10 1016@j Scitotenv 2019 03 013

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Science of the Total Environment 668 (2019) 1064–1076

Contents lists available at ScienceDirect

Science of the Total Environment

journal homepage: www.elsevier.com/locate/scitotenv

Metal solubility and transport at a contaminated landfill site – From the


source zone into the groundwater
Uddh Söderberg T. a,⁎, Berggren Kleja D. b,c, Åström M. a, Jarsjö J. d, Fröberg M. b, Svensson A. a, Augustsson A. a
a
Department of Biology and Environmental Science, Linnaeus University, Kalmar, Sweden
b
Swedish Geotechnical Institute, Olaus Magnus väg 35, Linköping, Sweden
c
Department of Soil and Environment, Swedish University of Agricultural Sciences, Box 7014, Uppsala, Sweden
d
Department of Physical Geography, Stockholm University, Stockholm, Sweden

H I G H L I G H T S G R A P H I C A L A B S T R A C T

• Metal solubility and transport from a


contaminated landfill site were investi-
gated.
• Numerous methods were adopted in
this high-resolution field- and labora-
tory study.
• Efficient cation retention, likely due to
high pH. As and Sb showed higher mo-
bility.
• Waste zone Kd values were highly vari-
able but not correlated to basic geo-
chemistry.
• Soil metal concentrations decreased
abruptly below the landfill.

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

Article history: Risks associated with metal contaminated sites are tightly linked to material leachability and contaminant mobil-
Received 20 December 2018 ity. In this study, metal solubility and transport were characterized within a glass waste landfill through
Received in revised form 25 February 2019 i) lysimeter-collection of pore water and standardized batch leaching tests, ii) soil profiles extending from the
Accepted 1 March 2019
landfill surface, through unsaturated soil underneath, and into the groundwater zone, and iii) groundwater sam-
Available online 3 March 2019
ples upstream, at, and downstream of the landfill. The soil analyzes targeted both pseudo-total and geochemi-
Editor: Jay Gan cally active concentrations of contaminant metals (As, Cd, Pb, Sb) and basic soil geochemistry (pH, org. C, Fe,
Mn). Water samples were analyzed for dissolved, colloid-bound and particulate metals, and speciation modelling
Keywords: of the aqueous phase was conducted. The results revealed a highly contaminated system, with mean metal con-
Soil and groundwater metal pollution centrations in the waste zone between 90 and 250 times the regional background levels. Despite severe contam-
Glass waste ination of the waste zone and high geochemically active fractions (80–100%) of all contaminant metals as well as
Soil:Solution partitioning (Kd) elevated concentrations in landfill pore water, the concentrations of Cd and Pb decrease abruptly at the transition
Leachability between landfill and underlying natural soil and no indication of groundwater contamination was found. The ef-
Mobility
ficient cation retention is likely due to the high pH. However, the sorption of As and Sb is weaker at such high pH,
Colloids
which explains their higher mobility from the pore water zone into groundwater. The field soil:solution

⁎ Corresponding author.
E-mail address: terese.uddh-soderberg@lnu.se (T. Uddh Söderberg).

https://doi.org/10.1016/j.scitotenv.2019.03.013
0048-9697/© 2019 The Authors. Published by Elsevier B.V. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).
T. Uddh Söderberg et al. / Science of the Total Environment 668 (2019) 1064–1076 1065

partitioning (Kd) displayed a high spatial variability within the waste zone (the highest Kd variability was seen
for Pb, ranging from 140 to 2,900,000 l kg−1), despite little variability in basic geochemical variables, which we
suggest is due to waste material heterogeneity.
© 2019 The Authors. Published by Elsevier B.V. This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/by-nc-nd/4.0/).

1. Introduction assess contaminant travel times and/or total transport in pore water
and groundwater systems (Jarsjö et al., 2017; Malmstrom et al., 2008;
According to the European Environment Agency, there are about Persson et al., 2011).
2.5 million potentially contaminated sites within the EU member states, To contribute to filling the knowledge gaps regarding metal solubil-
and the contamination has in the majority of cases arisen from waste ity, partitioning and transport at sites with waste material contamina-
disposal (Van Liedekerke et al., 2014). A key factor for the risks posed tion, the first aim of this paper is to present how metals are released
by metal pollutants at such settings is the potential for leaching, mobili- to pore water in the waste zone (landfill) of an industrial site affected
zation and subsequent spreading from the primary source areas. The by N100 years of glass production, and the subsequent metal migration
characterization of contaminant spreading thus plays a central role in through the natural soil under the landfill and finally to groundwater.
the risk assessment frameworks adopted by many national environ- Additional aims are to define within-site variability in soil:solution
mental agencies. The generic approaches are, however, both simplified partitioning (Kd values) and to investigate how well standardized
and uncertain, where a lack of holistic studies with field measurements batch-leaching tests reflect metal solubility in the waste zone.
of metal release from contaminated soil/waste horizons and subsequent
migration into hydrological systems (Pandey et al., 2014; Sungur et al.,
2015) leaves us with critical knowledge gaps regarding field-scale 2. Method
metal mobility and transport, and net effects in downstream recipients.
In this context, the present paper addresses, firstly, waste zone metal 2.1. Study site
release. In internationally well-accepted and spread risk assessment
models, such as those from the US EPA and Dutch RIVM, soil:solution The study was conducted at Pukeberg glassworks site (Fig. 1), where
partitioning (defined by the Kd value; l kg−1) are used to approximate coloured and crystal glass have been manufactured since 1871. This site
the contaminant mobilization in the source zone (Brand et al., 2007; is one of the most extensively investigated glassworks settings in Kal-
Swedish, 2009; U.S. EPA., 1991). The Kd value is expressed as a constant mar and Kronoberg counties in south-eastern Sweden. The major raw
based on the metal concentration associated with the solid soil phase materials used by the glassworks were sand of quartz and feldspar,
(mg/kg) and that of the soil solution (mg/l) (Allison and Allison, 2005; soda (NaCO3), potash (K2CO3), calcite (CaCO3), dolomite (MgCO3
U.S. EPA., 1999). Site specific Kd values may be determined from labora- ·CaCO3), witherite (BaCO3) and oxides of different metals, for example
tory leaching tests; either in batch or column set-ups or according to ex- As, Cd, Co, Cr, Cu, Ni, Pb, Sb and Zn (Hermelin and Welander, 1986;
traction protocols (TS 14405, ISO/TS 21268-2, TS 14429, TS 149977, SS- Magnusson, 1971). While the majority of modern glass waste (bottles/
EN 12457-2), or sometimes from field sampling of soil and pore water. jars and industrial glass waste) does not contain toxic metals in signifi-
Few have, however, compared the laboratory and field approaches cant quantities, and thus does not release such elements through
(Sauvé et al., 2000). Anyway, we know that soil:solution partitioning leaching (Disfani et al., 2012; Imteaz et al., 2012), coloured artistic
differs greatly between sites, and in general it is well linked to differ- glass and crystal glass of the kind produced in the study region often
ences in basic soil or pore water geochemistry (Carlon et al., 2004; contain high amounts of a range of metals (Hermelin and Welander,
Degryse et al., 2009; Janssen et al., 1997; Sauvé et al., 2000). Considering 1986; Magnusson, 1971). Most of the waste generated (crushed glass,
this naturally high variability in soil:solution partitioning, it may not discarded raw material batches, grinding and acid sludges, old oven
come as a surprise that soil contamination is sometimes followed by and demolition debris etc.) has been deposited on the factory's landfill,
groundwater contamination as well (Khalil, 2012; Kumari et al., 2014; which covers an area of approximately 15,000 m2 and has an average
Zheng et al., 2013), and sometimes not (Ahmed and Sulaiman, 2001; depth of 1.6 m. The waste is currently mixed with natural soil, partly
Clausen et al., 2011; Jensen et al., 2000). However, while several studies due to various forms of excavation work at the property, but also as a re-
have characterized metal solubility or Kd values for different media and sult of the establishment of vegetation and soil at the surface of the
their connection to soil properties, a review of 245 scientific papers and landfill since it was taken out of use in the 1970s.
reports by Allison and Allison (2005) concluded that none of the inves- The site is located in the northern part of the Nybro esker, in an area
tigated papers addressed waste systems (waste piles, landfills etc.). Al- mainly containing sandy glacifluvial materials (SGU, 2017b). The pri-
most 15 years later that conclusion still seems to hold. The studies mary rock is granite (SGU, 2017b), with a 4–9 m thick overlying layer
which have addressed Kd variability have dealt mainly with natural of Quaternary deposits (SGU, 2017b). The area around the landfill gently
soil, and exclusively characterized differences between sites (de Souza slopes from southwest to northeast, which also is the general ground-
Braz et al., 2013; Fanger et al., 2006; Janssen et al., 1997), thus leaving water flow direction (Fig. 1). The hydraulic gradient is about 2% and
intra-site variability poorly investigated. From a risk assessment per- the hydraulic conductivity on average 3.5 × 10−6 m/s. These parameters
spective, this means uncertainties regarding the representativeness were calculated from water level measurements in groundwater sam-
when applying single Kds. pling wells at the site and the Bouwer and Rice Slug-test (Bouwer,
Laboratory experiments have contributed to a relatively detailed un- 1989). The local catchment area is approximately 0.30 km2, estimated
derstanding of processes involved in metal(loid) leaching from contam- from topographical maps (Dingham, 2002; Stone et al., 2015). The land-
inated soil and waste matrices under different conditions, e.g. according fill is located at the northeast end of this catchment (Fig. 1), where dis-
to de Matos et al. (2001), but an outlook beyond these small lab scales charge into the St Sigfrid stream occurs. The theoretical dilution (under
and well-defined conditions reveals unanswered questions regarding full mixing) of pore water that penetrates from the landfill into the
the whole process chain from waste material leaching and metal release groundwater beneath is thus about 20, calculated as the ratio between
to pore water, transport in the unsaturated zone, and finally migration the catchment area upstream of the landfill (0.30 km2) and the area of
into groundwater. Consequently, it is not surprising that available the landfill (0.015 km2). The water infiltration of the region is estimated
models that aim to simulate metal transport often fail to adequately to about 0.18 m/year (Törnqvist, 2015), which gives an annual water
1066 T. Uddh Söderberg et al. / Science of the Total Environment 668 (2019) 1064–1076

