1986; Vogler, 1998). Above a certain water content, the properties of the particle surface are overbalanced by the properties of the water lm, which become more and more similar to those of free water (Clifford, 1975; Zettlemoyer et al., 1975; Goebel et al., 2004). These studies show that the water content alone is not the sole determinant for water repellency expression in soils prone to repellency. Therefore, additional factors like the SOM had been suggested to inuence water repellency, e.g. organic substances such as aliphatic hydrocarbons and amphiphilic molecules have been related to water repellency (e.g. Mashum and Farmer, 1985; Franco et al., 2000; Horne and McIntosh, 2000). Capriel et al. (1995) suggested a correlation between water repellent character of SOM and the content of aliphatic hydrocarbons for arable soils in Bavaria using diffuse reectance infrared Fourier transform spectroscopy (DRIFT). In contrast, Hurra and Schaumann (2006) did not nd such a correlation using Fourier transforminfrared spectroscopy (FT-IR) for urban soil samples. However, the contribution of individual substance classes and the total organic carbon content to water repellency remains uncertain. The presence of certain organic substances may thus impart water repellency, but additional factors are decisive for its expression. Mashum and Farmer (1985) or more recently Hurra and Schaumann (2006) suggested that the conformation of organic molecules is relevant for the water repellency. In case of amphiphilic molecules the orientation of the hydrophilic functional groups relative to the particle surface controls the wetting character of the soil (Horne and McIntosh, 2000). Ellerbrock et al. (2005) further considered the effectiveness of SOM for the expression of water repellency introducing a weighing factor that depends on the ratio of soil organic carbon (SOC) to mineral matrix (SOC/mineral in g/g). This ratio is considered of high importance as it may describe the orientation possibilities of SOM molecule: with a low SOC/mineral ratio, hydrophilic functional groups are supposed to be orientated to the mineral surface and hydrophobic groups outward, leading thus to water repellency. With higher SOC/mineral ratio SOM molecules may be forced in a more upright position, so that also hydrophilic groups can be outward orientated, which may lead to higher wettability (Ellerbrock et al., 2005). Rearrangement in the orientation of such functional groups is assumed to follow
Copyright 2007 John Wiley & Sons, Ltd.

slow kinetics, but may lead to major changes in surface wettability (Mashum et al., 1988; Roy and McGill, 2000). To identify the processes occurring during wetting, process orientated studies are necessary. The investigation of process kinetics is one possibility to determine the nature of the participating sub-processes. For chemical and physical reactions, an energy barrier has to be overcome to initiate the reaction. With rising temperatures the energy input increases and, therefore, the energy barrier can be overcome, and reactions proceed faster with higher temperatures. Isothermal drying of soil samples may induce conformational changes in SOM after water removal, and increasing drying temperatures may accelerate these processes. Upon remoistening dried samples, hysteretic behaviour of the water drop penetration time (WDPT) often occurs (Doerr and Thomas, 2000; Quyum et al., 2002; Hurra and Schaumann, 2006). Following the assumption that drying activates conformational changes, the isothermal hysteresis of WDPT may be caused by the relatively slower rearrangement compared to the conformational changes induced by drying (Hurra and Schaumann, 2006). Besides the water content and SOM, the soil pH may affect the wettability (e.g. Wallis and Horne, 1992). It has been observed, that the water uptake of water repellent soils was improved by the application of lime or sodium hydroxide in high concentrations (e.g. Karnok et al., 1993; Blackwell, 1997). Additionally, Roberts and Carbon (1971), Steenhuis et al. (2001) and Hurra and Schaumann (2006) observed that water repellent samples were more likely to have lower pH than wettable samples. Roberts and Carbon (1971) found this relation for sand samples from south-western Australia, Steenhuis et al. (2001) for more than 3000 garden and agricultural samples from New York State and Hurra and Schaumann (2006) for 47 samples from an inner city park in Berlin. On the other hand, no signicant correlation between water repellency and pH has been found for samples from a former sewage eld in Berlin (Hurra and Schaumann, 2006). One approach to explain the inuence of the pH on the wettability is to relate the pH to the state of the functional groups of SOM (e.g. Tschapek, 1984) and to the ratio of dissolved to particulate organic matter (e.g. Babejova, 2001): as fulvic acids are soluble in a wide pH range and humic acids are only soluble at high pH, it is more probable that fulvic acids are leached out and humic acids remain bound to the particle surface. In principle, both, humic and fulvic acids contain carboxylic and phenolic groups and, hence, are potentially hydrophilic (Chen and Schnitzer, 1978). The increasing deprotonation of those polar functional groups with increasing pH, however, may lead to an increase in wettability (Terashima et al., 2004; Hurra and Schaumann, 2006). Terashima et al. (2004) reported that at pH 4 only 31% of the carboxylic groups in humic acids were deprotonated, whereas at pH 7 all carboxylic groups were dissociated. In addition, humic acids rather than fulvic acids are supposed to have the character of amphiphilic
Hydrol. Process. 21, 2266 2275 (2007) DOI: 10.1002/hyp

