Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Journal of Hazardous Materials 363 (2019) 394–400

Contents lists available at ScienceDirect

Journal of Hazardous Materials


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

Mobility and retention of indium and gallium in saturated porous media T


Kerstin Ringering, Yasmine Kouhail, Yinon Yecheskel, Ishai Dror , Brian Berkowitz

Department of Earth and Planetary Sciences, Weizmann Institute of Science, Rehovot, Israel

ARTICLE INFO ABSTRACT

Keywords: Transport of indium and gallium is reported in laboratory column experiments using quartz sand as a model
Technology-critical elements porous medium representative of a groundwater system. With increased use of indium and gallium in recent
Transport years, mainly in the semiconductor industry, concerns arise regarding their environmental effects. The transport
Complexation and retention behavior of these two metals were quantified via batch and column experiments, and numerical
Humic acid
modeling. The effect of natural organic matter on indium and gallium mobility was studied by addition of humic
Modeling
Groundwater
acid (HA). Measured breakthrough curves from column experiments demonstrated different binding capacities
Soil-water between indium and gallium, stronger for indium, with the presence of HA affecting retention dynamics. For
Environmental contamination indium, the binding capacity on quartz decreases significantly in the presence of HA, leading to enhanced
mobility. In contrast, gallium exhibits slightly higher retention and lower mobility in the presence of HA. In all
cases, the binding capacity of gallium to quartz is much weaker than that of indium. These results are consistent
with the assumption that indium and gallium form different types of complexes with organic ligands, with
gallium complexes appearing more stable than indium complexes. Quantitative modeling confirmed that metal
retention is controlled by complex stability.

1. Introduction Increased industrial use often leads to increased environmental


concentrations [3] and development of new potential environmental
Indium and gallium are metals of low abundance in the Earth’s sources such as e-waste [11] and biosolids [12]. For example, Toku-
crust, with typical concentrations of 0.11–0.25 mg/L indium and maru et al. [11] reported an extremely high degree of indium con-
15–19 mg/L gallium [1,2]. These two chemical elements were of only tamination in e-waste recycling soil with an enrichment factor > 40
minor technical relevance in the past. Anthropogenic release was compared to control site soil. Similarly, Poledniok [13] reported sys-
mainly via pollution-intensive processes such as mining and burning of tematic higher concentrations of Ga in soil from an industrial area in
fossil energy carriers in which indium and gallium were present as trace Poland compared to agricultural soil from the same region. With re-
metals [3]. Correspondingly, there has been only moderate interest in spect to groundwater, Chen [4] reported that wells in an industrial area
these elements from the environmental sciences community. in Taiwan had a mean Ga concentration of 19.34 μg/L, compared to
However, over the last two decades, both indium and gallium have 0.05 and 0.02 μg/L in groundwater from two non-industrialized zones,
played a key role in the modern semiconductor industry. They are and concentrations of 9.25 μg/L vs. 0.04 and 0.01 μg/L, respectively,
central to integrated circuits and photoelectric devices in products of for In. These findings raise serious concern, given that gallium and
mass production, such as smartphones, tablet computers, and digital especially indium have negative impacts on living organisms, like re-
cameras [4,5]. The predominant Ga compounds in semiconductors are duction in root length and weight for Arabidopsis thaliana and rice
GaAs and GaN [6,7] while the most common use (> 50%) of indium is [3,14–16]. In addition to natural ecosystems, toxicity to various or-
as indium tin oxide (ITO), a transparent, heat reflective and electrically ganisms was reported [17–20] and human health might also be affected
conductive coating used on screens in electrical equipment [7–9]. Both if indium and/or gallium reach groundwater systems that provide
elements are extremely useful in these applications due to their rela- drinking water [14]. In a different context, gallium has been applied to
tively high electron mobility and charge transfer efficiency [10] thus diagnose tumors in medical tests as gallium citrate complexes [21] and
they are often near irreplaceable. Consequently, there has been a can thus be released from hospital wastewater plants.
sudden and dramatic increase in the demand for indium and gallium in Currently, the environmental speciation and the ecotoxicity of in-
the last decade (Fig. S1). dium and gallium are not well constrained. However, these elements


Corresponding author at: Department of Earth and Planetary Sciences, Weizmann Institute of Science, 234 Herzl St. P.O. Box 26, Rehovot, 7610001, Israel.
E-mail address: ishai.dror@weizmann.ac.il (I. Dror).

