Anda di halaman 1dari 7

Environmental Pollution 175 (2013) 110e116

Contents lists available at SciVerse ScienceDirect

Environmental Pollution
journal homepage: www.elsevier.com/locate/envpol

Isosteric heats of sorption and desorption of phenanthrene in soils and


carbonaceous materials
Guohui Wang 1, Peter Grathwohl*
Center for Applied Geoscience, University Tübingen, Hölderlinstr. 12, D-72074 Tübingen, Germany

a r t i c l e i n f o a b s t r a c t

Article history: Isosteric heats (DH) of sorption/desorption of phenanthrene were determined for carbonaceous materials
Received 26 June 2012 (Pahokee peat, lignite, and high-volatile bituminous coal) and two soils based on reported equilibrium
Received in revised form sorption/desorption isotherms at four different temperatures (4, 20, 46 and 77  C). In addition, DH for
20 December 2012
desorption of native phenanthrene was determined to elucidate the “aging” effect by equilibrating
Accepted 21 December 2012
samples with water at six temperatures (20, 40, 53, 61, 73, and 86  C). Isosteric heats decreased with
increasing solute concentration and were in a range of 19e35 kJ mol1. Values higher than the heat of
Keywords:
octanolewater phase transfer for phenanthrene (19 kJ mol1) imply that both partitioning and
Isosteric heat
Sorption/desorption
adsorption processes are involved for these materials, where the sorptive contributions from both
Phenanthrene processes were estimated based on the phenanthrene thermodynamic data. Moreover, on the basis of DH
Carbonaceous material values of desorption, release of native and spiked phenanthrene from our samples was similar.
Ó 2013 Elsevier Ltd. All rights reserved.

1. Introduction isosteric heats (DH) transferred in the sorption/desorption pro-


cesses, contribute predominantly to the sorption. It will be valuable
Sorption and desorption are major processes influencing the if the sorption contributions resulting from partitioning and
fate, transport and bioavailability of hydrophobic organic com- adsorption could be distinguished through isosteric heat mea-
pounds (HOC) in the environment. In the literature, organic con- surements. Theoretically partitioning and adsorption processes
taminants partitioning into “soft” organic matter and adsorption on involve different sorbate and sorbent interactions, thus different
“hard” organic sorbent surface are well accepted as the two main DH values could be expected.
sorption mechanisms in the natural environment (Weber et al., The isosteric heats of sorption for various organic chemical
1992; Xing and Pignatello, 1997; Xia and Ball, 1999). However, compounds on different types of natural sorbents, including in-
a quantitative separation of sorption contributions resulting from teractions on mineral surfaces and polymers in vapour phase and
these two sorption mechanisms is lacking in the literature. Tradi- aqueous systems, have been reported previously (Adams et al.,
tionally, sorption has been classified into three categories according 2007; Arp et al., 2008; Chiou et al., 1979, 1985; Costanza and
to the attractive forces governing the molecular interactions be- Pennell, 2007; Farrell et al., 1999; Haftka et al., 2010; Kleineidam
tween solute and sorbent, namely, physical, chemical and electro- et al., 2004; Krauss and Wilcke, 2001; Lüres and Ten Hulscher,
static. Thermodynamically, at equilibrium the contaminant 1996; Mader et al., 1997; Madlener et al., 2003; Pitta et al., 1996;
distribution between the solid and aqueous phases is ultimately Pre et al., 2002; Ten Hulscher and Cornelissen, 1996; Tremblay et al.,
governed by the isosteric heats (also called enthalpy, DH) and en- 2005; Werth and Reinhard, 1997; Woodburn et al., 1989). Table 1
tropies of sorption or desorption. The magnitude and sign of DH compiles literature values of DH for PAHs as well as other hydro-
values gives valuable insight to the molecular interactions between phobic compounds to different sorbents. Most of these studies
sorbate and sorbent. For sparingly soluble HOC molecules, the focused on the temperature dependence of sorption without con-
London dispersion or Van der Waals forces between solutes and sidering separation of the heat contributions resulting from either
sorbents, which can be determined from measurements of the partitioning or adsorption processes. Sorption mechanisms can be
distinguished by varied DH values in the data compiled in Table 1, in
which the isosteric heats on pure substances or surfaces could be
* Corresponding author.
E-mail address: grathwohl@uni-tuebingen.de (P. Grathwohl).
used as a reference in separating different sorption mechanisms.
1
Present address: Pacific Northwest National Laboratory, P.O. Box 999, MS P7-54, Relatively low values of 13 to 23 kJ mol1 are reported for
Richland, WA 99352, USA. phenanthrene on hydrophilic mineral surfaces (such as Al2O3 or

0269-7491/$ e see front matter Ó 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.envpol.2012.12.027
G. Wang, P. Grathwohl / Environmental Pollution 175 (2013) 110e116 111

Table 1
List of reported DH of PAHs as well as other hydrophobic compounds in different sorbents (1 standard deviation).