Fig. 1. The location of Pukeberg glassworks site in Nybro municipality, southeastern Sweden. The dashed line marks the local catchment area and the dotted line marks the approximate
landfill borders. Sampling site U1 is thus located upstream. Sampling sites L1 to L5 are located at the landfill, and D1 to D4 are located downstream.

flow of ca 2700 m3 through the contaminated wastes (0.18 m/year 1.0 g of the milled material with Aqua Regia (1HNO3:3HCl) at 90 °C in
×15,000 m2). a microprocessor controlled digestion block for 2 h.
Geochemically active concentrations of metal cations (Cd, Pb) were
determined after 16 h extraction of 1.0 g sieved soil with 30 ml 0.1 M
2.2. Soil sampling and analyzes HNO3 (Gustafsson and Berggren Kleja, 2005). The corresponding frac-
tion of As, Sb (which form anions), and of Al, Fe and Mn, were deter-
In June 2015, a combination of Sonic drilling and DTH (down the mined after extraction of 1.0 g with 100 ml 0.2 M oxalate/oxalic acid
hole) drilling was used to retrieve intact soil cores. These were collected buffer at pH 3.0, shaken for 4 h in darkness (van Reeuwijk, 2002). The
at five locations at the landfill (Fig. 1), and consisted of material from extracts were analyzed with ICP-SFMS according to SS EN ISO 17294-
i) the unsaturated waste zone, where glass waste is mixed with natural 1, 17294-2 and EPA-method 200.8.
soil, ii) the unsaturated natural soil zone, underneath the waste zone Finally, batch leaching tests were performed on all soil samples col-
and iii) the saturated zone, i.e. the groundwater zone. Three cores lected around the lysimeters and in soil profiles reaching below the
downstream of the landfill were also retrieved, collected ca 4 m from groundwater table. The aim was to obtain data on waste zone material
the landfill slope, and subsampled every 0.5 m. All in all, 48 soil samples leachability according to standard lab protocols, for comparison with
from the 8 cores were obtained. After the pore water sampling was the leachability observed in the field and the leachability along the soil
completed, soil samples were also collected within a 10 cm radius of profile. These batch leaching tests were conducted according to SS-EN
the five lysimeters that were installed for collection of pore water 12457-2 and 21268-2, where 22.5 g of field-moist soil was shaken
from the waste zone (see below). with 225 ml Milli-Q water alternative 1 mM CaCl2-solution (hence a L/
All soil samples were sieved to 2.0 mm (ISO 11464:2006) and mea- S ratio of 10) for 24 h. All batch leaching test were run on duplicate sam-
sured for pH and dry weight (ISO 10390:2005 and SS-ISO 11465). A por- ples. The collected eluates were analyzed for metal concentrations after
tion of 8.0 g was milled with a centrifugal ball mill before determination both membrane filtering (b0.45 μm) and ultra-filtration (b10 kDalton)
of organic C (org.C) and metal concentrations. Organic C was deter- with ICP-SFMS according to SS EN ISO 17294-1, 17294-2 and EPA-
mined using a Carlo Erba CN analyzer (Flash1112 series) according to method 200.8, i.e. according to the same protocols as for the pore
ISO 10694 and 13878. Pseudo-total metal concentrations were deter- water and groundwater samples (see below). The eluates were also an-
mined with a Perkin Elmer Sciex ELAN 9000 ICP/MS after digesting alyzed for dissolved organic carbon (DOC) and pH, again according to
T. Uddh Söderberg et al. / Science of the Total Environment 668 (2019) 1064–1076 1067

the same methodologies as for pore water and groundwater samples samples). Solution speciation was estimated using the thermodynamic
(see below). data of the Visual database, which mostly relies on the NIST Critical Sta-
bility constants database (Smith et al., 2003). The site-specific data used
2.3. Quality control and representativeness of sampled soil material in the model were pH, alkalinity, DOC, and concentrations of F, Cl, SO4,
PO4, Ca, Mg, Na, K, Fe, Mn, Al, Si, As, Ba, Cd, Co, Cu, Pb, Sb and Zn. To de-
To assess the homogeneity of the sieved soil samples from the waste scribe the binding of metals to DOC, the Stockholm Humic Model (SHM)
zone, eight samples were prepared as triplicates before analyzes of was used (Gustafsson, 2001; Gustafsson et al., 2003) and the Davies
pseudo-total and geochemically active metal concentrations were con- equation was used for activity correction. Saturation indices were
ducted (i.e. three different sub-portions of the material were digested/ assessed for all As, Cd, Pb and Sb mineral phases included in the
extracted and analyzed separately). All relative standard deviations Visual MINTEQ database, and also for duftite (PbCuAsO 4(OH)),
(RSD) were between 4 and 10%, implying stable metal concentrations hydroxymimetite (Pb5(AsO4)3OH), mimetite (Pb5(AsO4)3 Cl), lead
in individual sieved subsamples. This further indicates that differences silicate (PbSiO3) and shultenite (PbHAsO4). The latter four minerals
in metal concentrations between locations (subsamples from different were added as previous investigations have pointed to their poten-
cores and different depths) reflect true material heterogeneity. The sta- tial for affecting Pb and As solubility (Bajda, 2010; Inegbenebor
bility and precision of the sample digestion procedure and of the instru- et al., 1989; Lee and Nriagu, 2007; Lindsay, 1979; Magalhaes and
ment (method repeatability) was evaluated by analyzing duplicates of Dejesus, 1988; Sjöstedt et al., 2018).
14 milled samples. Again, RSDs were below 10%. Recovery rates of the
analyzed reference materials were between 83 and 112% (QC; USGS 2.6. Statistical analyzes
Geochemical Exploration Reference Materials GXR-1, GXR-4, GXR-6,
SdAR-M2), indicating a satisfactory precision in methodology. Soil:solution metal partitioning coefficients (Kd values) were ana-
lyzed with linear regression models using the program R 3.0.2 (R Core
2.4. Water sampling and analyzes Team, 2017). For each metal, a model with four predictor variables
was fitted, namely: soil pH, soil organic C, soil Fe content, and soil con-
Pore water from the waste zone was collected with tension lysime- centration of the metal in question. Kd and pH values were repeatedly
ters (Prenart Super Quartz, 2 μm pore size), installed at three locations sub-sampled at each station. Therefore, the data was analyzed as linear
at the landfill (Fig. 1) at two depths (30 and 60 cm). Sampling was con- mixed effects models with the sampling station included as a random
ducted on seven occasions, in July, August, September, and November factor. Predictors and responses were log transformed when necessary
(twice) 2015, and in February and May 2016. A total of 31 pore water to obtain homoscedastic model residuals. Models were fitted with re-
samples were collected. stricted maximum likelihoods (REML), using the function lme in the R
Six groundwater sampling tubes (30 mm PEH tubes) located up- package nlme (Pinheiro et al., 2017). We used model simplification by
stream of, directly at, and downstream of the landfill were installed stepwise elimination of non-significant predictors (criterion for dele-
(Fig. 1). The three tubes at the landfill (L2, L4 and L5) and one of the tion p N 0.05; (Crawley, 2007), so that final models only contained sig-
tubes downstream (D3) were installed in the sonic drilling holes nificant predictors (if any). The coefficients of significant predictors
where soil cores were also retrieved. Tube filters were placed to extend and the model intercepts were used to build predictive model equa-
from approximately one meter above to one meter below the ground- tions. Marginal R-squared values of the final models were calculated
water table. Sampling was conducted according to U.S. Geological Sur- using the function r.squaredGLMM in R package MuMIn (Barton, 2013).
vey Techniques of Water-Resources Investigations (USGS, 2006), using
a peristaltic pump connected to a flow cell (YSI 659 MDS), on the 2.7. Regional background concentrations
same occasions as the sampling of the tension lysimeters. A total of 34
groundwater samples were collected. Regional background concentrations in soil and groundwater
Field measurements of pH, conductivity (cond.), dissolved oxygen (delimited to a 45 km radius around Pukeberg) were retrieved from
(dis. oxygen), ORP (Oxidation-Reduction Potential) and temperature the Geological Survey of Sweden (SGU, 2017a). The soil background
were conducted on groundwater samples. pH and conductivity on pore data refer to the fine fraction of natural (C horizon) till soil. Data on Sb
water samples from the lysimeters were measured using a Hanna pH- in groundwater in the region were not available at SGU. Hence addi-
meter 8519 N and Jenwey 4510 Conductivity meter (methods SS-EN tional groundwater data were collected from the Swedish Nuclear
ISO 10523:2012 and SS-EN 27888). Alkalinity (alk.), orto-phosphate Fuel and Waste Management C. (SKB) Site Characterization database
(PO4) and UV-absorbance (254 nm) were measured on all water samples (SICADA) for Simpevarp (SKB, 2016). Simpevarp has the same primary
according to SS 028139, SS-EN ISO 6878:2005 and US EPA Method 415.3 bedrock as Pukeberg and may give representative regional background
(Perkin Elmer UV/VIS spectrometer Lambda XLS+). Dissolved organic values for Sb.
carbon was measured with IR detection (SS-EN 1484) and macro anions,
fluoride (F), chloride (Cl) and sulphate (SO4), were determined with ion 3. Results and discussion
chromatography (SS-EN ISO 10304-1). The specific UV-absorbance
(SUVA) was then determined as the UV-absorbance divided by DOC. 3.1. Solid phase characterization
Metal concentrations in both pore water and groundwater were de-
termined with ICP-SFMS according to SS EN ISO 17294-1, 17294-2 and In accordance with consultant investigations conducted between
EPA-method 200.8. The distribution of the metals and metalloids in dis- year 2002 and 2012 (Elert and Höglund, 2012; Fanger et al., 2003), the
solved, particulate or colloid bound forms, were investigated by analyz- metals As, Cd, Pb and Sb were clearly elevated in the landfill waste
ing digested, membrane filtered (0.45 μm) and ultra-filtered (b10 kD) zone, with mean pseudo-total metal concentrations exceeding the
samples, respectively. background levels of regional till soil by factors of about 120, 160, 90
and 250, respectively (Table 1). The fraction of geochemically active
2.5. Modelling of metal speciation in water with Visual MINTEQ forms ranged from 80% (As) to 100% (Cd) (Fig. 2a), while concentrations
after batch leaching with CaCl2 were below 1.5%.
Visual MINTEQ (Gustafsson, 2016) was used to calculate the free ion The highest metal concentrations of the waste zone were found to-
activity, saturation indices for minerals and percentage distribution of wards the bottom of the landfill, but they decreased abruptly in the un-
dissolved species in all water samples (membrane filtered and ultra- derlying unsaturated natural soil zone (Fig. 2a); reaching background
filtered samples) and eluates from batch leaching tests (ultra-filtered concentrations already within the upper 20 cm for Cd and Pb and
1068 T. Uddh Söderberg et al. / Science of the Total Environment 668 (2019) 1064–1076