2268

J. V. BAYER AND G. E. SCHAUMANN

molecules, and thus, are more likely to cause water repellency. The addition of humic acids to soil samples has in fact been shown to increase their water repellency (Tschapek, 1984; Babejova, 2001). Furthermore, stable humic acid complexes were found to be mainly responsible for the water repellency of sand samples (Roberts and Carbon, 1972). The outward orientation of nonpolar groups from humic acids on soil particle surfaces may thus be responsible for soil water repellency (Tschapek, 1984). Regarding the inuence of pH, the question arises whether a low pH and its inuence on the state of SOM is a cause or a consequence of soil water repellency. The aim of this study was to elucidate the inuence of pH changes and drying processes on the wettability of soil samples taken from two urban locations in Berlin (Buch and Tiergarten). To identify the above discussed inuences on the development of wettability, such as water content and conformation of SOM (Goebel et al., 2004; Hurra and Schaumann, 2006), we investigated the development of the wettability in the course of the drying process at different temperatures. Therefore, sub-samples were taken within the drying process and their wettability was determined by the WDPT test (e.g. Letey, 1969). The second part of this study focussed on the inuence of the pH on sample wettability. We induced articial pH changes in the samples and observed their effect on the WDPT. Following previous studies, we hypothesized that raising the pH of the samples would lead to a decrease of water repellency and reducing the pH would increase the repellency (Karnok et al., 1993; Steenhuis et al., 2001).

MATERIALS AND METHODS Soil samples The soil samples were taken from 10 to 20 cm depth of the research locations Berlin-Tiergarten and BerlinBuch, which have been intensively investigated in a series of previous studies (e.g. Reemtsma et al., 2003; T umer et al., 2005; Hurra and Schaumann, 2006; Diehl a and Schaumann, 2007). Berlin-Buch is a former sewage

eld which was used for more than 80 years until 1985. Berlin-Tiergarten is an inner-city park built in the 17th century. After bomb damages and deforestation during World War II it had to be rearranged in the 1950s. From Buch, we investigated seven samples with varying degrees of water repellency, and from Tiergarten four samples were investigated. In order to minimize the effect of heterogeneity and in accordance with the suggestion by Hurra and Schaumann (2006), we examined sample pairs taken at directly adjacent positions. The sample pairs comprised of one sample that was wettable and one that was water repellent at eld moist conditions. In this contribution, the results are shown exemplarily for one representative sample pair from each location, as the results for the other samples are comparable. The sample pairs are named T for Tiergarten and B for Buch with the sufxes w for actually wettable and h for actually water repellent (hydrophobic) samples. In the following the terms initially water repellent and initially wettable are used, if the text refers to the eld state of the samples. The characteristics of the samples are shown in Table I. The water contents (WC) of all samples were determined gravimetrically and are related to the dry mass. FTIR spectras of the samples were taken and the CH/CO groups ratio is given in Table I. Additionally, data from soil extracts as given in Hurra and Schaumann (2006) are listed in Table I. These are based on other samples, however, they are comparable to our samples as they were taken on the same sample sites. Further sample properties of comparable samples from the same sample sites are shown in T umer et al. (2005), Hurra and a Schaumann (2006). Determination of wettability The wettability of the samples was determined by the WDPT test (e.g. Letey, 1969). When possible three water drops of 50 l where placed on the sample surface; if the available amount of individual sub-samples was too small to place three drops on the surface only one drop was used. To ensure comparable conditions for all

Table I. Water content, WDPT, organic carbon content, pH and CH/CO ratio (measured by FT-IR) of the soil sample pairs Th/Tw and Bh/Bw; properties of soil extracts as given in Hurra and Schaumann (2006). Further information about the locations are given in T umer et al. (2005), Hurra and Schaumann (2006) a Tiergarten Th Water content [%] (dry mass) WDPT [min] pH Corg [%] CH/CO ratio DOC/SOC [mg/g]a Absorption254nm /DOC [L/(cmg)]a Surface tension [mN/m]a LC-OCD main peak position [kg/mol]a Conductivity [S/cm]a
a Data

Buch Tw Bh 475 197 46 37 0104 0001 104 003 46 6 679 05 42 01 97 5 Bw 975 0 46 48 0300 0001 093 002 47 5 715 05 43 01 105 18

40 258 40 221 0348 0001 045 001 47 12 590 05 23 01 134 14

143 0 43 166 0192 0002 038 003 50 17 647 05 38 02 76 1

taken from Hurra and Schaumann (2006), see there for details.

Copyright 2007 John Wiley & Sons, Ltd.

Hydrol. Process. 21, 2266 2275 (2007) DOI: 10.1002/hyp

DRYING AND PH EFFECTS ON SOIL WATER REPELLENCY

2269

experiments and to minimize the changes of environment, WDPT experiments were carried out at 20 C in a cooled incubator at a relative humidity of (50 5)%. WDPT above 5 min were investigated with the aid of a camera, taking pictures every 3 min. For WDPT up to one hour, the time of complete penetration was determined with an uncertainty of three pictures (e.g. 9 min), while above 1 hour the uncertainty was ve pictures (e.g. 15 min). Effect of drying temperature on wettability 10 g of each sample were oven-dried at 35 C, 65 C and 105 C in covered containers with a small hole in the lid to slow down the drying process. In the course of drying, sub-samples were taken every hour within the rst 6 h and with increasing intervals after 6 h, and their WDPTs were determined at 20 C after cooling the samples for 20 min. This short period of time was long enough for cooling to 20 C and short enough to minimize additional transformations in the sample. The development of WDPT was monitored for at least 24 h, depending on the drying temperature. Effect of the pH on wettability The eld pH was determined with a CaCl2 solution (0,01 mol l 1 ) directly after sampling. For the pH determined, the soil : solution ratio was 1:25 and the equilibration time 1 hour. The pH of the samples was modied articially in the laboratory by addition of NaOH and HNO3 in different concentrations (each 005 mol l 1 , 025 mol l 1 , 05 mol l 1 , 25 mol l 1 , 5 mol l 1 ). 2 ml of the respective solutions were added to 10 g of a sample and mixed intensively. To determined the inuence of the addition of liquids on the WDPT, a control sample was used: sub-samples were mixed with 2 ml of water. All samples were incubated for 1 week at 20 C in closed containers and then air-dried for 2 days in a cooled incubator (temperature 20 C, relative humidity (50 5%), in order to achieve constant weight before the rst WDPT measurement. The samples were then incubated for another week in a water saturated atmosphere and the WDPT were re-evaluated.