Sorbents Compounds DH (kJ mol1) Temperature ( C) Reference


Mineral oxides (a-Al2O3/a-Fe2O3) Naphthalene 12.5 (3.6)/26 (10) 5e35 Mader et al., 1997
Phenanthrene 18.2 (1.0)
Anthracene 16 (2.8)/17.2 (1.9)
Fluoranthene 19.5 (0.4)/16.7 (3.0)
Pyrene 20.8 (0.4)/20 (0.9)
Silica gel (100e200 mm)/Al2O3 Phenanthrene 16.1/16.8 25e45 Su et al., 2006
(w150 mesh) Pyrene 23.2/18.8
Silica gels (150/40) Phenanthrene 12.6 to 17.9/19.6 to 22.8a 5e45
Graphite surface Phenanthrene 47.1a 5e45
Silty loam soil Phenanthrene 19 (2.2) 14e35 Woodburn et al., 1989
Anthracene 22.7 (0.6)
Fluoranthene 24.5 (0.8)
Pyrene 37.7 (10.7)
Aquifer sediment containing low Naphthalene 11/1.1 4e26 Pitta et al., 1996
organic carbon (batch/column) Phenanthrene 3.3/5.6
Pyrene 14/14
Soils (Bayreuth, Germany) PAHs and PCBs 0 to 20 20e80 Krauss and Wilcke, 2001
Demolition waste and harbour sediments PAHs 32 to 58 25e100 Madlener et al., 2003
Aquifer sediments and rock fragments Phenanthrene 23 to 36 20e40 Kleineidam et al., 2004
Suspended particles Phenanthrene 4.5 to 3.2 2e20 Tremblay et al., 2005
Dissolved organic carbon (DOC) Fluoranthene 18.3 16e45 Lüres and Ten Hulscher, 1996
Benzo[ghi]perylene 40.7
Polyethylene Phenanthrene 18 (1) 16e30 Adams et al., 2007
Pyrene 29 (2)
Polydimethylsiloxane Phenanthrene 16.2 (0.48)
Anthracene 15.3 (1.2)
Fluorene 14.7 (0.84)
Fluoranthene 18.9 (1.7)
Pyrene 22.1 (0.25) 3e36 Haftka et al., 2010
Low density polyethylene Acenaphthene 26
Phenanthrene 22
Fluorene 43
pyrene 46 2e30 Lohmann, 2012
Natural sorbents Hydrophobic pollutants 13 to þ3.8 3e37 Ten Hulscher and
Cornelissen, 1996
Granular activated carbon VOCs 40 to 80 20 Pre et al., 2002
Silica Gel TCE 9.5 to 45 15e60 Werth and Reinhard, 1997
Natural solids 0 to 34
Silica gel TCE 22.2 (1.5) to þ20.8 (1.7) 5e90 Farrell et al., 1999
Zeolite type NaX þ20.3 (12.1)
Silty clay soil PCE 12 (0.3) 25e95 Costanza and Pennell, 2007
a
DH values for phenanthrene of 12.6 to 17.9 kJ mol1 for silica gel 150, 19.6 to 22.8 kJ mol1 for silica gel 40, and 47.1 kJ mol1 for graphite surface were rein-
terpreted by Su et al. (2006) based on experimental data from Huang and Weber (1997).

silica gel), where London forces are the main cause for adsorption in natural geosorbents through isosteric heat measurements.
(Su et al., 2006). Polymer sorbents, such as low density polyethyl- Meanwhile, sorption/desorption hysteresis is also discussed by
ene (LDPE, often used as a passive samplers), are generally assumed comparing sorption/desorption isosteric heats. Carbonaceous ma-
to comprise amorphous and crystalline regions where both parti- terials (CM) and soils containing carbonaceous materials were
tioning and adsorption may occur (Lohmann, 2012). A relatively selected in this study because of (1) their relative pure substances,
small range of DH values (16 to 22 kJ mol1) has been reported (2) their relevance in strong sorption of hydrophobic organic
for phenanthrene with LDPE. Graphite provides a relative homo- compounds in soils and in sediments, and (3) values of DH for
geneous hydrophobic surface for strong adsorption of organic sorption of PAHs on CM materials are rarely reported. High isosteric
compounds from aqueous solution, with the reported DH value of heat values are expected if strongly sorbing carbonaceous materials
47.1 kJ mol1 typical for solute condensation (Su et al., 2006). In are present (Arp et al., 2008). CM is non-carbonate, carbon-con-
contrast to LDPE and graphite, a wide range of DH values (3.3 to taining matter and the broad category of CM includes residues from
36 kJ mol1 in Table 1) has been observed for soils and sediments. combustion and pyrolysis processes, such as soot and char, as well
The variations observed in aqueous phase DH values among soils as geologic processes (diagenesis and catagenesis), such as coal and
and sediments could originate from two sources. Firstly, different kerogen (Allen-King et al., 2002). In this study, we evaluated a data
geosorbents vary in physico-chemical properties depending on set on sorption/desorption isotherms measured at different tem-
their origin, which again results in different values for DH. Sec- peratures (Wang et al., 2007) to determine DH values of phenan-
ondly, experimental artifacts resulting from differences in exper- threne for CM and CM-containing natural geosorbents. Based on
imental procedures, such as variations in equilibration time periods the heat of solution of phenanthrene in water, the energy contri-
from a few hours to days or weeks, could result in different DH butions from both partitioning and adsorption processes were
values even in the same sorbent-sorbate system (Wang et al., 2007). estimated, which can help to elucidate the sorption/desorption
Non-equilibrium during sorption, for instance, would result in mechanisms. In addition, it should be noted that, besides the
higher DH-values of desorption because the activation energy of determined sorption and desorption isosteric heats for phenan-
diffusion would also be included (Wang and Grathwohl, 2009). threne on selected materials, DH for desorption of the native
The main objective of this study was to distinguish sorption (“aged”) phenanthrene in these samples was also determined over
contributions resulting from partitioning and adsorption processes a similar temperature range of 20e86  C, and compared with those
112 G. Wang, P. Grathwohl / Environmental Pollution 175 (2013) 110e116