Table 1
Soil pseudo-total metal concentrations in the three analyzed zones at the landfill and the unsaturated natural soil zone downstream the landfill, together with basic physicochemical pa-
rameters. Metal concentrations in mg/kg dry weight. Data from each individual soil core is provided in the Supplementary information (Table S-1).

Contaminant metals

As Cd Pb Sb Fe Al Mn pH (H2O) Org. C (%)

Regional background (45 km radius around Pukeberg), N = 34. Depth: approximately 0.6–1.0 m
Min-Max 0.57–2.6 0.031–0.18 6.2–59 0.057–0.32 6300–32,500 4900–22,500 92–450 4.3–5.6
Mean ± SE 1.2 ± 0.091 0.078 ± 0.006 18 ± 2.3 0.15 ± 0.0091 16,400 ± 770 10,800 ± 920 230 ± 14 5 ± 0.056

Unsaturated glass waste zone, N = 13. Depth: approximately 0–1.5 m


Min-Max 35–390 3.0–22 560–4300 9.9–78 10,300–25,000 3000–12,000 190–1900 6.8–7.6 1.2–7.8
Mean ± SE 140 ± 30 13 ± 1.8 1700 ± 270 38 ± 5.1 16,000 ± 1100 9100 ± 710 510 ± 120 7.2 ± 0.1 3.0 ± 0.5

Unsaturated natural soil zone under the landfill, N = 20. Depth: approximately 1.5–4.0 m
Min-Max 0.8–36 0.005–0.3 11–41 0.40–9.8 11,500–18,000 6500–15,000 170–320 6.8–8.3 0.006–0.42
Mean ± SE 5.2 ± 1.8 0.087 ± 0.021 19 ± 1.8 2.2 ± 0.49 14,000 ± 410 10,000 ± 510 240 ± 8.0 7.6 ± 0.090 0.089 ± 0.024

Saturated natural soil zone under the landfill, N = 4. Depth: approximately 4.0–5.5 m
Min-Max 2.8–3.9 0.053–0.10 6.4–20 0.38–0.75 15,500–22,000 11,300–18,000 280–490 8.0–8.3 0.006–0.027
Mean ± SE 3.3 ± 0.23 0.088 ± 0.012 16 ± 3.1 0.48 ± 0.090 18,000 ± 1600 14,000 ± 1500 370 ± 45 8.1 ± 0.058 0.014 ± 0.006

Unsaturated natural soil zone downstream the landfill, N = 11. Depth: approximately 0–2.0 m
Min-Max 0.8–2.5 0.005–0.10 12–28 0.10–0.40 12,000–19,000 8000–13,000 200–530 5.2–7.3 0.012–3.3
Mean ± SE 1.6 ± 0.18 0.055 ± 0.015 17 ± 1.4 0.21 ± 0.028 14,000 ± 620 9600 ± 470 260 ± 29 6.0 ± 0.21 0.68 ± 0.30

70 cm for As and Sb. All contaminant metal concentrations in all soil solution. This makes Pb the element with by far the lowest fraction as
samples downstream of the landfill slope, about 4 m away, were also free ions. The complexation with Cd was also mainly with DOC (27%)
comparable or close to regional background levels (Table 1). and carbonates (9.4%). Minor complexation (b1%) was simulated with
Concentrations of Fe and Al represented levels that are normal for hydroxides, sulphates, phosphates, chlorides and fluorides. In general,
the natural till soil of the region and Mn was slightly elevated the distribution between free dissolved ions, dissolved inorganic ions
(Table 1). The variable which stands out as non-representative is the and organic complexes was similar in the waste zone pore water sam-
pH value, which was clearly elevated in both the waste zone and ples and in groundwater (Fig. 3b), even though the surrounding solid
below (Table 1). phase material differ significantly in composition.
Despite the relatively high contaminant metal concentrations in
3.2. In situ aqueous phase characterization pore water of the waste zone, none of the mineral phases plausible for
the four metals according to the Visual MINTEQ database were assessed
Within the waste zone, elevated metal concentrations were also to reach supersaturation. All groundwaters were undersaturated with
found in the lysimeter pore waters, where the concentrations of all respect to the mineral phases investigated. Saturations indices for min-
metals were significantly higher than in both local and regional ground- erals, simulated in Visual MINTEQ, are presented in the Supplementary
water (Table 2). The difference between pore water and landfill ground- information (Table S-4).
water concentrations were on average 200 and 910 times for Cd and Pb,
respectively (Fig. 2b); thus, much larger than one would expect from di- 3.3. Leaching of metals from glass waste material and subsequent migration
lution (applying the theoretical dilution factor of 20 presented in the to groundwater
site description (Section 2.1)). The concentrations of Cd and Pb in
groundwater wells at the landfill were comparable to regional back- The observed high fraction of geochemically active metal forms
ground levels, and also not elevated when compared to the sampling (Fig. 2a) in association with elevated metal concentrations in pore wa-
well upstream (Table 2). They also complied to the drinking water stan- ters of the waste zone (Table 2; Fig. 3) imply that a significant part of
dards of the World Health Organization (WHO, 2011). Groundwater As the waste zone material is susceptible to leaching. That metal concen-
and Sb were, however, elevated compared to both regional background trations clearly increase towards the bottom of the landfill (Fig. 2a) is,
levels and drinking water criteria in several sampling tubes at and however, more likely due to older and younger wastes differing in com-
downstream of the landfill (Table 2). Thus, these elements were less position than to the migration of contaminants from upper landfill hori-
strongly retained than the cationic metals. zons. This, we argue, is because of the sharp disappearance of metal
The most distinct features of the basic water chemistry, summarized contaminants in the unsaturated soil underneath the landfill (Fig. 2a).
in Table 3, were the high alkalinity – matching the high pH values of the Before 1969, for example, the glass raw material was prepared in the
system - and the high fluoride concentrations. The remaining variables glass works factory. It involved mixing of a variety of powdered metal
were within normal ranges for till soils of the region (SGU, 2016; SKB, oxides, and discarded batch mixtures were thrown at the landfill, prob-
2016), and all groundwater samples were oxic (Eh N290 mV). The ably leading to both higher total metal concentrations than in younger
high metal concentrations in the waste zone pore water were mainly waste materials and a higher leachability of the material.
(N67%) associated with dissolved forms, even though colloids were sig- Although leachable in the waste zone, the sharp decrease in concen-
nificant carriers for Pb (24% of the total content) and Sb (15%) (Fig. 3a). trations of cation-forming metals (Pb and Cd) immediately under the
Association with colloids was, however, non-existent for all metals in landfill (Fig. 2a) implies that these metals are efficiently trapped over
the groundwater (Fig. 3a). very short distances (decimeters) as the metal-rich pore water perco-
From the Visual MINTEQ modelling we see that the soluble As and lates into the natural soil. It has previously been shown that little or
Sb, in both pore water and groundwater, occurred entirely as free, reac- no impact from metal leaching may be detected in groundwater down-
tive ions (HAsO42−, H2AsO4−, Sb(OH)6−) in the soluble (b10 kD) fraction. stream sites with landfilled mixed municipal waste (Christensen et al.,
In contrast, the two cations showed some degree of complex formation 2001; Kjeldsen et al., 2002; Mor et al., 2006), although examples to
(Fig. 3b). In pore water, the organic complexes dominated for Pb (86% of the contrary also exist (Kiddee et al., 2014). Our results demonstrate
the total dissolved Pb). The second most important Pb complexing li- how efficient, and over how short distances, the immobilization of
gand was carbonate, accounting for 12% of the total Pb speciation in aqueous species may be under favorable conditions, at least for cations.
T. Uddh Söderberg et al. / Science of the Total Environment 668 (2019) 1064–1076 1069