Figure 1. Water drop penetration time of initially water repellent and initially wettable samples from Tiergarten and Buch as a function of water content during oven drying at 35 C

RESULTS AND DISCUSSION Effect of drying temperature on wettability In Figure 1, the WDPT of the samples Th, Tw, Bh and Bw are shown for the drying temperature of 35 C at different drying stages. The changes in WDPT are a function of both water content and drying time. The experiment starts at high water content and, thus, on the right side of the diagram. The WDPT of the initially wettable samples from both sites rst increased strongly with decreasing water content and reached a maximum water repellency. Thus, sample Tw reached a WDPTmax of 159 15 min at WC D 61% and Bw a WDPTmax of 59 9 min at WC D 08%. The WDPT then decreased again, until all
Copyright 2007 John Wiley & Sons, Ltd.

samples were completely wettable. The maximum water repellency of the Tiergarten sample Tw was distinctly more severe and appeared at higher water contents than that of the Buch sample Bw. In contrast, the WDPT of initially water repellent samples from Buch and Tiergarten rst continuously decreased with decreasing water content. Between water contents of 3% and 1%, the WDPT of both water repellent samples uctuated (40 min), but reincreased to their eld WDPTs at water contents just below 1% (Figure 1). While the shape of the WDPTwater content curves were strongly affected by their initial wettability, no major differences between the development of WDPT for the two locations were found. The initial differences in minutes of WDPT between water repellent and wettable samples were broadly retained throughout the drying for the sample site Buch. The minimum difference was 59 min at a water content around 8% and higher for all other water contents. The WDPT of the wettable sample from Tiergarten approximated the minimum WDPT of the water repellent sample (159 min) at one water content (WC D 8%) during the whole drying cycle. At all other water contents a minimum difference of 82 min can be observed between the WDPTs of the water repellent and the wettable sample. The development of the WDPT upon oven drying at 65 C is illustrated in Figure 2. The WDPT of all wettable samples tended to increase with decreasing water content during the whole drying experiment. The sample Tw reached its maximal WDPT at a water content of 105% (WDPTTw D 114 15 min) within the rst hour of drying. The WDPT of the wettable sample Bw increased almost linearly with decreasing water content and reached its maximal water repellency (WDPTBw D 109 15 min) at a water content of 03%. The development of the WDPT of all initially water repellent samples was contrary to the WDPT from that of the initially wettable samples for both sample sites. With decreasing water content, the WDPT of the water repellent samples decreased slowly and approached a
Hydrol. Process. 21, 2266 2275 (2007) DOI: 10.1002/hyp

2270

J. V. BAYER AND G. E. SCHAUMANN

Figure 2. Water drop penetration time of sample pairs from Tiergarten and Buch as a function of water content during oven drying at 65 C

minimum at the end of the drying period (WDPTTh,min D 164 15 min; WDPTBh,min D 155 15 min). Figure 3 shows the dependence of the WDPT on water content (top) and the drying time (bottom) for drying at 105 C. All samples reached equilibrium dryness (gravimetrically determined WC <001%) after 2 h of drying, except for the sample Th. The WDPT of all samples continued changing despite constant water content. Both initially wettable samples tended to develop severe water repellency upon drying even though the WDPT of the sample Tw varied strongly between 107 15 min and 9 min at the end of the drying experiment. The WDPT of the sample Bw varied only slightly at a water content of 0% (WDPTmax D 51 9 min and WDPTmin D 42 9 min). The WDPT of the actually water repellent samples rst decreased with decreasing water content within

Figure 3. Water drop penetration time of sample pairs from Tiergarten and Buch as a function of water content during oven drying at 105 C Copyright 2007 John Wiley & Sons, Ltd.

the rst 2 h of drying. Between a drying time of 25 and 6 h, both water repellent samples showed a local maximum in WDPT (WDPTTh D 206 15 min and WDPTBh D 292 15 min). The nal WDPT of the samples Th and Bh (WDPTTh D 157 15 min and WDPTBh D 93 15 min) were lower than their initial WDPT (WDPTTh, eld D 258 15 min and WDPTBh, eld D 197 15 min). In general, drying changed the wettability of the soil samples at all drying temperatures and samples from both sites reacted in comparable manner at each respective temperature. However, water repellent and wettable samples showed differing development of WDPT. The nal wettability, as well as the development of the WDPT during the drying process, differ for each drying temperature. Even though an overall trend could be found for 65 C and 105 C, where actually wettable samples tended to become water repellent and actually water repellent samples became less repellent, the initial differences between the water repellent and wettable samples never disappeared fully (Figures 2 and 3). Thus, the higher water repellency persistence of the initial water repellent samples seems to be independent of the water content, and the differences in initial repellency were not completely overcome by drying at 105 C. These initial differences are most probably not caused by the amount of aliphatic hydrocarbons present in SOM, as our own data from FT-IR measurements (see Table I) and FT-IR data obtained by Hurra and Schaumann (2006), do not allow to differentiate clearly between the amount of CH groups present in water repellent and wettable samples. The WDPTwater content characteristics also show that the water content cannot be the sole reason for the observed changes in wettability. Especially for the initially wettable samples, differences in the WDPTs are most pronounced at low water contents (WC <1%, compare Figures 1 to 3). Thus, the wettability is partly dependent on the water content but additional factors may play a role. We speculate that sample wettability is governed by different mechanisms at high and low water contents. Also, drying temperature appears to have a stronger inuence at low, compared to high water contents. In the course of drying, the water repellency of the initially wettable samples passed through a maximum at low water contents (61% for Tw and 08% for Bw). This observation is in agreement with King (1981) and Goebel et al. (2004), although it was made only for the drying temperature 35 C. Under the assumption that the initial wettability of the water repellent samples corresponds with their maximum water repellency, their WDPT follow the same curve at 65 C: the WDPT are decreasing continuously with drying at 65 C. This assumption was made as the eld water contents of the water repellent samples were already very low. The development of the wettability of the water repellent samples at 35 C and 105 C did not support this interrelation, as no distinct decrease in water repellency was found. Figure 1 to 3, however, show that the development of the WDPT at
Hydrol. Process. 21, 2266 2275 (2007) DOI: 10.1002/hyp