from the laboratory spiked samples. To our knowledge, no previous Table 2


studies have reported isosteric heat of desorption for native con- Physico-chemical properties of phenanthrene.

taminants from such relatively pure carbonaceous materials. Melting point ( C) (Tm) 99.5a
Boiling point ( C) 340.2a
Heat of fusion [kJ mol1] (DHfus) 18.6b
2. Data evaluation Heat of solution in water [kJ mol1] (DHsol) (solid)c 36.7e39.1
Heat of solution in water [kJ mol1] (subcooled liquid)d 18.1e20.5
2.1. Sorption/desorption isotherms on spiked samples a
Melting point and boiling point are from Verschueren (1983).
b
Data from Chiou (2002, page 71).
In this study, we are analyzing sorption/desorption isosteric c
Reported by May et al. (1983) based on a series of water solubilities (Ssolid)
heats by interpreting the sorption/desorption isotherms reported measured between 4e29.9  C (May et al., 1983), 25e73.4  C (Wauchope and Getzen,
in Wang et al. (2007). Sample selection, characterization, and iso- 1972), and 8e30  C (Schwarz, 1977); heat of condensation (DHcond) is the reverse
heat of solution with DHcond ¼ DHsol.
therm experiments were described in details in Wang et al. (2007). d
Calculated as the difference between DHsol and DHfus.
Briefly, three carbonaceous samples (Pahokee peat, lignite, and
high-volatile bituminous coal (HC)) and two different soil samples
were measured at increasing temperatures (20, 40, 53, 61, 73, and
(anthropogenic soil (AS5) and mineral soil (MS4)) were selected to
86  C) using an automated accelerated solvent extraction (ASE)
cover a wide range of natural geosorbents. Phenanthrene (98%,
device (Dionex ASE 300). The time for each water equilibration step
Aldrich Chemical Corp.) was used as the probe compound in their
was 99 min, which was tested to be sufficient for equilibration of
study as a representative PAH. By using a mass-balance-monitored
these pulverized samples at low water to solid ratios as confirmed
new batch sorption protocol, sorption and desorption isotherms of
by preliminary experiments. Note that equilibration at low water to
phenanthrene on above selected materials at four temperatures (4,
solid ratios is much faster than that at large water to solid ratios,
20, 46 and 77  C) were determined. All the sorption/desorption
because the latter result in a more exhaustive desorption of the
isotherms in their study were nonlinear (with 1/n of 0.61e0.86)
sorbate (Grathwohl, 2011). Also, comparison of aqueous concen-
except AS5, which showed a relatively slight nonlinearity (1/
trations obtained after 30 min and 99 min equilibration showed no
n ¼ 0.91). Sorption nonlinearity indicates that physical adsorption
significant differences at a given temperature in several pre-
played an important role in the soluteesorbent interactions
liminary experiments, indicating that 99 min was a conservative
(especially pronounced in coals). In all samples, the sorption ca-
duration for equilibration. This is consistent with Lou et al. (1997)
pacity decreased with increasing temperature as expected for
and Wennrich et al. (2000), where 20e30 min was used for
exothermic sorption and endothermic desorption processes. Fig. 1
optimal extraction of hydrophobic organic compounds with
shows lignite as an example for decreasing Kd values with
solvent or water using ASE.
increasing temperatures over an aqueous concentration range
spanning 0.01%e50% of the phenanthrene solubility. Detailed
isotherm parameters at each temperature are listed in Wang et al. 2.3. Thermodynamic parameters of phenanthrene
(2007).
Thermodynamic parameters of phenanthrene relevant for this
study are listed in Table 2. Emphasis here is on the heats of solution
2.2. Desorption of native phenanthrene at elevated temperatures
of phenanthrene in water (DHsol) based on measured water solu-
bilities (Ssolid) between 4  C and 73.4  C (May et al., 1983; Schwarz,
All samples used in Wang et al. (2007) contain trace concen-
1977; Wauchope and Getzen, 1972). The heat of solution for solid
trations of native phenanthrene as listed in Table 3. As described in
phenantherene (DHsol) includes the heat of fusion (DHfus) which
Wang et al. (2007), similar to the classical batch system, the equi-
indicates the heats involved during phenanthrene transfer between
librium concentrations of desorbed native phenanthrene in water
solid to liquid phases. The difference between DHsol and DHfus is the
heats of solution of the subcooled liquid states of phenanthrene,
which represents the heats involved in the dissolution of liquid
200000 0.5S phenanthrene in water.
0.1S
0.01S 2.4. Data evaluation
150000 0.001S
The molar isosteric heat of sorption (DH) is a general expression,
Kd [L/Kg]