https://doi.org/10.1016/j.jhazmat.2018.09.079
Received 31 May 2018; Received in revised form 27 September 2018; Accepted 28 September 2018
Available online 06 October 2018
0304-3894/ © 2018 Elsevier B.V. All rights reserved.
K. Ringering et al. Journal of Hazardous Materials 363 (2019) 394–400

are widely accepted to be predominantly trivalent, and hydrolyze in interact with various metals [35]. All working solutions that are de-
solution (to Ga/In(OH)2+, Ga/In(OH)2+, Ga/In(OH)3° and Ga/In scribed hereafter were prepared from stock solutions of 10 mg/L indium
(OH)4−). Moreover, these elements complex to hard inorganic ligands or gallium, including citrate, and referred to hereafter as indium or
when available, and precipitate readily in pure water [1]; both ele- gallium solution. A molar ratio of 1:4.5 metal to citrate was shown to
ments also complex with organic molecules rich in hydroxyl and car- yield stable complexes in aqueous solution and was thus applied in all
boxy groups [15]. Analogy to other metals suggests that toxicity may be cases. For solutions that contain HA, the above working (stock) solu-
exhibited when concentrations in the environment change abruptly [3]. tions were mixed with 10 mg/L HA.
The transport behavior of indium and gallium in groundwater sys-
tems remains poorly understood; to date it has in essence not been
2.3. Sorption isotherms
reported in the literature. Moreover, although the sorption behavior of
indium and gallium has been studied, to some extent, for technical
Sorption equilibrium experiments were performed to quantify the
sorbents utilized for recovery (e.g., from aluminum and zinc ores
mass of indium and gallium that adsorbs to sand under equilibrium
[22–25]), research on their interaction with typical minerals of soil
conditions. Full details of these experiment protocols, and the results
systems is only in its infancy [10,26].
and interpretation, are described in the Supporting Information (SI).
The objectives of this study were to provide insights into the re-
tention and transport behavior of indium and gallium organic com-
plexes in porous media that are representative of some groundwater 2.4. Column experiments
environments. Trisodium citrate was used a as complexation agent that
stabilizes indium and gallium in solution; it is a naturally prevalent Transport behavior of indium and gallium was studied by vertical
molecule representative of short-chain organic acids present in soil. sand column experiments. The experiments were carried out in dupli-
Study of this system is a crucial step towards fuller understanding of the cates with solution concentrations of 1 mg/L metal and a molar ratio of
environmental fate of these elements. Quartz sand was chosen as a 1:4.5 metal to citrate, or subsequently also including 10 mg/L of HA
simple model porous medium. Sorption equilibrium tests were first (which is considered a common environmental level). In addition, each
performed and interpreted in terms of Langmuir and Freundlich sorp- stock solution contained 500 μg/L bromide (NaBr) which served as a
tion isotherms. Transport behavior in saturated quartz sand columns nonreactive tracer. All relevant figures depict the ratio of inlet con-
was then studied, with examination also of the effect of humic acid centration to measured outlet concentration C/C0 vs. pore volume (PV;
(HA) – representative of natural organic matter – on indium and gal- 1 PV = total volume of fluid in the column).
lium mobility. Because of the hydrolysis of indium(III) and gallium(III) Two polycarbonate columns with lengths of 18.2 cm and 19 cm, and
in a pH range typical of sand and soils, indium(III) and gallium(III)- inner diameter of 3 cm were used. Each packed column underwent a
organic acid complexes are not expected to be significant [1]. However, 24 h saturation phase, which involved saturation with DDW, from the
indium(III) and gallium(III) are known to form relatively strong poly- bottom of the column, via a peristaltic pump to displace air in the pores.
nuclear complexes with α-hydroxyl acids in acidic media [27]. More- The DDW reservoir was then substituted by a reservoir solution with the
over, HAs are known to form complexes with metal ions [28–30] and respective metal complexes, and fraction collection at the column outlet
they can thus influence significantly the binding and transport behavior began. After running the experiment for 90–130 pore volumes, the re-
of metal ions in aqueous environments [30–32]. It is further noted that servoir was switched back again to DDW. For the entire experiment, the
citrate is an extremely soluble low molecular weight organic acid [33], pump delivery rate was 1.1 mL/min.
whereas humic acid has a much higher molecular weight, is far less The pH of each collected sample was measured online with an
soluble (particularly under acidic conditions), and is frequently found electrode placed at the outlet of each column. The pH was stable and
in colloidal material [34]. Hence, their roles in these experiments are comparable within less than 0.25 pH units for all experiments as shown
almost mutually exclusive. Finally, quantitative numerical modeling of in Figs. S3–S5.
the experiments confirmed the factors controlling the retention of in-
dium and gallium.
2.5. Concentration analyses
2. Materials and methods
All samples were analyzed via ICP-MS (Agilent 7700 s) for indium,
gallium and bromide concentrations. Before measurement, each sample
2.1. Materials
was filtered through a 0.45 μm PVDF (polyvinylidene fluoride) mem-
brane and then diluted with nitric acid, resulting in a total acid content
Indium(III) chloride (InCl3 98%), gallium(III) nitrate hydrate an-
of 2%.
hydrous basis (Ga(NO3)3·H2O 99.9%), trisodium citrate dihydrate
(C6H5Na3O7·2H2O), sodium bromide (NaBr ≥99.5%), hydrochloric
acid (HCl ≥37%), nitric acid (HNO3 70%), and humic acid (HA) so- 2.6. Modeling
dium salt were purchased from Sigma-Aldrich (for more information on
the HA, see Section 3 in the SI). Medium size sand (UNIMIN, USA) was The transport of indium and gallium complexes was quantified
used in all experiments (for more information on the sand, see Section 3 using the software package Hydrus 1D [36], which can model transport
in the SI). Cobalt (10 mg/L stock solution) was purchased from and retention of multiple solutes in water-saturated porous media. An
Inorganic Ventures and used as an internal standard in analytical inverse mode was applied to fit sorption parameters to experimental
measurements. All solutions were prepared using double deionized breakthrough curves (BTCs), using a nonlinear least squares optimiza-
water (DDW, 18.2 MΩ cm). tion routine, to evaluate the relevance of various nonequilibrium ki-
netic sorption models. Fits of tracer BTCs to the advection-dispersion
2.2. Stock solution equation, associated with each indium or gallium transport experiment,
were used to obtain the dispersivity. Water content and bulk density
Indium and gallium complexes with added trisodium citrate dihy- were calculated from gravimetric measurements and the geometry of
drate, hereafter referred as citrate, were found to produce stable com- the column. For these simulations, steady state, uniform flow through a
plexes in aqueous solutions. Citrate is a typical component of soil or- saturated column was assumed. The effect of more complicated flow
ganic matter and often used as a model molecule, for small natural behavior, such as mobile and immobile water, was neglected in this
organic ligands, that are prevalent in the subsurface environment and study. Full details of the models are described in the SI.