DRYING AND PH EFFECTS ON SOIL WATER REPELLENCY

2271

each drying temperature showed distinct trends for each the wettable and water repellent samples. Therefore, we extracted trend lines from the overall behaviour in which we distinguish between drying temperatures and initial wettability. These general trend lines are shown in Figure 4, where the trend lines for the wettable samples at drying temperatures of 35 C and 65 C are depicted at the top. The general trend lines in the middle represent the development of the WDPT of the water repellent samples at 35 C and 65 C. At 35 C, the WDPTwater content curve has a slight minimum, but at low water contents, the water repellency is again comparable to its initial value. In contrast, the WDPTwater content curve at 65 C declines continuously. The bottom of Figure 4 shows the combined trend lines from drying at 35 C and 65 C resulting in one curve with a maximum water repellency, which corresponds with the observations of King (1981) and Goebel et al. (2004) that some samples express a maximum peak in water repellency at low water contents but are wettable at even lower ones. The samples used in our study showed additional development of wettability upon further heating, even when the water content remained constant (Figure 3). Therefore the model has to be supplemented with an additional factor. This factor is associated with time and assumed to be a slow, process,

with a temperature-dependent rate, partly competing with the inuence of the water content. To explain the initial differences between the water repellent and wettable samples, which never disappear completely, we assume that the expression of water repellency in samples that have the potential to display water repellent characteristics, is controlled by (1) a water lm as proposed by Goebel et al. (2004) and (2) variable soil particle surface characteristics. The latter are most probably controlled by the orientation of polar functional groups relative to the surface (e.g. Horne and McIntosh, 2000; Ellerbrock et al., 2005). These different orientations may inuence the water properties on the surface (Vogler, 1998) and control the thickness of the water lm (Clifford, 1975). The rst three molecular water layers adsorbed on a surface develop stronger hydrogen bonds than liquid water and are to some extent ice-like (Zettlemoyer et al., 1975). With increasing water content several water layers build-up and the properties of the adsorbed water approach those of free water (Zettlemoyer et al., 1975; Schreiber et al., 2001; Gruenberg et al., 2004). The permanent differences between water repellent and wettable samples are represented in Figure 5(a) in the curves 1 and 2, where curve 1 stands for the wettable

Figure 4. General trend lines for the development of the WDPT of all tested initially wettable samples (top) and all initially water repellent samples (middle) for the drying temperatures 35 C and 65 C. The bottom curve follows from the overlay of all other curves. The arrow on the left results from the observations of drying the samples at 105 C: most initially water repellent samples showed rst an increased wettability at water contents of 0% but developed water repellency later with increasing drying time. Most initially wettable samples developed water repellency Copyright 2007 John Wiley & Sons, Ltd.

Figure 5. (a) Dependence of water repellency on water content, with 1 depicting an initially wettable and 2 an initially water repellent sample. These curves can be slowly transformed into each other over time with a preferential direction (arrows). The roman numbers I to IV represent certain conformational states of SOM and the extent of coverage with water, specied in Figure 5(b). (b) I: a thick water lm is attached to the particle surface, independent of the conformation of SOM, corresponding with high water content. II: monomolecular water adsorption on the particle/SOM surface. III: water lm evolves gaps with further decreasing water content, for curve 1 free adsorption sites for water are gained, less are gained for curve 2 due to the conformation of SOM. IV: free adsorption sites for water at a water content of 0%. This results in more adsorption sites for curve 1 and, therefore, better wettability than curve 2 Hydrol. Process. 21, 2266 2275 (2007) DOI: 10.1002/hyp