0.0001S
which represents the enthalpy change involved in the transfer of
100000 a solute from the dissolved to the sorbed state while holding the
solid-phase concentration (Cs) constant. It indicates the difference
in binding energies between the sorbent and the sorbate and be-
50000 tween the solvent (water) and the solute (sorbate). DH can be
calculated using the ClausiuseClapeyron equation at a given Cs
(Chiou, 2002):

0  n
dlog Cw Cs
77°C 46°C 20°C 4°C DHa=d ¼ 2:303R with Cw ¼ (1)
dð1=TÞ KFr
Temperature
where DHa/d [kJ mol1] is the molar isosteric heat of sorption or
Fig. 1. Lignite is shown as an example of decreasing sorption capacity (Kd) with desorption (DHa and DHd have opposite signs). R is the gas constant
increasing temperature at 5 selected aqueous concentration levels (0.0001, 0.001, 0.01, and T is the temperature in Kelvin. Cs [mg kg1] and Cw [mg L1]
0.1, and 0.5 of the aqueous solubility, S). The Kd was calculated based on the measured
sorption isotherms covering the same concentration range. The solid aqueous solu-
denote the concentrations in the solid and the aqueous phase,
bility of phenanthrene is 0.361, 0.788, 3.35, and 17.37 mg L1 at temperatures of 4, 20, respectively. KFr [mg kg1:(mg L1)1/n] and 1/n [e] are the
46, and 77  C, respectively (Wang et al., 2007). Freundlich coefficients. The change in Cw at a given Cs level was
G. Wang, P. Grathwohl / Environmental Pollution 175 (2013) 110e116 113

Table 3
Isosteric heats of sorption and desorption (DHa, DHd) for five geosorbents calculated from sorption/desorption isotherms at five spiked concentration levels (Cs) over 4e77  C
and for the release of native phenanthrene over 20e86  C (1 standard deviation). Fadsorption indicates the contribution of adsorption to the total sorption/desorption isosteric
heat.

Pahokee peat Cs level [mg kg1] 1.38a 333 833 2083 4167 8333
DHa [kJ mol1] 26.0  0.5 25.0  1.1 23.9  1.7 23.1  2.2 22.3  2.6
R2 [e] 1.0 1.0 0.99 0.99 0.99
DHd [kJ mol1] 35.6  2.4 27.1  1.9 25.9  2.5 24.6  3.0 23.6  3.4 22.7  3.8
R2 [e] 0.98 0.99 0.99 0.99 0.98 0.97
Fadsorption 59% 25% 21% 18% 14% 11%
Lignite Cs level [mg kg1] 0.49a 256 769 2308 6923 20513
DHa [kJ mol1] 30.6  1.9 27.8  1.9 24.9  1.9 22.1  2.0 19.3  2.1
R2 [e] 0.99 0.99 0.99 0.98 0.98
DHd [kJ mol1] 40.5  2.1 31.2  1.6 28.2  1.6 25.1  1.6 22.0  1.7 19.0  1.7
R2 [e] 0.99 0.99 0.99 0.99 0.99 0.98
Fadsorption 77% 43% 32% 21% 11% 1%
HC Cs level [mg kg1] 16.94a 1389 3056 6944 13889 20833
DHa [kJ mol1] 33.5  3.2 29.7  2.6 25.7  2.0 22.4  1.5 20.4  1.2
R2 [e] 0.98 0.99 0.99 0.99 0.99
DHd [kJ mol1] 34.8  4.9 30.2  4.0 25.4  2.1 21.3  2.4 19.0  2.0
R2 [e] 0.96 0.97 0.97 0.98 0.98
Fadsorption 53% 39% 25% 11% 4%
AS5 Cs level [mg kg1] 9.29a 119 476 1905 7143 23810
DHa [kJ mol1] 31.2  2.1 31.9  2.0 32.7  2.1 33.4  2.3 34.1  2.5
R2 [e] 0.99 0.99 0.99 0.99 0.99
DHd [kJ mol1] 41.0  2.3 31.3  1.8 32.1  1.9 32.9  2.1 33.7  2.3 34.4  2.5
R2 [e] 0.98 0.99 0.99 0.99 0.99 0.99
Fadsorption 78% 44% 46% 49% 52% 54%
MS4 Cs level [mg kg1] 4.02a 59 235 882 2941 11765
DHa [kJ mol1] 30.3  0.6 28.8  0.9 27.3  1.3 26.0  1.8 24.4  2.3
R2 [e] 1.0 1.0 1.0 0.99 0.98
DHd [kJ mol1] 31.0  3.4 30.1  1.1 28.2  0.8 26.5  1.0 24.9  1.5 23.0  2.0
R2 [e] 0.97 1.0 1.0 1.0 0.99 0.98
Fadsorption 43% 39% 36% 28% 25% 18%
a
Native phenanthrene loadings; All the Cs loading were normalized by geosorbent organic carbon contents.