Fig. 2. a. Depth profiles of metal concentrations in soil at the landfill area. The concentration given for a certain depth interval represents the average for all sampling sites. The error bars
represent the standard error (SE). Profiles for each soil core are presented in the Supplementary information (Fig. S-1). b. Depth profiles of metal concentrations in aqueous phase at the
landfill area. The concentration given for a certain depth interval represents the average for all sampling sites. The error bars represent the standard error (SE).
1070 T. Uddh Söderberg et al. / Science of the Total Environment 668 (2019) 1064–1076

Table 2
Metal concentrations in membrane filtered water (b0.45 μm). Data from each individual lysimeter and groundwater monitoring wells is provided in the Supplementary information
(Table S-2).

Contaminant metals

As Cd Pb Sb Ca Mg Na K Fe Al Mn

μg/l μg/l μg/l μg/l mg/l mg/l mg/l mg/l μg/l μg/l μg/l

Pore water collected with lysimeters in the waste zone of the landfill (sample N = 31)
Min-Max 0.77–97 2.1–32 2.1–860 6.6–530 4.3–180 0.81–31 11–330 6.6–83 3.0–620 0.030–7800 0.0045–2800
Mean ± SE 24 ± 4 13 ± 1.7 84 ± 30 190 ± 26 89 ± 7.9 5.9 ± 1.3 62 ± 15 29 ± 4.5 50 ± 22 650 ± 340 270 ± 130

Groundwater from monitoring wells at the landfill (sample N = 21)


Min-Max 0.035–15 0.012–0.28 0.0055–0.87 4.8–28 34–72 7.5–11 18–57 12–17 0.17–11 13–66 0.10–510
Mean ± SE 5.3 ± 1.2 0.065 ± 0.013 0.092 ± 0.042 13 ± 1.6 52 ± 2.2 9.0 ± 0.24 24 ± 2.0 15 ± 0.34 2.2 ± 0.52 37 ± 2.9 12 ± 6.6

Groundwater from monitoring wells downstream the landfill (sample N = 11)


Min-Max 0.19–3.1 0.066–0.23 0.0029–0.17 9.9–47 22–79 5.6–12 14–25 6.6–17 0.10–19 7.4–98 5.3–6.1
Mean ± SE 1.5 ± 0.34 0.15 ± 0.018 0.029 ± 0.016 29 ± 4.7 48 ± 6.6 8.2 ± 0.65 19 ± 1.2 11 ± 1.2 3.4 ± 1.7 45 ± 11 5.9 ± 0.082

Groundwater from monitoring wells upstream the landfill (sample N = 3)


Min-Max 0.74–2.2 0.020–0.042 0.013–0.038 0.33–0.50 12–16 2.8–3.7 7.4–8.8 3.0–3.1 0.91–1.6 29–38 1.1–9.8
Mean 1.5 0.028 0.026 0.41 14 3.2 8.1 3.1 1.2 33 5.5

Regional background concentrations in groundwater


N 114 143 142 291 759 764 738 737 686 436 667
Min-Max 0.010–1.0 0.003–0.31 0.005–4.4 0.012–2.2 0.92–110 0.79–14 3.2–71 0.12–5.7 0.30–9100 1.8–1100 0.10–1600
Mean 0.23 0.038 0.61 0.1 14 2.9 12 1.7 220 99 150

3.4. Retention mechanisms Pukeberg, as indicated by relatively high concentrations of Fe, Al and
Mn and fair amounts of organic carbon (Table 1). That As and Sb are
The waste zone solid material in Pukeberg, as well as the pore water less efficiently retained is probably also largely due to the high pH of
and groundwater sampled at the landfill area, are clearly affected by the the system, where the decrease in positive charged functional groups
extensive use of various carbonates in glass production. We have previ- is not beneficial for the immobilization of these anion-forming elements
ously described a system of high pH, alkalinity and conductivity, with (Goldberg, 2002; Johnson et al., 2005; Leuz et al., 2006; Wang and
high concentrations of base cations (Table 2), and with carbonates dom- Mulligan, 2006; Violante et al., 2010).
inating the inorganic complexation in the aqueous phase (Fig. 3b). The The low contribution of potentially mobile aqueous complexes of Cd
majority of the metal cation retention in the waste zone is therefore (Fig. 3) also favors immobilization processes. For Pb, complexes with
probably due to the high pH. Within the current pH interval (7.2–8.2), dissolved organic matter are important in the pore waters of the
the cation sorption is expected to be highly efficient for Cd and Pb; waste zone as well as in the groundwater. This organic matter complex-
nearly 100% has previously been reported (Jalali and Najafi, 2018; ation obviously facilitates mobility, as described by e.g. Christensen et al.
Ndiba and Axe, 2010; Rijkenberg and Depree, 2010; Tahervand and (1996), Clausen et al. (2011) and Rijkenberg and Depree (2010). The
Jalali, 2016; Yang et al., 2003; Zang et al., 2017). This pH-depending re- overall low Pb mobility, however, points to a strong binding of DOC
tention is explained mainly by intensified electrostatic adsorption to complexes to the solid phase. In support of this hypothesis there are
negatively charged functional groups and/or co-precipitation with Fe/ studies which have shown that the release of DOC to pore water may
Al/Mn oxides (Alloway, 2013; Bolan et al., 2014; Naidu et al., 1997; occur parallel to efficient immobilization processes (Kaiser and
Violante et al., 2010). The pre-conditions for this process are good in Kalbitz, 2012). The potential of DOC retention in soil systems is also

Table 3
Physicochemical parameters in membrane filtered water (b0.45 μm). Data from each individual lysimeter and groundwater monitoring wells is provided in the Supplementary informa-
tion (Table S-3).

pH Cond. Alk. Dis. oxygen Eh DOC SUVA (254 nm) F Cl SO4 PO4

μS/cm mg HCO3/l DO % mV mg/l l/mg-m mg/l mg/l mg/l μg/l

Pore water collected with lysimeters in the waste zone of the landfill (sample N = 31)
Min-Max 6.1–7.6 110–1900 16–1000 8.7–58 1.4–4.3 0.64–15 0.50–12 2.5–220 7.1–140
Mean ± SE 6.9 ± 0.084 760 ± 86 350 ± 55 16 ± 2.1 2.5 ± 0.13 7.7 ± 0.71 3.2 ± 0.57 19 ± 7.2 50 ± 6.7

Groundwater from monitoring wells at the landfill (sample N = 21)


Min-Max 5.9–6.6 160–570 120–250 40–83 290–510 1.9–5.5 0.48–1.9 1.2–3.6 14–22 21–65 1.0–18
Mean ± SE 6.3 ± 0.042 400 ± 25 180 ± 8.2 63 ± 3.0 430 ± 12 3.6 ± 0.20 1.2 ± 0.10 2.1 ± 0.14 19 ± 0.52 40 ± 2.2 5.9 ± 1.3

Groundwater from monitoring wells downstream the landfill (sample N = 11)


Min-Max 5.3–6.1 170–520 69–240 54–76 400–520 1.9–4.7 0.53–1.7 0.63–5.2 13–29 12–49 1.0–17
Mean ± SE 5.9 ± 0.082 330 ± 31 150 ± 21 66 ± 1.9 520 ± 11 2.8 ± 0.26 1.2 ± 0.13 2.8 ± 0.57 19 ± 1.2 28 ± 3.5 7.6 ± 1.8

Groundwater from monitoring wells upstream the landfill (sample N = 3)


Min-Max 5.8–5.9 79–120 28–22 67–77 440–480 1.9–2.3 0.94–1.5 0.36–0.44 11–14 14–15 23–36
Mean ± SE 5.9 ± 0.045 100 ± 22 31 ± 2.5 72 ± 5.1 460 ± 24 2.1 ± 0.23 1.2 ± 0.30 0.40 ± 0.040 12 ± 1.3 14 ± 0.25 30 ± 6.5

Regional background concentrations in groundwater


N 897 895 688 38 238 674 738 735 545
Min-Max 4.6–8.5 35–622 0.20–300 280–530 0.6–26 0.040–1.8 2.1–135 2.4–88 2.0–110
Mean 6.3 150 34 430 3.2 0.31 19 15 14
T. Uddh Söderberg et al. / Science of the Total Environment 668 (2019) 1064–1076 1071