2272

J. V. BAYER AND G. E. SCHAUMANN

and curve 2 for the water repellent sample. Both curves have a comparable shape, but, the gradient of curve 2 is steeper. The points I to IV represent certain states in the development of wettability and are valid for all curves. A supposed schematic representation of the interactions between water and the surface of a soil particle at these points is given in Figure 5(b). Point I represents the state of the samples with high water contents, where all samples would show complete wettability. A water lm of approximately 4 complete water layers is present on the particle surface, thus, leading to good wettability, as the adsorbed water shows properties of free water and controls the surface properties (Clifford, 1975; Zettlemoyer et al., 1975; Vogler, 1998). The difference in surface wettability is determined by the number of outward orientated hydrophilic functional groups, as suggested by e.g. Mashum et al. (1988), Roy and McGill (2000). Upon drying, the water content decreases and the water repellency increases until reaching its maximum at point II. Here, a small number of water molecules in strongly structured order are adsorbed on the particle surface, reducing the surface free energy and, hence, increasing the water repellency (Derjaguin and Churaev, 1986; Vogler, 1998). With further decrease in the water content, the water lm breaks up and free adsorption sites for water are available (Clifford, 1975), so that the wettability increases (point III). The nal point IV of the curve is the point, at which the water content reaches about 0% and soil samples show complete or at least improved wettability, which is then solely determined by the orientation of the polar groups. The energetic state is probably high for outward orientated hydrophilic molecules in a quasi water free environment. To improve the energetic state, hydrophilic molecules may tend to orientate inwards or bend together, and, thus, leading to a lower density of available hydrophilic functional groups, which would increase water repellency. So, over the time of drying, a sample following originally curve 1 in the development of wettability may approach curve 2 in point IV (Figure 5(a)). If an initially water repellent sample is forced in contact with water over a long period, hydrophilic functional groups will orientate outwards (e.g. Tschapek, 1984, Mashum and Farmer, 1985) and so the wetting behaviour of the sample would approach curve 1 in point I. Conformational changes are, however, not necessarily restricted to low water contents. A change in conformation of SOM possibly takes place when the orientation of functional groups does not correspond to the equilibrium orientation, which may be a function of temperature and water cover. Examples of transformations between the curves are represented by the arrows in Figure 5(a) with a preferential direction given by the arrowhead. The above described mechanisms may also be relevant at eld conditions as some water repellent samples have very low eld water contents and top layer soils are exposed to high temperatures during summer months and can well reach temperatures around 30 C (Scheffer and Schachtschabel, 2002). Conformational changes may be
Copyright 2007 John Wiley & Sons, Ltd.

slower in the eld state than with drying but are probable to take place within long term periods. Such conformational changes may explain the initially unexpected development of wettability of water repellent samples upon drying at 35 C: at the starting point II, due to the loss of water and, therefore, increasing number of sorption sites for water, the water repellency rst decreases. As the energetic state at this point is unfavourable and the drying temperature too low to remove the water layer completely, the inward orientation of hydrophilic groups initiates already at this point, leading again to increased repellency. For the initially wettable samples, water repellency does not develop at low water contents upon drying at 35 C, because probably more hydrophilic functional groups are orientated outwards than for the initially water repellent sample. Drying would remove the water lm and free water sorption sites were available. At this stage, as for the water repellent sample, an inward orientation of the hydrophilic functional groups may be initiated. However, due to the initial state of more available functional groups, no or only slow development of water repellency would take place within the observed time frame. Regarding the development of the wettability of the wettable samples upon drying at 65 C, it can be assumed that the state of maximum water repellency does not appear in the WDPTwater content curves, as the drying process is too fast and the energy input from the drying probably is high enough to start the conformational changes at an early state. The same mechanism can be assumed for all samples when drying at 105 C as the energy input is even higher. This may lead to more rapid conformational changes, competing with the effect of decreasing water content. The conformation of functional groups depicted in curve 2 in Figure 5(a) does not represent the endpoint of conformational changes, so that a further inward orientation of hydrophilic functional groups over time may be possible. This means that drying of samples may lead towards a similar wetting behaviour of initially water repellent and wettable samples, as the conformations of SOM functional groups on the surface approach each other. Effect of the pH on wettability Figure 6 shows the WDPT as a function of the pH between pH D 15 and pH D 11. The initial pH of the samples (Bh, Bw, Th, Tw) was between pH 40 and 46 (details see Table I). The pH of all control samples did not change after the addition of pure water, but the WDPT of the water repellent samples corresponding to these pH decreased strongly. Therefore, the initial WDPT of the samples Bh and Th are identied additionally in Figure 6. As the development of WDPT upon addition of HNO3 and NaOH requires separate discussion, these two aspects are described and discussed separately. The WDPT of all samples from both locations developed in a comparable manner after the addition of HNO3 :
Hydrol. Process. 21, 2266 2275 (2007) DOI: 10.1002/hyp

DRYING AND PH EFFECTS ON SOIL WATER REPELLENCY

2273

Figure 6. WDPT of sample pairs from Tiergarten and Buch as a function of articially changed pH. pH was changed by adding different amounts of HNO3 and NaOH. Initial eld WDPT for the samples Th and Bh are additionally indicated, WDPT of the control samples of Th and Bh were both 0s

only below a pH of 3 did all samples develop severe water repellency, with WDPT up to 300 min. It is therefore speculated that the development of WDPT in this region is attributed to comparable mechanisms. We suggest that the increase of water repellency is caused by chemical reactions such as acid catalysed condensation reactions and oxidation reactions of the organic substances, because HNO3 in high concentrations is a good oxidant (E0 D C096 V). As such low pH values are rarely reached under natural conditions, these reactions are considered of low relevance for the development of natural wettability in the eld. The high wettability between pH 3 and 5 for Buch and Tiergarten may be attributed to the superimposed effect of the liquid addition itself, as found for the addition of pure water to the control samples. As shown in Figure 6, the WDPT of the water repellent and wettable sample from Buch did not increase with rising pH (addition of NaOH): all samples were completely wettable above the initial pH. In contrast, the WDPT of water repellent and wettable samples from Tiergarten rst increased with rising pH, considering the WDPT of both the control samples as well as the initial WDPT. Between pH 45 and 65, all Tiergarten samples show a distinct maximum in WDPT with values between 400 and 550 min. At pH above 10, all Tiergarten samples were completely wettable (WDPT <1 s). In general, articial pH changes within the range of natural soil pH affected only the WDPT of the Tiergarten samples. This is in accordance with Hurra and Schaumann (2006), who stated that pH may be relevant only for the wettability of Tiergarten, as they found a relation between low eld pH and water repellency for this location, whereas for samples from Buch no such correlation was found. Considering the differences between water repellent and wettable samples upon pH changes also conrms the expectation, that bringing samples to the same pH should adjust their wettability to each other as the structure of SOM may be altered in a similar manner. But the qualitative shape of the WDPTpH curve of the Tiergarten samples does not exactly correspond
Copyright 2007 John Wiley & Sons, Ltd.