calculated based on the equilibrium isotherm parameters (KFr and by the charcoal-like substance, which originated from biomass
1/n) determined at each temperature. In addition to the batch ex- burning and got embedded into the peat.
periments using spiked samples, DHd values of native phenan- For (nonlinear) adsorption of nonionic solutes on porous ad-
threne from the samples using ASE was also determined by using sorbents, the total molar isosteric heat of adsorption for a partially
Eq. (1), where the measured aqueous concentrations after DOC- miscible solute in a water solution is the sum of the heats of con-
correction at each temperature were used instead of isotherm densation (to form a separate condensed phase) and the net mo-
calculated Cw values. The change in native phenanthrene Cs during lecular attraction energy between the condensed solute and
the leaching experiments was found to be insignificant (less than adsorbent (Chiou, 2002). The condensation of sorbed phenan-
2% of the total mass over the temperature range) and thus was threne on the micropore-containing carbonaceous materials (see
neglected. The DH difference between native and spiked samples at Wang et al., 2007) likely occurred by a pore-filling mechanism. For
the same Cs level would indicate different sorbateesorbent a partially water-soluble solute, the basic Polanyi pore-filing theory
interaction forces for native and spiked phenanthrene. expects the solute to condense into the adsorption space as a liquid
or a solid phase, depending on the state of the pure solute at the
3. Results and discussion system temperature (Chiou, 2002). For phenanthrene with a melt-
ing point of 99.5  C, a condensed solid phase (rather than a discrete
Fig. 2 shows the regression of log Cw versus 1/RT for the deter- molecular state) should be formed in the temperature range of 4e
mination of DH for sorption and desorption of phenanthrene on the 77  C. Ordinarily, nonlinear solute adsorption is reflective of either
five sorbents each with five different Cs levels. The linear re- adsorbate condensation to form separate phase or a specific surface
gressions (Eq. (1)) fit the data very well with a coefficient of interaction of the solute with the solid. However, the later effect
determination r2 > 0.96 in all cases. The isosteric heats of sorption/ seems unlikely for non-polar compounds (such as phenanthrene).
desorption with standard deviations determined for our samples For non-polar solutes, a weak linear adsorption occurs usually on
are compiled in Table 3. They range from 19 kJ mol1 to hydrophilic mineral surfaces covered with a water film (that re-
35 kJ mol1, which is equal to or more exothermic than the heat of duces the net attraction between solute and mineral surfaces) to
octanolewater phase transfer for phenanthrene (about prevent solute condensation (Su et al., 2006). Due to net attractions
19 kJ mol1, Lei et al., 2000; representing a partitioning process). between carbonaceous sorbent surfaces and phenanthrene mole-
Yet, the observed range of isosteric heats is less than the heat of cules and the implied phenanthrene condensation (as indicated by
condensation (the reverse heat of solution) of solid phenanthrene, nonlinear isotherms), the isosteric heat of phenanthrene would be
which is about 38 kJ mol1 in the temperature range of interest more exothermic than the heat of condensation. Therefore, in this
(see Table 2). It should be noted that the Pahokee peat used in this study, the total molar isosteric heat of adsorption of phenanthrene
study is known to contain a small amount (w8%) of charcoal-like in solid state is expected to be larger than 38 kJ mol1. Thus,
materials (Karapanagioti et al., 2001), thus a significant sorption the observed isosteric heats (19 to 35 kJ mol1, which are
nonlinearity (1/n ¼ 0.65) and higher heat effect at low loadings are equal to or more exothermic than the heat of octanolewater phase
anticipated for this sorbent because of the adsorptive contributions transfer, but less than the heat of condensation) indicate that both
114 G. Wang, P. Grathwohl / Environmental Pollution 175 (2013) 110e116

10 10 10

1 1 1

0.1 0.1 0.1

C w [mg/L]
C w [mg/L]

C w [mg/ L]
0.01 0.01 0.01

0.001 0.001 0.001

0.0001 0.0001 0.0001

0.00001 0.00001 0.00001


HC MS4
Pahokee peat
0.000001 0.000001 0.000001
0.32 0.34 0.36 0.38 0.4 0.42 0.44 0.32 0.34 0.36 0.38 0.4 0.42 0.44 0.32 0.34 0.36 0.38 0.4 0.42 0.44
1/TR 1/TR 1/TR
10 10

1 1

0.1 0.1
C w [mg/L]
C w [mg/L]

0.01 0.01

0.001 0.001

0.0001 0.0001

0.00001 lignite 0.00001


AS5
0.000001 0.000001
0.32 0.34 0.36 0.38 0.4 0.42 0.44 0.32 0.34 0.36 0.38 0.4 0.42 0.44
1/TR 1/TR

Fig. 2. Van’t Hoff plots for the determination of isosteric heats of sorption and desorption of phenanthrene at five different concentration levels onto lignite, high-volatile bitu-
minous coal (HC), Pahokee peat, and the soils AS5 and MS4. Squares, diamonds, triangles, circles and horizontal bars represent different levels of spiked concentrations (Cs) during
sorption and crosses () denote the corresponding desorption steps. Plus signs (þ) indicate desorption of native phenanthrene for samples where aqueous concentrations were
above the detection limit.