Fig. 3. a): Average metal partitioning in pore water (N = 7) and groundwater samples (N = 8), distributed among dissolved, colloid bound and particulate bound fractions. The
distributions were determined from ultra-filtered (10 kD), membrane filtered (0.45 μm) and digested samples. b): Distribution among free ion, inorganic complexes and DOC
complexes in dissolved (ultra-filtered, 10 kD) water samples. Results from Visual MINTEQ modelling. The As and Sb forms (HAsO24, H2AsO− −
4 , Sb(OH)6 ) are treated as free ions, despite
their association with OH−, since they behave as free (reactive) ions. The inorganic complexes involve carbonates, hydroxides, sulphates, phosphates, chlorides and fluorides.

implied by studies which present how DOC at deeper soil depths origi- around the same lysimeters, we see that concentrations are consistently
nates mainly from decomposition at these depths and not from DOC higher in the lysimeter-collected pore waters for Cd, Pb and Sb (Fig. 2b
production near the surface (Fröberg et al., 2006). + Fig. 4). This finding implies that the field leachability of metals in the
From the much lower occurrence of colloids in the groundwater waste zone is not adequately captured in the evaluated standard
zone compared to that of the waste zone in Pukeberg (Fig. 3a) one leaching tests, despite the extensive use and high acceptance of these
could hypothesize that the decrease in metal concentration could to protocols. This effect appears to be especially prominent at high concen-
some degree be due to colloid retention. While metal transport may trations, where the implications of erroneous interpretations are gener-
be governed by binding to colloidal particles (Baumann et al., 2006; ally the highest too. The maximum concentration in pore water
Karathanasis, 1999; Shang and Li, 2011), colloid association may also collected in the field was three (Sb), four (Pb) and eight (Cd) times
prevent spreading in cases where these small units are physically the maximum concentration after batch leaching. Perfect correlations
trapped (Baumann et al., 2006; Buffle and Leppard, 1995; Ranville between field and lab measurements can obviously not realistically be
et al., 1999). However, since the same decrease in metal concentrations expected since samples taken for laboratory testing represent only a
is seen for membrane filtered and ultra-filtered samples in Pukeberg, sub-portion of the total soil volume which has been in contact with
colloid filtering is not plausible as a key controlling factor for the ob- the pore water collected in the lysimeters. However, the consistent ap-
served metal retention in the waste zone. pearance below the 1:1 line of Cd, Pb and Sb indicates a systematic un-
derestimation of the material leachability in the leaching tests. The
3.5. Assessing solubility – batch leaching vs field measurements probability of the observed difference being of a systematic nature is
also strengthened by similar observations found for the groundwater
Comparing metal concentrations in lysimeter-collected pore water zone (not displayed in Fig. 4). Concentrations in groundwater samples
with eluates from laboratory batch leaching of soil samples collected from the installed tubes were between two (Pb and Sb) and five (Cd)
1072 T. Uddh Söderberg et al. / Science of the Total Environment 668 (2019) 1064–1076

Fig. 4. Relationship between average metal concentration (μg/l) in ultra-filtered (10 kD) lysimeter-collected pore water and in eluates from batch leaching tests (with H2O and CaCl2).
Ultra-filtered samples should provide the most relevant basis for comparison since batch leaching can result in a high release of colloids.

times those received after batch leaching of the soil removed upon the in batch leaching set-ups. The results from our study, however,
installation of these tubes. imply the opposite net effect despite a confirmed release of colloids.
Since the same difference applied to both ultra-filtered (b10 kD) and While Meers et al. (2007) presented a good correlation between Cd
membrane filtered (b0.45 μm) samples, the higher concentration in ly- concentrations in pore water collected with Rhizon soil moisture sam-
simeter pore waters cannot be the effect of a higher occurrence of col- plers and in eluates after CaCl2 leaching, underestimated metal concen-
loid particles in these waters. Moreover, the metal association to trations upon batch leaching have been presented before. For example,
colloids was higher after the batch leaching tests – either due to a de Groot et al. (1998) showed that metal concentrations in 0. 01 M CaCl2
lower average grain size of the sieved material, or more probably due extracts at L/S = 10 where consistently lower than in soil pore water,
to the release of fine particles during shaking (Bergendahl, 2005; and most pronouncedly so for metals with a high tendency of organic
Bergendahl and Grasso, 1998). While the colloidal association was vir- complexation – such as Pb and Cu. Since dilution results in lower DOC
tually non-existent for all metals except Pb and Sb in the pore waters concentrations in CaCl2 extracts (Yin et al., 2002), organic complexation
(Fig. 3a), it still corresponded to 91% (Pb), 55% (Cd), 27% (As) and 11% is also reduced. Also Degryse et al. (2009) have suggested that a de-
(Sb) of the total concentrations in membrane-filtered water after crease in DOC complexation can result in lower aqueous phase metal
batch leaching with H2O – supporting the idea of a significant colloidal concentrations after batch leaching. In our dataset this mechanism is
release with this method. For the CaCl2 extraction colloidal associa- supported by a higher average DOC concentration in lysimeter-
tion compared to the total metal concentration were lower (64, 4, collected pore waters (16 mg/l) than in the solutions after CaCl2 and
14 and 4%, respectively), probably due to the aggregating and bridg- H2O leaching (1.7 and 2.4 mg/l, respectively). However, if this was the
ing potential of the Ca2+ ions which can result in the formation of sole mechanism, one would expect the effect to be clearer for Pb than
some particles that are too large to pass through a 0.45 μm filter for, eg, Sb, which showed no DOC complexation in the system (Fig. 3).
(Dahlgren and Marrett, 1991; Grathwohl and Susset, 2009; Klitzke Other pore water characteristics than DOC content that are important
and Lang, 2009; Kretzschmar and Sticher, 1997). Grathwohl and for the solid-liquid partitioning – such as the pH and ionic strength –
Susset (2009) speculated, but without performing any tests to con- can also differ between field and lab conditions and may thus affect
firm their hypothesis, that the mobilization of fine particles during partitioning assessments (Di Bonito et al., 2008; Fest et al., 2008; van
shaking was likely to result in an overestimation of metal leaching der Sloot et al., 1996), but the difference is marginal in our case,
T. Uddh Söderberg et al. / Science of the Total Environment 668 (2019) 1064–1076 1073

where both the natural pore waters and leaching eluates were be irrelevant, but it can of course be that the different pore systems in
circumneutral (pH = 7.1 in the leaching tests and pH = 6.9 in the the waste deposit are not in true equilibrium.
field). Besides differences in the aqueous phase chemistry, alternative
explanations for the lower leachability assessed from batch leaching 3.6. Substantial intra-site variability in soil:solution partitioning – Kd
may be a relatively short contact time or a higher L/S ratio (Di Bonito values
et al., 2008; Ndiba and Axe, 2010; Schuwirth and Hofmann, 2005); but
a recent study by Löv et al. (2019) implies that the contact time in stan- Fig. 5 presents ranges in partitioning coefficients from a selection of
dardized leaching tests with both CaCl2 and H2O, and at different L/S, is previously published articles and reports, together with Kds from
enough for near-equilibrium conditions to arise. Another contributing Pukeberg. According to several of the earlier studies, which have fo-
factor may be the removal of coarse grain sizes by standard procedure cused on Kd variability between different sites, Kd values can be fairly
pre-treatment (sieving) of the laboratory samples. Under true equilib- well predicted from soil and/or pore water characteristics that are
rium conditions, the particle size distribution of the solid phase should well known to influence metal partitioning. pH is typically the strongest

Fig. 5. Variability in Log10 Kd values from Pukeberg (ref 1–3), in relation to a selection of other studies which have focused on compiling Kd data from multiple sites (ref 4–8). The figures
show min-max intervals (bars), with the red asterisks indicating the arithmetic mean values. Most Kd values are based on pseudo-total concentrations of both the solid and aqueous phase.
In the ideal case, free-ion concentrations of the pore water and geochemically active concentrations (representing the sorbed pool) of the solid phase are preferred, since these are the ones
participating in the equilibrium reactions relevant for the Kd coefficient. The improvement in prediction when using the latter concentrations instead of total ones is, however, small
(Degryse et al., 2009).
1074 T. Uddh Söderberg et al. / Science of the Total Environment 668 (2019) 1064–1076

Table 4
Linear mixed-effects models of soil:solution partitioning coefficients (Kd) for As Cd, Pb and Sb as explained by the following predictor variables: soil pseudo-total metal concentration, pH,
organic C and pseudo-total Fe content. P-values for all predictors are presented, as well as parameter estimates from the final models which only contained statistically significant predictor
variables.