to the expectations expressed by Steenhuis et al. (2001) or Hurra and Schaumann (2006), who suggested an increase in WDPT with decreasing pH, and a decrease in WDPT with increasing pH. In contrast, in our study we found a maximum water repellency at pH above the eld pH (pH 4565). This behaviour may be explained by assuming that the conformation of SOM in all samples is altered with the addition of small amounts of NaOH. For macromolecules with hydrophobic side chains, the existence of hydrophobic shielded intramolecular nanodomains at low pH was observed (Deo et al., 2005) and similar mechanisms were suggested by e.g. Mashum and Farmer (1985) and Valat et al. (1991) for water repellent soils. With increasing pH, the carboxylic groups dissociate, and therefore increase in charge. This leads to more stretched molecules and an increasing outward exposure of hydrophobic domains (Deo et al., 2005). Assuming further that molecules within SOM behave in comparable manner, the addition of small amounts of NaOH would induce a conversion of soil organic molecules to a more stretched form, in which more hydrophobic chains are orientated outwards. This would explain the observed increase of the water repellency with the rst addition of NaOH. The complete wettability for pH >10 may be explained with the complete deprotonation of functional groups in SOM and, therefore, higher polarity (Terashima et al., 2004; Hurra and Schaumann, 2006), or even a dissolution of organic matter (Kononova and Belchikova, 1961). The inuence of pH on the wettability of the Tiergarten samples is, thus, most probably based on the interplay of two partly competing processes: rstly, the disruption of SOM aggregates with hydrophobic nanodomains, which are stabilized by the interaction of hydrophilic functional groups, leads to an increase in water repellency with increasing the pH. Secondly, the deprotonation and, therefore, increasing polarity with increasing pH improves wettability. For samples from Buch, the process of SOM disruption may either be overcompensated by the addition of liquid, or the deprotonation of functional groups is faster than in Tiergarten, so that no water repellency could be detected for the samples upon pH changes. These results would additionally indicate that the amount of pH active functional groups in SOM is smaller in Buch samples than in Tiergarten samples. This assumption, however, requires further evidence from structural investigations. For both sample sites, the inuence of sesquioxides on the surface charge of mineral and organomineral matter may be of additional relevance, especially in combination with changes in SOM.

CONCLUSIONS The results of this study conrmed that the development of water repellency is not exclusively based on the water content, as already suggested in a number of studies (e.g. Doerr et al., 2000; Hurra and Schaumann, 2006). It is suggested that the repellency characteristic of
Hydrol. Process. 21, 2266 2275 (2007) DOI: 10.1002/hyp

2274

J. V. BAYER AND G. E. SCHAUMANN

a soil sample is partly controlled by a water lm of variable thickness. The thickness of the water lm is probably mainly controlled by the water content. However, the change in wettability after reaching constant water content implies that additionally slow changes in surface properties are responsible for the wetting behaviour. These surface properties are most probably controlled by the orientation of polar functional groups in SOM, which may change with time at temperature-dependent rates. We suggest that the effects of water content and conformational changes of SOM are competing and partly antagonistic at low water contents, as the direction of wettability development at low water contents depends on the temperature. This temperature dependence further leads us to assume that the rate of the conformational changes in SOM is controlled by the temperature: higher drying temperatures lead to faster changes in the conformation, and low drying temperatures only induce slow conformational changes, depending, therefore, even more strongly on the drying time. The suggested model may be relevant also at eld conditions as some water repellent samples have very low initial water contents at which conformational changes may take place. At these low water contents the water lm present on the surface may not have liquid-like properties any more, and continued dry periods during the summer may even lead to a disruption within this water lm. This could induce partly irreversible conformational changes or even condensation reactions like esterication. The disruption of the water lm may further allow organic contaminants to sorb in a rather irreversible way directly to the unprotected surface (Todoruk et al., 2003) and hide below the water lm after remoistening. The conformational changes may additionally be governed by pH. The WDPT of water repellent and wettable samples from each sample site react identically to changes in pH, suggesting, therefore, that the conformation in SOM is altered by the addition of acidic or basic solutions and with that becoming similar in both water repellent and wettable samples. Further, it can be assumed that a deprotonation of carboxylic and phenolic functional groups is induced by base addition. The rst deprotonation of functional groups may lead to a repulsion of charged groups and, hence, to more stretched molecules that enhance water repellency. Further increasing deprotonation leads to increasing negative charge and with that overcompensate this rst effect and lead to increased wettability. Additionally, dissolution of SOM molecules may occur, leading eventually to clean spots on the mineral surface, leading to further increased wettability. The results of pH changes also conrmed that specic sample location dependent factors have to be considered for the development of water repellency. Inuencing factors may be the composition of SOM itself or the number of available pH active functional groups. Additionally, the differences between laboratory and eld conditions regarding the inuence of pH on the wettability have to be claried. The water lm on soil particle surfaces which is supposed to be responsible for the wetting behaviour of
Copyright 2007 John Wiley & Sons, Ltd.

the samples, is currently being examined in our group by the means of 1 H-MAS-NMR and Differential Scanning Calorimetry (DSC). The results from this study give insight into the relevance of temperature, moisture conditions and therefore climate for the expression of soil water repellency. They further indicate that the mechanisms leading to water repellency are partly controlled by the location itself. Therefore, a deeper understanding of these mechanisms is required for each location of interest prior to development of remediation measures.