adsorption and partitioning into organic matter are involved in (i.e., the molar exothermic heat) at the lowest concentration (Chiou,
sorption/desorption of phenanthrene in the samples studied. The 2002). From the five sorbents used in this study, the high-volatile
determined DH values thus represent partitioning and adsorption bituminous coal shows the strongest dependency of DHa/d on
processes with different weights. Chiou et al. (1979, 1985) inves- concentration in coincidence with the strongest nonlinearity of
tigated partitioning and adsorption mechanisms and suggested sorption for this sorbent. Here, a maximum DHa/d value of
that they can be separated by comparing the isosteric heat of 55 kJ mol1 is reached if Cs decreases below 10 mg kg1 (i.e., the
sorption with the heat of condensation as well as by monitoring the natural loading of phenanthrene on this sample). These findings
concentration dependency of the isosteric heat on loading levels. If apply in general to the other carbonaceous samples as well as for
we use 47.1 kJ mol1 (adsorption heat of phenanthrene onto the char-containing soil sample MS4. The AS5 sample, however,
a graphite surface, see Su et al., 2006) as the value for the standard shows an opposite trend of a slightly increasing isosteric heat with
adsorption isosteric heat, and 19 kJ mol1 as the value for the increasing concentration levels.
standard partitioning isosteric heat, then the adsorption con- Regressions for the release of native phenanthrene in cases with
tribution (Fadsorption) from the total sorption/desorption isosteric Cw above the detection limit are shown in Fig. 2, and the calculated
heat can be calculated. The Fadsorption data in Table 3 show that large isosteric heats are provided in Table 3 and plotted against Cs in
values of DH are commonly dominated by adsorption, especially for Fig. 3. The desorption isosteric heats of native phenanthrene (DHd)
the desorption of native phenanthrene. The partitioning con- for the soils, Pahokee peat, and lignite samples range between
tribution increases with increased loading, i.e., with increasing 31 kJ mol1 and 41 kJ mol1 and thus are in reasonable agreement
aqueous concentrations Cw. For example, the adsorption con- with the DHa/d values obtained from the spiked experiments (see
tribution almost vanishes at a loading of 8000 mg kg1 in lignite, Table 3 and Fig. 3) at low concentration levels. In the case of high-
which corresponds to Cw of 38% of the phenanthrene aqueous volatile bituminous coal, Cw for native phenanthrene was below the
solubility at 20  C. This is supported by the commonly assumed detection limit at all temperatures.
superposition of sorption mechanisms where adsorption (or pore- In general, the determined DH values on the studied carbona-
filling) dominates at low concentrations while linear partitioning ceous materials and carbonaceous material-containing soils are
takes over at higher concentrations. Together, partitioning and higher than those reported for other natural aquifer sediments and
adsorption results in Freundlich type isotherms (Kleineidam et al., soil samples (Table 1) indicating strong adsorption of phenanthrene
2002; Xia and Ball, 1999). onto carbonaceous materials. The adsorptive contribution could
The dependency of the DH on the solid phase concentration (Cs) become significant (up to 78% in this study) at low loading levels
is summarized in Fig. 3. For the three carbonaceous materials (such as native phenanthrene loading levels). However, the deter-
(Pahokee peat, lignite, and HC), the isosteric heat becomes less mined heat contribution from adsorption process at all spiked
exothermic or endothermic with increasing Cs during sorption and aqueous concentration levels was still less than 50% except sample
desorption, respectively. This effect is consistent with a nonlinear HC (53%). For Pahokee peat, which contains w8% charcoal-like
adsorption mechanism, where DHa has the largest negative value materials, even the DH at Cs ¼ 160 mg kg1 is no more than 15e
G. Wang, P. Grathwohl / Environmental Pollution 175 (2013) 110e116 115

45 desorption from these same samples during column leaching ex-


periments with stepwise temperature increases which are reported
lignite in Wang and Grathwohl (2009). Overall DH indicates no significant
40
sorption/desorption hysteresis of phenanthrene within the samples
investigated in this study.
Isosteric heat [kJ/mol]

35 HC
Acknowledgements
30
Pahokee This research was funded by Deutsche Forschungsgemeinschaft
25 peat (DFG) under project number GR 971/16-1 and “AquaTerra”, a Eu-
ropean Union FP6 integrated project (Project no. 505428 (GOCE)).
20 The authors also thank Renate Seelig, Bernice Nisch and Renate
Riehle for their assistance in the geochemistry laboratory, Tuebin-
gen University.
15
0.1 1 10 100 1000 10000 100000
References
Cs [mg/kg]
45 Adams, R.G., Lohmann, R., Fernandez, L.A., Macfarane, J.K., Gschwend, P.M., 2007.
Polyethylene devices: passive samplers for measuring dissolved hydrophobic
organic compounds in aquatic environments. Environ. Sci. Technol. 41, 1317e
40
AS5 1323.
Allen-King, R.M., Grathwohl, P., Ball, W.P., 2002. New modeling paradigms for the
sorption of hydrophobic organic chemicals to heterogeneous carbonaceous
Isosteric heat [kJ/mol]