Metal Predictor variables Final regression model Final model marginal R2

log [Me] pH log org C log Fe

As 0.4517 0.0429* 0.8588 0.2091 logAsKd = 3.7 + 0.68 × pH 7%


Cd 0.0643 0.1428 0.3947 0.5312 – –
Pb 0.0013** 0.5864 0.5482 0.1797 logPbKd = −2.0 + 1.74 × logPbSoil 87%
Sb 0.2820 0.0076** 0.6833 0.2011 logPbKd = 8.7–0.46 × pH 9%

*, ** represents p b 0.05 and p b 0.01, respectively.

determinant, together with organic matter, Fe, and/or the solid phase i) The long lasting (N100 years) deposition of metal-containing
concentration of the metal in question (Carlon et al., 2004; de Vries glass waste to the investigated landfill site is not only reflected
et al., 2011; Degryse et al., 2009; Janssen et al., 1997; Sauvé et al., by high total concentrations of As, Cd, Pb and Sb, but also by a
2000). Under ideal conditions, the Kd value should be insensitive to high fraction of geochemically active forms and elevated concen-
changes in the solid phase metal concentration (since it is equivalent trations in pore water. However, despite their leachability in the
to the numerator of the Kd expression). In practice, however, metal waste zone, soil metal concentrations decrease abruptly to back-
sorption is often not linear – and regression model performance may ground concentrations (within decimeters) as the landfilled ma-
therefore be improved by including the solid phase metal concentration terial transcends into the natural soil underneath. The main
(Sauvé et al., 2000). controlling factor is most probably the high pH of the waste ma-
From Fig. 5 we see that the Kd values calculated for the Pukeberg terial, which has the strongest effect on the cationic metals Cd
waste zone are highly variable, even though they were calculated for a and Pb.
single site that has received waste materials from only one specific ii) Field leachability may not be adequately assessed when
type of industry, and with little variation across the site in basic geo- translated from standard batch leaching tests. In this study we
chemical properties. For most metals the variability is comparable to found consistently higher concentrations of Cd, Pb and Sb in
that found in studies which have included many different sites, with ev- lysimeter-collected pore water than in eluates from the evalu-
ident differences in terms of pollution load and general geochemistry. ated laboratory batch leaching experiments.
The highest Kd variability in the Pukeberg waste zone was seen for Pb, iii) The soil:solution partitioning (Kd values) varied by several
ranging from 140 to 2,900,000 l kg−1 (field Kd; ref. 1 in Fig. 5). The orders of magnitude within the landfill waste zone; basically
variability then decreased in the order As N Cd N Sb, with interval ranges within the same range as previously reported for different sites,
of [450–98,000], [22–4700] and [62–1300], respectively. Our linear with marked differences in basic soil geochemistry. In our
models further imply that traditional soil attributes play a relatively study, however, basic soil properties vary little, and our regres-
low role in the soil:solution partitioning in the system (Table 4). Even sion models imply that traditional soil attributes explain little
though there was a significant relationship between soil pH and the of the variability observed in soil:solution partitioning. Instead,
Kd values of As in Pukeberg, the model R2 values (5–9%) were low com- we propose that waste material heterogeneity is more important
pared to previous studies considering data from multiple sites. For ex- in this case, where the contaminated waste zone is characterized
ample, Sauvé et al. (2000) (ref 8 in Fig. 5) found R2 values in the range by a high fraction of waste (or non-soil material). The generaliz-
of 29–58% when using only pH as the predictor variable. Combining ability of a “waste heterogeneity effect” for metal partitioning,
pH with log SOM and/or log [Me soil], R2 increased to 56–88% (Sauvé however, needs to be further evaluated.
et al., 2000). Similar or even higher R2 values, after considering pH iv) Since risk assessment protocols and remediation strategies at
and log OC, are presented in the review by Degryse et al. (2009), pre- contaminated sites focus to a great degree on the conditions for
senting data from multiple studies. Janssen et al. (1997), too, presented contaminant spreading, efforts to pinpoint critical flaws or un-
regression models of good prediction ability, with R2 values from 71 to necessarily uncertain assumptions are of high practical rele-
85% for a range of metals. vance. In this context, both the potentially high metal retention
The relatively poor correlation between field Kds and traditional and the intra-site variability in Kd values have direct implica-
predictor variables (soil pH, org-C and Fe content; Table 4) in Pukeberg tions. For example, the measured concentrations in groundwater
is probably due to the high variability in Kd combined with a relatively are much lower than one would predict by applying the national
low variability in the predictor variables. Exemplifying this by pH, the generic risk assessment model (Swedish, 2009). The difference
Pukeberg waste zone sample of lowest pH (6.8) had a hydrogen ion ac- was most evident for Cd and Pb, where simulated average con-
tivity only 6.3 times the sample with the highest pH (7.6). By compari- centrations in groundwater are 3800 and 57,000 times those actu-
son, Janssen et al. (1997) presented a pH range of 3.8–7.9 for the 20 ally found in the field - when applying conservative (minimum)
Dutch contaminated soils for which they estimated Kd values, corre- site-specific Kd values. Applying average Kd values in the model
sponding to a 12,500 times difference in hydrogen ion activity. And in instead, however, the difference was reduced to 54 and 30
the study by Groenenberg et al. (2010) there was a 39,800 times differ- times. This shows the significance of choosing a representative es-
ence in hydrogen ion activity (pH range 3.7–8.3) when comparing 216 timate for soil:solution partitioning, as well as the need to im-
different soils, representing both clean and heavily contaminated soils prove our understanding about which factors that control metal
from the Netherlands and the UK. Therefore, the data at hand for partitioning. In this study, we have touched upon possible reten-
Pukeberg indicates that the waste material itself has a more prominent tion mechanisms, but further analyzes are needed.
influence on metal partitioning than the basic soil properties generally
attributed to this role.

4. Conclusion CRediT authorship contribution statement

The most important conclusions that can be drawn from this high- Uddh Söderberg: T.Conceptualization, Methodology, Validation, Formal
resolution field and laboratory study are that: analysis, Investigation, Writing - original draft, Visualization. Berggren
T. Uddh Söderberg et al. / Science of the Total Environment 668 (2019) 1064–1076 1075