ACKNOWLEDGEMENTS

The experiments were performed at the Technical University Berlin, Dept. Environmental Chemistry. The authors want to thank Prof. Rotard for the use of the Environmental Chemistry laboratories at TU Berlin, where both authors were employed until recently. We would like to thank the whole HUMUS group for the valuable discussions in the course of this study. We also want to thank Dr. Ruth Ellerbrock und Ralf Rath from the ZALF M ncheberg for the use of the IR spectrometer, Annett u Hapka who assisted us with the literature research and all members of the collaborate research group INTERURBAN (DFG research group INTERURBAN, HUMUS, SCHA 849/4-2).

REFERENCES Babejova N. 2001. An inuence of changing the humic acids content on soil water repellency and saturated hydraulic conductivity. Journal of Hydrology and Hydromechanics 49(5): 291 300. Blackwell P. 1997. Water Repellent Soils: Managing Water Repellent Soils, Farmnote 109/96. Western Australian Department of Agriculture; 2. Capriel P, Beck T, Borchert H, Gronholz J, Zachmann G. 1995. Hydrophobicity of the organic matter in arable soils. Soil Biology and Biochemistry 27(11): 1453 1458, DOI: 101016/0038-0717(95)00068. Chen Y, Schnitzer M. 1978. The surface tension of aqueous solutions of soil humic substances. Soil Science 125(1): 715. Clifford J. 1975. Properties of water in capillaries and thin lms. In Water in Disperse Systems, Vol. 5, Franks F (ed). Plenum Press: New York; 75131. Dekker LW, Doerr SH, Oostindie K, Ziogas AK, Ritsema CJ. 2001. Water repellency and critical soil water content in a dune sand. Soil Science Society of America Journal 65(6): 1667 1674. Deo P, Deo N, Somasundaran P, Jockusch S, Turro NJ. 2005. Conformational changes of pyrene -labeled polyelectrolytes with pH: effect of hydrophobic modications. Journal of Physical Chemistry B 109: 20714 20718, DOI: 101021/jp051743f. Derjaguin BV, Churaev NV. 1986. Properties of water layers adjacent to interfaces. In Fluid Interfacial Phenomena, Croxton CAe (ed). John Wiley & Sons: New York; 663 738. Diehl D, Schaumann GE. 2007. The nature of wetting on urban soil samples: wetting kinetics and evaporation assessed from sessile drop shape. Hydrological Processes 21(17): 2255 2265. Doerr SH, Thomas AD. 2000. The role of soil moisture in controlling water repellency: new evidence from forest soils in Portugal. Journal of Hydrology 231 232: 134147, DOI: 10101016/S00221694(00)00190-6. Doerr SH, Shakesby SH, Walsh RPD. 2000. Soil water repellency: its causes, characteristics and hydro-geomorphological signicance. Earth-Science Reviews 51(1 4): 3365, DOI:101016/S00128252(00)00011-8. Doerr SH, Shakesby RA, Dekker LW, Ritsema CJ. 2006. Occurrence, prediction and hydrological effects of water repellency Hydrol. Process. 21, 2266 2275 (2007) DOI: 10.1002/hyp

DRYING AND PH EFFECTS ON SOIL WATER REPELLENCY amongst major soil and land-use types in a humid temperate climate. European Journal of Soil Science 57(5): 741754, DOI:101111/j.1365 2389200600818.x. Ellerbrock RH, Gerke HH, Bachmann J, Goebel M-O. 2005. Composition of organic matter fractions for explaining wettability of three forest soils. Soil Science Society of America Journal 69(1): 5766. Franco CMM, Clarke PJ, Tate ME, Oades JM. 2000. Hydrophobic properties and chemical characterisation of natural water repellent materials in Australian sands. Journal of Hydrology 231 232: 47 58, DOI: 10101016/S0022-1694(00)00182-7. Goebel M-O, Bachmann J, Woche SK, Fischer WR, Horton R. 2004. Water potential and aggregate size effects on contact angle and surface energy. Soil Science Society of America Journal 68(2): 383 393. Gruenberg B, Emmler T, Gedat E, Shenderovich I, Findenegg GH, Limbach HH, Buntkowsky G. 2004. Hydrogen bonding of water conned in mesoporous silica MCM-41 and SBA-15 studied by 1H solid-state NMR. Chemistry-A European Journal 10(22): 5689 5696, DOI: 101002/chem.200400351. Horne DJ, McIntosh JC. 2000. Hydrophobic compounds in sands in New Zealandextraction, characterisation and proposed mechanisms for repellency expression. Journal of Hydrology 231232: 35 46, DOI: 101016/S0022-1694(00)00181-5. Hurra J, Schaumann GE. 2006. Properties of soil organic matter and aqueous extracts of actually water repellent and wettable soil samples. Geoderma 132(1 2): 222239, DOI: 101016/j.geoderma. 200505012. de Jonge LW, Jacobsen OH, Moldrup P. 1999. Soil water repellency: effects of water content, temperature, and particle size. Soil Science Society of America Journal 63(3): 437442. Karnok KA, Rowland EJ, Tan KH. 1993. High pH treatments and the alleviation of soil hydrophobicity on golf greens. Agronomy Journal 85(5): 983 986. Keizer JJ, Doerr SH, Malvar MC, Ferreira AJD, Pereira VMFG. 2007. Temporal and spatial variations in topsoil water repellency throughout a crop-rotation cycle on sandy soil in north-central Portugal. Hydrological Processes 21(17): 2317 2324. King PM. 1981. Comparison of methods for measuring severity of water repellence of sandy soils and assessment of some factors that affect its measurement. Australian Journal of Soil Research 19: 275285. Kononova MM, Belchikova NP. 1961. Rapid methods of determination the humus composition of mineral soils. Pochvovedenie 10: 75. Leighton-Boyce G, Doerr SH, Shakesby RA, Walsh RPD, Ferreira AJD, Boulet AK, Coelho COA. 2005. Temporal dynamics of water repellency and soil moisture in eucalypt plantations, Portugal. Australian Journal of Soil Research 43(3): 269280, DOI: 101071/SR04082. Letey J. 1969. Measurement of contact angle, water drop penetration time, and critical surface tension. Proceedings of the Symposium on Water -Repellent Soils, May 1968. University of California; 4347. Mashum M, Farmer VC. 1985. Origin and assessment of water repellency of a sandy South Australian soil. Australian Journal of Soil Research 23: 623 626. Mashum M, Tate ME, Jones GP, Oades JM. 1988. Extraction and characterization of water-repellent materials from Australian soils. Journal of Soil Science 39: 99110. Quyum A, Achari G, Goodman RH. 2002. Effect of wetting and drying and dilution on moisture migration through oil contaminated