395
K. Ringering et al. Journal of Hazardous Materials 363 (2019) 394–400

KF = 10.7 mg/kg and n = 0.3 with SSD = 3.4. As for gallium, the
Freundlich model for indium provides a better fit (Fig. S2B).
Following the switch from the injection phase experiments to their
flushing (desorption) phases (marked as a dashed line in Fig. 1; recall
Section 2.4), the indium concentration decreased rapidly, similar to
that of bromide. This behavior implies that of the total mass of indium
that was retained in the column during the retention phase, indium is
not released during flushing of the column with DDW. In comparison, a
long tail over many PVs is observed for the gallium, indicating rela-
tively slow release of gallium during the flushing phase; note that the
experiment ended prior to complete release of all the retained gallium.
Comparison of the indium and gallium BTCs shown in Fig. 1 in-
dicates different transport and retention behaviors under similar con-
ditions. Indium shows a much greater degree of retardation than gal-
lium, which is represented by its higher Rf value. Moreover, the
retention of indium appears to be irreversible. The tailing pattern (with
the onset of flushing) of indium and the inert (nonsorbed) bromide
tracer suggests that no retained indium is desorbed once its injection is
Fig. 1. Breakthrough curves of bromide tracer ( ) compared to indium ( ) and stopped. These findings suggest that the interactions of indium with the
gallium ( ) citrate complexes in sand column experiments. Dashed grey line quartz sand surface are stronger than those of gallium. This result is
represents the beginning of the flushing phase. somewhat surprising, given that relative to indium, gallium has a lower
ionic radius (0.47 Å and 0.62 Å for Ga3+ and In3+, respectively [38])
3. Results and discussion and a stronger ionic potential (ratio of ionic charge to ionic radius). Bi
and Westerhoff [10] demonstrated, for these elements, differences in
3.1. Transport of indium and gallium chemisorption potential with speciation and higher tendency to form
bonds under acidic conditions with positively charged SiO2 surface.
A set of BTCs for the column experiments with indium and gallium A possible explanation for this result is that at pH 6.5, speciation
citrate complexes and bromide tracer are presented in Fig. 1, showing calculations (see Section 4 SI and Fig. S3) indicate that In(III) is present
relative concentration as a function of pore volume. For clarity only one as In(OH)3aq, which can lead to formation of hydrogen bonds with the
BTC is shown here for each case; replicates demonstrating very similar quartz surface sites. Under the same conditions, Ga(III) is present
behavior are shown in the SI. The BTCs display typical sigmoidal pro- mostly as Ga(OH)4− leading to electrostatic repulsion with the quartz
gression during the first pore volumes, increasing until equilibrium sites that are negatively charged. Additionally, citrate complexation
conditions are reached. Note that the initial breakthrough and sub- with each of the elements can alter the retention dynamics. Two pos-
sequent increases in indium and gallium concentrations are much sible, different behaviors of the BTC experiments can thus be expected:
slower than those of the (nonreactive) bromide tracer, which show competition between citrate and quartz sand to complex the metal, and
classical behavior and elutes around 1 PV. It is further noted that for all a ternary system where the metal can be bound to citrate and quartz
cases the first signs of elution of indium and gallium appear at 1 PV; simultaneously. Ivanova et al. [27] found that indium(III) citrate and
indium transport is particularly slow, and reaches equilibrium near C/ indium(III) hydroxo complexes are not as stable as similar gallium(III)
C0 = 0.8, while gallium elutes much faster and reaches equilibrium complexes. Hence, it might be hypothesized that the lower amount of
near C/C0 = 0.95. The drop in all three BTCs (at about 130 PVs) retained gallium, and its lower retardation factor as exhibited in the
marked by the dotted line indicates the end of injection of the metal BTC, are caused by a greater stability of gallium citrate complexes,
solutions. Throughout the experiment, the pH values of the indium and which lead to weak binding of gallium on the sand surface. In contrast,