35 matter in soils, sediments, and rocks. Adv. Water Resour. 25, 985e1016.
Arp, H.P.H., Schwarzenbach, R.P., Goss, K.U., 2008. Ambient gas/particle partitioning.
2: the influence of particle source and temperature on sorption to dry terrestrial
30 aerosols. Environ. Sci. Technol. 42, 5951e5957.
Chiou, C.T., Peters, L.J., Freed, V.H., 1979. A physical concept of soil-water equilibria
MS4 for nonionic organic compounds. Sci. (Washington, DC) 206, 831e832.
25 Chiou, C.T., Shoup, T.D., Porter, P.E., 1985. Mechanistic roles of soil humus and
minerals in the sorption of nonionic organic compounds from aqueous and
organic solutions. Org. Geochem. 8, 9e14.
20 Chiou, C.T., 2002. Partition and Adsorption of Organic Contaminants in Environ-
mental Systems. John Wiley & Sons, Inc., New Jersey.
Costanza, J., Pennell, K.D., 2007. Distribution and abiotic degradation of chlorinated
15 solvents in heated field samples. Environ. Sci. Technol. 41, 1729e1734.
Farrell, J., Hauck, B., Jones, M., 1999. Thermodynamic investigation of trichloro-
0.01 0.1 1 10 100 1000 10000 ethylene adsorption in water-saturated microporous adsorbents. Environ.
Toxicol. Chem. 18, 1637e1642.
Cs [mg/kg] Grathwohl, P., 2011. Fate and transport of organic compounds in(to) the subsurface
environment (Chapter 8). In: Xing, Baoshan, Senesi, Nicola, Huang, Pan Ming
Fig. 3. Isosteric heats of sorption and desorption of phenanthrene (with standard error (Eds.), Biophysico-chemical Processes of Anthropogenic Organic Compounds in
bars) as a function of sorbed concentrations (Cs). Squares, triangles, diamonds, stars Environmentals Systems. Wiley and Sons, pp. 215e231.
and crosses represent the phenanthrene sorption onto lignite, high-volatile bitumi- Haftka, J.J.H., Govers, H.A.J., Parsons, J.R., 2010. Influence of temperature and origin
nous coal (HC), Pahokee peat, and the soils AS5 and MS4, respectively, whereas the of dissolved organic matter on the partitioning behavior of polycyclic aromatic
corresponding open circles indicate isosteric heats of desorption. Filled circles repre- hydrocarbons. Environ. Sci. Pollut. Res. 17, 1070e1079.
sent the isosteric heats for desorption of native phenanthrene from the different Huang, W., Weber Jr., W.J., 1997. Thermodynamic considerations in the sorption
sorbents except for HC which could not be determined. of organic contaminants by soils and sediments. 1. The isosteric heat
approach and its application to model inorganic sorbents. Environ. Sci.
Technol. 31, 3238e3243.
20% higher than that at Cs ¼ 4000 mg kg1, implicating the domi- Karapanagioti, H.K., Childs, J., Sabatini, D.A., 2001. Impacts of heterogeneous organic
matter on phenanthrene sorption: different soil and sediment samples. Envi-
nant partitioning effect. The separated isosteric heat contributions ron. Sci. Technol. 35, 4684e4690.
from this study reinforce the importance of organic contaminant Kleineidam, S., Schüth, C., Grathwohl, P., 2002. Solubility-normalized combined
uptake via partitioning by natural carbonaceous materials at pore-filling-partitioning sorption isotherms for organic pollutants. Environ. Sci.
Technol. 36, 4689e4697.
moderate-to-high concentrations. For common soils with less or no Kleineidam, S., Rügner, H., Grathwohl, P., 2004. Desorption kinetics of phenan-
carbonaceous substances, the isotherms are expected to be more threne in aquifer material lacks hysteresis. Environ. Sci. Technol. 38, 4169e4175.
linear because of the dominant partition uptake with the soil Krauss, M., Wilcke, W., 2001. Prediction soil-water partitioning of polycyclic aro-
matic hydrocarbons and polychlorinated biphenyls by desorption with
organic matter. methanol-water mixtures at different temperatures. Environ. Sci. Technol. 35,
The determined sorption/desorption isosteric heats can also be 2319e2325.
used to observe sorption/desorption hysteresis in a thermodynamic Lei, Y.D., Wania, F., Shiu, W.Y., Boocock, D.G.B., 2000. HPLC-based method for esti-
mating the temperature dependence of n-octanol-water partition coefficients.
view. The DH values should differ significantly between sorption J. Chem. Eng. Data 45, 738e742.
and desorption if hysteresis is significant. In this study, comparison Lohmann, R., 2012. Critical review of low-density polyethylene’s partitioning and
of the absolute values of isosteric heats for sorption and desorption diffusion coefficients for tracer organic contaminants and implication for its use
as a passive sampler. Environ. Sci. Technol. 46, 606e618.
of phenanthrene (DHa and DHd, respectively) shows only slight
Lou, X., Janssen, H.G., Cramers, C.A., 1997. Parameters affecting the accelerated
deviations that are not significant considering error bars (Fig. 3). solvent extraction of polymeric samples. Anal. Chem. 69, 1598e1603.
This indicates that no significant changes occurred concerning the Lüres, F., Ten Hulscher, Th.E.M., 1996. Temperature effect on the partitioning of
polycyclic aromatic hydrocarbons between natural organic carbon and water.
interaction mechanism of sorbate and sorbent during the temper-
Chemosphere 33, 643e657.
ature driven sorption/desorption cycles. In addition, the isosteric Mader, B.T., Uwe-Goss, K., Eisenreich, S.J., 1997. Sorption of non-ionic, hydrophobic
heats of desorption of the native phenanthrene are also in rea- organic chemicals to mineral surfaces. Environ. Sci. Technol. 31, 1079e1086.
sonable agreement with the DHa,d values obtained from the spiked Madlener, I., Henzler, R., Grathwohl, P., 2003. Material investigations to determine
the leaching behaviour of PAH at elevated temperatures. In: Halm, D.,
experiments (see Fig. 3 and Table 3). These findings are further Grathwohl, P. (Eds.), Proceedings of the 2nd International Workshop on
confirmed by activation energies determined during exhaustive Groundwater Risk Assessment at Contaminated Sites (GRACOS) and Integrated
116 G. Wang, P. Grathwohl / Environmental Pollution 175 (2013) 110e116