2275

hydrophobic soils. The Science of The Total Environment 296(1 3): 77 87, DOI:101016/S0048-9697(02)00046-3. Reemtsma T, Savric I, Jekel M. 2003. A potential link between the turnover of soil organic matter and the release of aged organic contaminants. Environmental Toxicology and Chemistry 22(4): 760766, DOI: 101897/1551-5028(2003)022<0760:APLBTT>20.CO;2. Ritsema CJ, Dekker LW. 1994. Soil moisture and dry bulk density patterns in bare dune sands. Journal of Hydrology 154(1 4): 107131, DOI:101016/0022-1694(94)90214-3. Roberts FJ, Carbon BA. 1971. Water repellence in sandy soils of southwestern Australia. 1. Some studies related to eld occurence. Field Station Record Division Plant Industry CSIRO (Australia) 10: 13 20. Roberts FJ, Carbon BA. 1972. Water repellence in sandy soils of southwestern Australia. II. Some chemical characteristics of hydrophobic skins. Australian Journal of Soil Research 10(1): 35 42. Roy JL, McGill WB. 2000. Flexible Conformation in organic matter coatings: an hypothesis about soil water repellency. Canadian Journal of Soil Science 80: 143152. Scheffer F, Schachtschabel P. 2002. Lehrbuch der Bodenkunde, Vol. 15. Spektrum Akademischer Verlag GmbH: Berlin, Heidelberg; 1593. Schreiber A, Ketelsen I, Findenegg GH. 2001. Melting and freezing of water in ordered mesoporous silica materials. Physical Chemistry Chemical Physics 3: 1185 1195, DOI: 101039/b010086m. Steenhuis TS, Rivera JC, Hern ndez CJM, Walter MT, Bryant RB, a Nektarios P. 2001. Water Repellency in New York State Soils. International Turfgrass Society Research Journal 9: 624 628. T umer K, Stoffregen H, Wessolek G. 2005. Determination of repellency a distribution using soil organic matter and water content. Geoderma 125(1 2): 107 115, DOI:101016/j.geoderma.200407004. Terashima M, Fukushima M, Tanaka S. 2004. Inuence of pH on the surface activity of humic acid: micelle-like aggregate formation and interfacial adsorption. Colloids and Surfaces A: Physicochemical and Engineering Aspects 247: 7783, DOI:101016/j.colsurfa.200408028. Todoruk TR, Litvina M, Kantzas A, Langford CH. 2003. Low-eld NMR relaxometry: a study of interactions of water with water-repellent soils. Environmental Science and Technology 37: 2878 2882, DOI: 101021/es026295t10. Tschapek M. 1984. Criteria for determining the hydrophilicity hydrophobicity of soils. Journal of Soil Science and Plant Nutrition 147(2): 137 149. Valat B, Jouany C, Riviere LM. 1991. Characterization of the wetting properties of air-dried peats and composts. Soil Science 152: 100107. Vogler EA. 1998. Structure and reactivity of water at biomaterial surfaces. Advances in Colloid and Interface Science 74: 69117, DOI:101016/S0001-8686(97)00040-7. Wallis MG, Horne DJ. 1992. Soil water repellency. Advances in Soil Science 20: 91146. Zettlemoyer AC, Micale FJ, Klier K. 1975. Adsorption of Water on WellCharacterized Solid Surfaces. In Water in Disperse Systems, Chapter 5, Franks F (ed). Plenum Press: New York; 249 289. Ziogas AK, Dekker LW, Oostindie K, Ritsema CJ. 2003. Soil water repellency in northeastern Greece. Soil Water Repellency: Occurence, Consequences and Amelioration, Vol. 1, Ritsema CJ, Dekker LW (eds). Elsevier: Amsterdam; 127 135.

Copyright 2007 John Wiley & Sons, Ltd.

Hydrol. Process. 21, 2266 2275 (2007) DOI: 10.1002/hyp

Você também pode gostar