Kleja: D.Conceptualization, Methodology, Validation, Formal analysis, In- de Groot, A.C., Peijnenburg, W.J.G.M., van den Hoop, M.A.G.T., Ritsema, R., van Veen,
R.P.M., 1998. Heavy metals in Dutch Field soils: An Experimental and Theoretical
vestigation, Writing - original draft, Visualization. Åström: M.Conceptu- Study on Equilibrium Partitioning. Report No 607220 001. National Institute of Public
alization, Investigation, Writing - review & editing, Funding acquisition. Health and the environment (RIVM), Bilthoven, The Netherlands.
Jarsjö: J.Conceptualization, Investigation, Writing - review & editing, de Matos, A.T., Fontes, M.P.F., da Costa, L.M., Martinez, M.A., 2001. Mobility of heavy
metals as related to soil chemical and mineralogical characteristics of Brazilian
Funding acquisition. Fröberg: M.Investigation, Writing - review & soils. Environ. Pollut. 111, 429–435.
editing, Funding acquisition. Svensson: A.Formal analysis, Writing - re- de Souza Braz, A.M.D., Fernandes, A.R., Ferreira, J.R., Alleoni, L.R.F., 2013. Distribution coef-
view & editing. Augustsson: A.Conceptualization, Methodology, Valida- ficients of potentially toxic elements in soils from the eastern Amazon. Environ. Sci.
Pollut. Res. 20, 7231–7242.
tion, Formal analysis, Investigation, Writing - original draft, Supervision, de Vries, W., McLaughlin, M.J., Groenenberg, J.E., 2011. Transfer functions for solid-
Project administration, Funding acquisition. solution partitioning of cadmium for Australian soils. Environ. Pollut. 159,
3583–3594.
Degryse, F., Smolders, E., Parker, D.R., 2009. Partitioning of metals (Cd, Co, Cu, Ni, Pb, Zn)
Acknowledgement in soils: concepts, methodologies, prediction and applications - a review. Eur. J. Soil
Sci. 60, 590–612.
Di Bonito, M., Breward, N., Crout, N., Smith, B., Young, S., 2008. Overview of selected soil
Commercial laboratory Actlabs in Canada, Forest Research labora- pore water extraction methods for the determination of potentially toxic elements
tory in England and ALS Scandinavia AB in Sweden are acknowledged in contaminated soils: operational and technical aspects. Environmental Geochemis-
try, pp. 213–249.
for excellent analyzes on soil and water samples. We thank Magnus Dingham, S.L., 2002. Physical Hydrology. Waveland Press Inc, Long Grove.
Törnqvist, a former Bachelor student in hydrology at Stockholm Univer- Disfani, M.M., Arulrajah, A., Bo, M.W., Sivakugan, N., 2012. Environmental risks of using
sity, for assistance in establishing the delineation of the local catchment recycled crushed glass in road applications. J. Clean. Prod. 20, 170–179.
Elert, M., Höglund, L.O., 2012. Huvudstudie Pukebergs Glasbruk. Report No Kemakta AR
area, and Gitte Laallam, a former Master student in hydrology at
2012-07. Kemakta Konsult AB, Stockholm.
Stockholm University, for assisting in the establishment of the hydraulic Fanger, G., Höglund, L.O., Jones, C., Svensson, H., 2003. Undersökning och fördjupad
conductivity at the landfill, and finally Malin Ekstedt (Environmental riskbedömning av fem glasbruk i Kalmar och Kronobergs län samt förslag på generell
metodik för riskbedömningar vid glasbruk. Report No Kemakta AR 2003-07. Kemakta
Office in Nybro Municipality) and David Lokranz (the County Adminis-
Konsult AB, Stockholm.
trative Board) for their engagement in the project. Fanger, G., Elert, M., Höglund, L.O., Jones, C., 2006. Laktester för riskbedömning av
Funding: Financial support for this study was provided by the Swed- förorenade områden - underlagsrapport 3. Report No 5558. Swedish EPA, Stockholm.
ish Geological Survey (grant number 36-1778/2014), the Faculty of Fest, E., Temminghoff, E.J.M., Comans, R.A.J., van Riemsdijk, W.H., 2008. Partitioning of or-
ganic matter and heavy metals in a sandy soil: effects of extracting solution, solid to
Health and Life Sciences at Linnaeus University and the Swedish Geo- liquid ratio and pH. Geoderma 146, 66–74.
technical Institute. Fröberg, M., Berggren, D., Bergkvist, B., Bryant, C., Mulder, J., 2006. Concentration and
fluxes of dissolved organic carbon (DOC) in three Norway spruce stands along a cli-
matic gradient in Sweden. Biogeochemistry 77, 1–23.
Appendix A. Supplementary data Goldberg, S., 2002. Competitive adsorption of arsenate and arsenite on oxides and clay
minerals. Soil Sci. Soc. Am. J. 66, 413–421.
Supplementary data to this article can be found online at https://doi. Grathwohl, P., Susset, B., 2009. Comparison of percolation to batch and sequential
leaching tests: theory and data. Waste Manag. 29, 2681–2688.
org/10.1016/j.scitotenv.2019.03.013. Groenenberg, J.E., Romkens, P., Comans, R.N.J., Luster, J., Pampura, T., Shotbolt, L., et al.,
2010. Transfer functions for solid-solution partitioning of cadmium, copper, nickel,
lead and zinc in soils: derivation of relationships for free metal ion activities and val-
References idation with independent data. Eur. J. Soil Sci. 61, 58–73.
Gustafsson, J.P., 2001. Modeling the acid-base properties and metal complexation of humic
Ahmed, A.M., Sulaiman, W.N., 2001. Evaluation of groundwater and soil pollution in a substances with the Stockholm Humic Model. J. Colloid Interface Sci. 244, 102–112.
landfill area using electrical resistivity imaging survey. Environ. Manag. 28, 655–663. Gustafsson, J.P., 2016. Visual MINTEQ Version 3.1. http://vminteq.lwr.kth.se/download/.
Allison, J.D., Allison, T.L., 2005. Partition Coefficients for Metals in Surface Water, Soil, and Gustafsson, J.P., Berggren Kleja, D., 2005. Modeling salt-dependent proton binding by or-
Waste. Report No EPA/600/R-05/074. U.S. Environmental Protection Agency Office of ganic soils with the NICA-Donnan and Stockholm humic models. Environ. Sci.
Research and Development, Washington, DC. Technol. 39, 5372–5377.
Alloway, B.J., 2013. Trace Metals and Metalloids in Soils and Their Bioavailability. Springer, Gustafsson, J.P., Pechova, P., Berggren, D., 2003. Modeling metal binding to soils: the role
Dordrecht, Netherlands. of natural organic matter. Environ. Sci. Technol. 37, 2767–2774.
Bajda, T., 2010. Solubility of mimetite Pb-5(AsO4)(3)Cl at 5-55 degrees C. Environ. Chem. Hermelin, C., Welander, E., 1986. Glasboken – Historia, teknik och form (in Swedish).
7, 268–278. Centraltryckeriet AB, Borås.
Barton, K., 2013. MuMIn: Multi-model Inference. R Package Version 1.9.13. Imteaz, M.A., Ali, M.M.Y., Arulrajah, A., 2012. Possible environmental impacts of recycled
Baumann, T., Fruhstorfer, P., Klein, T., Niessner, R., 2006. Colloid and heavy metal trans- glass used as a pavement base material. Waste Manag. Res. 30, 917–921.
port at landfill sites in direct contact with groundwater. Water Res. 40, 2776–2786. Inegbenebor, A.I., Thomas, J.H., Williams, P.A., 1989. The chemical-stability of mimetite
Bergendahl, J., 2005. Batch leaching tests: colloid release and PAH leachability. Soil Sedi- and distribution coefficients for pyromorphite mimetite solid-solutions. Mineral.
ment Contam. 14, 527–543. Mag. 53, 363–371.
Bergendahl, J., Grasso, D., 1998. Colloid generation during batch leaching tests: mechanics Jalali, M., Najafi, S., 2018. Effect of pH on potentially toxic trace elements (Cd, Cu, Ni, and
of disaggregation. Colloids Surf. A Physicochem. Eng. Asp. 135, 193–205. Zn) solubility in two native and spiked calcareous soils: experimental and modeling.
Bolan, N., Kunhikrishnan, A., Thangarajan, R., Kumpiene, J., Park, J., Makino, T., et al., 2014. Commun. Soil Sci. Plant Anal. 49, 814–827.
Remediation of heavy metal(loid)s contaminated soils–to mobilize or to immobilize? Janssen, R.P.T., Peijnenburg, W., Posthuma, L., VandenHoop, M., 1997. Equilibrium
J. Hazard. Mater. 266, 141–166. partitioning of heavy metals in Dutch field soils. 1. Relationship between metal par-
Bouwer, H., 1989. The Bouwer and Rice Slug Test - an update. Ground Water 27, 304–309. tition coefficients and soil characteristics. Environ. Toxicol. Chem. 16, 2470–2478.
Brand, E., Otte, P.F., JPA, Lijzen, 2007. CSOIL 2000: An Exposure Model for Human Risk As- Jarsjö, J., Chalov, S.R., Pietroń, J., Alekseenko, A.V., Thorslund, J., 2017. Patterns of soil con-
sessment of Soil Contamination. A Model Description. RIVM Report 711701054/2007. tamination, erosion and river loading of metals in a gold mining region of northern
National Institute for Public Health and the Environment (RIVM), Bilthoven, NL. Mongolia. Reg. Environ. Chang. 17, 1991–2005.
Buffle, J., Leppard, G.G., 1995. Characterization of aquatic colloids and macromdecules. 2. Jensen, D.L., Holm, P.E., Christensen, T.H., 2000. Soil and groundwater contamination with
-key role of physical structures on analytical results. Environ. Sci. Technol. 29, heavy metals at two scrap iron and metal recycling facilities. Waste Manag. Res. 18,
2176–2184. 52–63.
Carlon, C., Dalla Valle, M., Marcomini, A., 2004. Regression models to predict water–soil Johnson, C., Moench, H., Wersin, P., Kugler, P., Wenger, C., 2005. Solubility of antimony and
heavy metals partition coefficients in risk assessment studies. Environ. Pollut. 127, other elements in samples taken from shooting ranges. J. Environ. Qual. 34, 248–254.
109–115. Kaiser, K., Kalbitz, K., 2012. Cycling downwards - dissolved organic matter in soils. Soil
Christensen, J.B., Jensen, D.L., Christensen, T.H., 1996. Effect of dissolved organic carbon on Biol. Biochem. 52, 29–32.
the mobility of cadmium, nickel and zinc in leachate polluted groundwater. Water Karathanasis, A.D., 1999. Subsurface migration of copper and zinc mediated by soil col-
Res. 30, 3037–3049. loids. Soil Sci. Soc. Am. J. 63, 830–838.
Christensen, T.H., Kjeldsen, P., Bjerg, P.L., Jensen, D.L., Christensen, J.B., Baun, A., et al., Khalil, M.H., 2012. Magnetic, geo-electric, and groundwater and soil quality analysis over
2001. Biogeochemistry of landfill leachate plumes. Appl. Geochem. 16, 659–718. a landfill from a lead smelter, Cairo, Egypt. J. Appl. Geophys. 86, 146–159.
Clausen, J.L., Bostick, B., Korte, N., 2011. Migration of lead in surface water, pore water, and Kiddee, P., Naidu, R., Wong, M.H., Hearn, L., Muller, J.F., 2014. Field investigation of the
groundwater with a focus on firing ranges. Crit. Rev. Environ. Sci. Technol. 41, quality of fresh and aged leachates from selected landfills receiving e-waste in an
1397–1448. arid climate. Waste Manag. 34, 2292–2304.
Crawley, M.J., 2007. The R Book. John Wiley & Sons Ltd, Chichester. Kjeldsen, P., Barlaz, M.A., Rooker, A.P., Baun, A., Ledin, A., Christensen, T.H., 2002. Present
Dahlgren, R.A., Marrett, D.J., 1991. Organic-carbon sorption in Arctic and sub-alpine and long-term composition of MSW landfill leachate: a review. Crit. Rev. Environ. Sci.
Spodosol-B horizons. Soil Sci. Soc. Am. J. 55, 1382–1390. Technol. 32, 297–336.
1076 T. Uddh Söderberg et al. / Science of the Total Environment 668 (2019) 1064–1076