Soil and Water Protection (SOWA). Center for Applied Geoscience, Tübingen, Wang, G., Kleineidam, S., Grathwohl, P., 2007. Sorption/desorption reversibility of
Germany, pp. 193e198. http://w210.ub.uni-tuebingen.de/portal/tga_c. phenanthrene in soils and carbonaceous materials. Environ. Sci. Technol. 41,
May, W.E., Waslk, S.P., Miller, M.M., Tewarl, Y.B., Brown-Thomas, J.M., Goldberg, R.N., 1186e1193.
1983. Solution thermodynamics of some slightly soluble hydrocarbons in water. Wang, G., Grathwohl, P., 2009. Activation energies of phenanthrene desorption
J. Chem. Eng. Data 28, 197e200. from carbonaceous materials: column studies. J. Hydrol. 369, 234e240.
Pitta, J.J., Bachus, D.A., Capel, P.D., Eisenreich, S.J., 1996. Temperature-dependent Wauchope, R.D., Getzen, F.W., 1972. Temperature dependence of solubilities in
sorption of naphthalene, phenanthrene, and pyrene to low organic carbon water and heats of fusion of solid aromatic hydrocarbons. J. Chem. Eng. Data 17,
aquifer sediments. Environ. Sci. Technol. 30, 751e760. 38e41.
Pre, P., Delage, F., Faur-Brasquet, C., Cloirec, P.L., 2002. Quantitative structure- Weber, W.J., Mcginley, P.M., Katz, L.E., 1992. A distributed reactivity model for
activity relationships for the prediction of VOCs adsorption and desorption sorption by soils and sediments. 1. Conceptual basis and equilibrium assess-
energies onto activated carbon. Fuel Process. Technol. 77-78, 345e351. ments. Environ. Sci. Technol. 26, 1955e1962.
Schwarz, F.P., 1977. Determination of temperature dependence of solubilities of Wennrich, L., Popp, P., Moder, M., 2000. Determination of chlorophenols in soils
polycyclic aromatic hydrocarbons in aqueous solutions by a fluorescence using accelerated solvent extraction combined with solid-phase micro-
method. J. Chem. Eng. Data 22, 273e277. extraction. Anal. Chem. 72, 546e551.
Su, Y.H., Zhu, Y.G., Sheng, G., Chiou, C.T., 2006. Linear adsorption of nonionic organic Werth, C.J., Reinhard, M., 1997. Effects of temperature on trichloroethylene
compounds from water onto hydrophilic minerals: silica and alumina. Environ. desorption from silica gel and natural sediments. 1. Isotherms. Environ. Sci.
Sci. Technol. 40, 6949e6954. Technol. 31, 689e696.
Ten Hulscher, Th.E.M., Cornelissen, G., 1996. Effect of temperature on sorption Woodburn, K.B., Lee, L.S., Rao, P.S.C., Delfino, J.J., 1989. Comparison of sorption en-
equilibrium and sorption kinetics of organic micropollutants e a review. Che- ergetics for hydrophobic organic chemicals by synthetic and natural sorbents
mosphere 32, 609e626. from methanol/water solvent mixtures. Environ. Sci. Technol. 23, 407e413.
Tremblay, L., Kohl, S.D., Rice, J.A., Gagne, J.P., 2005. Effects of temperature, salinity, Xia, G.S., Ball, W.P., 1999. Adsorption-partitioning uptake of nine low-polarity
and dissolved humic substances on the sorption of polycyclic aromatic hydro- organic chemicals on a natural sorbent. Environ. Sci. Technol. 33, 262e269.
carbons to estuarine particles. Marine Chem. 96, 21e34. Xing, B., Pignatello, J.J., 1997. Dual-mode sorption of low-polarity compounds in
Verschueren, K., 1983. Handbook of Environmental Data on Organic Chemicals, glassy poly (vinyl chloride) and soil organic matter. Environ. Sci. Technol. 31,
second ed. van Nostrand Reinhold Company Inc., New York. 792e799.

Anda mungkin juga menyukai