Klitzke, S., Lang, F., 2009. Mobilization of soluble and dispersible lead, arsenic, and anti- Sjöstedt, C., Löv, A., Olivecrona, Z., Boye, K., Kleja, D.B., 2018. Improved geochemical
mony in a polluted, organic-rich soil - effects of pH increase and counterion valency. modeling of lead solubility in contaminated soils by considering colloidal fractions
J. Environ. Qual. 38, 933–939. and solid phase EXAFS speciation. Appl. Geochem. 92, 110–120.
Kretzschmar, R., Sticher, H., 1997. Transport of humic-coated iron oxide colloids in a SKB, 2016. Near Surface Groudwater Data in Simpevarp Retreived From the Site Charac-
sandy soil: influence of Ca2+ and trace metals. Environ. Sci. Technol. 31, 3497–3504. terization Database (SICADA). Swedish Nuclear Fuel and Waste Management (SKB).
Kumari, S., Singh, A.K., Verma, A.K., Yaduvanshi, N.P.S., 2014. Assessment and spatial dis- Smith, R.M., Martell, A.E., Motekaitis, R.J., 2003. NIST critically selected stability constants
tribution of groundwater quality in industrial areas of Ghaziabad, India. Environ. of metal complexes database. NIST Standard Reference Database 46, Editor. 7.0. Na-
Monit. Assess. 186, 501–514. tional Institute of Standards and Technology, U.S. Department of Commerce, Gai-
Lee, J.S., Nriagu, J.O., 2007. Stability constants for metal arsenates. Environ. Chem. 4, thersburg, MD.
123–133. Stone, A.L., Mitchell, F., Van de Poll, R., Rendall, N., 2015. APPENDIX G - Interpreting Topo-
Leuz, A.K., Monch, H., Johnson, C.A., 2006. Sorption of Sb(III) and Sb(V) to goethite: influ- graphic Maps and Drawing Watershed Boundaries. Method for Inventorying and
ence on Sb(III) oxidation and mobilization. Environ. Sci. Technol. 40, 7277–7282. Evaluating Freshwater Wetlands in New Hampshire (NH Method). University of
Lindsay, W.L., 1979. Chemical Equilibria in Soil. John Wiley & Sons, New York. New Hampshire, U.S. Department of Agriculture and N.H. counties cooperation.
Löv, A., Larsbo, M., Sjöstedt, C., Cornelis, G., Gustafsson, J.P., Kleja, D.B., 2019. Evaluating Sungur, A., Soylak, M., Ozcan, H., 2015. Investigation of heavy metal mobility and avail-
the ability of standardised leaching tests to predict metal(loid) leaching from intact ability by the BCR sequential extraction procedure: relationship between soil proper-
soil columns using size-based elemental fractionation. Chemosphere 222, 453–460. ties and heavy metals availability. Chem. Spec. Bioavailab. 26, 219–230.
Magalhaes, M.C.F., Dejesus, J.D.P., 1988. The chemistry of formation of some secondary ar- Swedish, E.P.A., 2009. Riktvärden för förorenad mark: Modellbeskrivning och vägledning.
senate minerals of Cu(II), Zn(II) and Pb(II). Mineral. Mag. 52, 679–690. Report No 5976. Swedish EPA, Stockholm.
Magnusson, J., 1971. Glas – Glasets egenskaper och tillverkning. Svenska Reproduktions Tahervand, S., Jalali, M., 2016. Sorption, desorption, and speciation of Cd, Ni, and Fe by
AB, Stockholm. four calcareous soils as affected by pH. Environ. Monit. Assess. 188, 322.
Malmstrom, M.E., Berglund, S., Jarsjo, J., 2008. Combined effects of spatially variable flow Törnqvist, M., 2015. Tungmetallers infiltration och utspädning i grundvatten - En
and mineralogy on the attenuation of acid mine drainage in groundwater. Appl. fallstudie om Pukebergs Glasbruk. Master. Stockholms University, Stockholm.
Geochem. 23, 1419–1436. U.S. EPA, 1991. Risk Assessment Guidance for Superfund. Volume I. Human Health Evalu-
Meers, E., Samson, R., Tack, F.M.G., Ruttens, A., Vandegehuchte, M., Vangronsveld, J., et al., ation Manual (Part B, Development of Risk-based Preliminary Remediation Goals).
2007. Phytoavailability assessment of heavy metals in soils by single extractions and Report No EPA/540/R-92/003. U.S. Environmental Protection Agency, Office of Emer-
accumulation by Phaseolus vulgaris. Environ. Exp. Bot. 60, 385–396. gency and Remedial Response, Washington, D.C.
Mor, S., Ravindra, K., Dahiya, R.P., Chandra, A., 2006. Leachate characterization and assess- U.S. EPA, 1999. Understanding Variation in Partition Coefficient, Kd Values - Volume I: The
ment of groundwater pollution near municipal solid waste landfill site. Environ. Kd Model, Methods of Measurement, and Application of Chemical Reaction Codes.
Monit. Assess. 118, 435–456. Report No EPA 402-R-99-004A. United States Environmental Protection Agency, Of-
Naidu, R., Kookana, R.S., Sumner, M.E., Harter, R.D., Tiller, K.G., 1997. Cadmium sorption fice of Air and Radiation, Washington, DC.
and transport in variable charge soils: a review. J. Environ. Qual. 26. USGS, 2006. Techniques of Water-resources Investigations Book 9. US Geological Survey
Ndiba, P.K., Axe, L., 2010. Risk assessment of metal leaching into groundwater from phos- (USGS).
phate and thermal treated sediments. J. Environ. Eng. 136, 427–434. van der Sloot, H.A., Comans, R.N.J., Hjelmar, O., 1996. Similarities in the leaching behav-
Pandey, M., Tripathi, S., Pandey, A.K., Tripathi, B.D., 2014. Risk assessment of metal species iour of trace contaminants from waste, stabilized waste, construction materials and
in sediments of the river Ganga. Catena 122, 140–149. soils. Sci. Total Environ. 178, 111–126.
Persson, K., Jarsjo, J., Destouni, G., 2011. Diffuse hydrological mass transport through Van Liedekerke, M., Prokop, G., S. R-B, Kibblewhite, M., Louwagie, G., 2014. Progress in the
catchments: scenario analysis of coupled physical and biogeochemical uncertainty ef- management of Contaminated Sites in Europe. Report No EUR 26376 EN. European
fects. Hydrol. Earth Syst. Sci. 15, 3195–3206. Commission Joint Research Centre, Institute for Environment and Sustainability,
Pinheiro, J., B. D., D. S., D. S, R Core Team, 2017. nlme:Linear and Nonlinear Mixed Effects Ispra (VA), Italy.
Models. R package version 3.1-131. van Reeuwijk, L.P., 2002. Procedures for Soil Analyses. Wageningen, Netherlands. .
R Core Team, 2017. R: A Language and Environment for Statistical Computing. R Founda- Violante, A., Cozzolino, V., Perelomov, L., Caporale, A.G., Pigna, M., 2010. Mobility and bio-
tion for Statistical Computing, Vienna, Austria. availability of heavy metals and metalloids in soil environments. J. Soil Sci. Plant Nutr.
Ranville, J.F., Chittleborough, D.J., Shanks, F., Morrison, R.J.S., Harris, T., Doss, F., et al., 1999. 10, 268–292.
Development of sedimentation Fleld-flow fractionation-inductively coupled plasma Wang, S., Mulligan, C.N., 2006. Natural attenuation processes for remediation of arsenic
mass-spectrometry for the characterization of environmental colloids. Anal. Chim. contaminated soils and groundwater. J. Hazard. Mater. 138, 459–470.
Acta 381, 315–329. WHO, 2011. Guidelines for Drinking-water Quality. fourth edition. World Health Organi-
Rijkenberg, M.J., Depree, C.V., 2010. Heavy metal stabilization in contaminated road- zation (WHO), Geneve.
derived sediments. Sci. Total Environ. 408, 1212–1220. Yang, J.K., Barnett, M.O., Jardine, P.M., Brooks, S.C., 2003. Factors controlling the bioacces-
Sauvé, S., Hendershot, W., Allen, H.E., 2000. Solid-solution partitioning of metals in con- sibility of arsenic(V) and lead(II) in soil. Soil Sediment Contam. 12, 165–179.
taminated soils: dependence on pH, total metal burden, and organic matter. Environ. Yin, Y.J., Impellitteri, C.A., You, S.J., Allen, H.E., 2002. The importance of organic matter dis-
Sci. Technol. 34, 1125–1131. tribution and extract soil: solution ratio on the desorption of heavy metals from soils.
Schuwirth, N., Hofmann, T., 2005. Comparability of and alternatives to leaching tests for Sci. Total Environ. 287, 107–119.
the assessment of the emission of inorganic soil contamination (11 pp). J. Soils Sedi- Zang, F., Wang, S., Nan, Z., Ma, J., Wang, Y., Chen, Y., et al., 2017. Influence of pH on the
ments 6, 102–112. release and chemical fractionation of heavy metals in sediment from a suburban
SGU, 2016. Environmental Monitoring of Groundwater. Open Access. The Geological Sur- drainage stream in an arid mine-based oasis. J. Soils Sediments 17, 2524–2536.
vey of Sweden (SGU), Uppsala. Zheng, J., Chen, K.H., Yan, X., Chen, S.J., Hu, G.C., Peng, X.W., et al., 2013. Heavy metals in
SGU, 2017a. Geochemistry Documentation and Dataset Over the Study Area. SGU 31- food, house dust, and water from an e-waste recycling area in South China and the
1382/2017. The Geological Survey of Sweden (SGU), Uppsala. potential risk to human health. Ecotoxicol. Environ. Saf. 96, 205–212.
SGU, 2017b. SGU Mapviewer - Open Geological Information. The Geological Survey of
Sweden (SGU), Uppsala.
Shang, S.B., Li, X.Q., 2011. Study on the co-migration of Pb with soil colloids under rainfall.
Adv. Mater. Res. 183-185, 437–441.

You might also like