gallium solutions were similar with plateau at 6.4 ± 0.1, as shown in the weaker indium citrate complexes can be broken up, forming other
Fig. S5. less stable indium species in solution, or In(OH)3s that precipitates on
The retention of indium and gallium in the columns can be char- the sand surface, leading to preferential, irreversible retention of in-
acterized via a retardation factor Rf, which quantifies the extent of in- dium in the sand column. These differences in stabilities of indium(III)
teractions between the respective solute and the sand matrix [37]. This and gallium(III) complexes, and the behavior of the BTCs suggest a
factor was calculated as the ratio of the number of pore volumes re- competition mechanism between the organic ligand and the mineral
quired for the metal to reach half of its equilibrium concentration, as surface. This competition mechanism is supported by Schwab et al.
compared to that of bromide. The retardation factors were Rf = 31.2 [39], who showed that citrate was adsorbed to soil and influence the
and 30.5 for indium, and Rf = 6.0 for gallium, for each of the two adsorption of lead on a soil composed of 35% sand, 42% silt, 23% clay,
replicates of the column experiments, thus indicating good experi- and 15 g/kg of organic carbon. The sorption of lead onto the soil was
mental reproducibility in terms of the column transport parameters. reduced due to the presence of Pb-citrate complexes in solution. A
Note that, as background, measurements of sorption isotherms and competition mechanism between the organic ligand and the soil was
their interpretation, in equilibrium experiments, appear in Section 2 of therefore assumed similarly to the current study. Citrate was also
the SI. Briefly, for gallium, the resulting parameter values for the shown to increase the leaching of iron and aluminum from soils
Langmuir model are total number of binding sites, ST = 8.8 mg/kg and [40,41], thus competing with metals sorbed on soils.
Langmuir adsorption constant, KL = 6.9 L/mg, with a sum of squared
deviations (SSD) of 5.5. Fits with the Freundlich model yielded values 3.2. Indium and gallium transport in the presence of humic acid
of free parameters KF = 9.2 mg/kg and n = 0.4, with SSD = 1.7. Based
on the SSD, the Freundlich model provides a better fit than the Lang- The BTCs in the experiments with indium (Fig. 2) show very dif-
muir model (Fig. S2A); these results suggest that the gallium sorption ferent behavior for transport of indium without and with the presence
on quartz sand may involve multilayer deposition and/or interaction of HA. BTC patterns, in the presence of HA, also indicate different be-
among sorbed species. For indium, the resulting Langmuir model havior for indium as compared to gallium (Figs. 2 and 3). In Fig. 2, the
parameter values are ST = 8.2 mg/kg and KL = 57.0 L/mg with BTC for indium with HA shows significantly higher mobility than for
SSD = 4.0, while the Freundlich model parameter values are indium without HA. The retardation factor for indium with HA

396
K. Ringering et al. Journal of Hazardous Materials 363 (2019) 394–400

because both indium and gallium are part of group 13 in the periodic
table, and located one directly above the other, similar trends in their
transport behavior might be expected. The pronounced difference in
transport behavior and interaction with HA requires further discussion.
Hereafter, the possible explanations for the behavior of each element
are examined.

3.2.1. Effect of humic acid on indium transport


For the experiments with indium, the most striking difference is in
the shape of the BTCs, which changed from a sigmoidal shape in the
absence of HA, to a strictly convex shape in the presence of HA (Fig. 2).
Clearly, the initial uptake of indium is much lower in the presence of
HA, which is also reflected by a lower retardation factor. HA is capable
of binding and stabilizing the indium [43] and is also known to sorb on
quartz surfaces [32,44,45]. The speciation of indium (and gallium) has
received little attention but several studies show that aluminum – in the
same column in the periodic table – forms complexes with HA [46–50].
Therefore, indium and gallium are also expected to form complexes
with HA. It is further noted that from the speciation calculations shown
in Section 4 of the SI, In(III) is predicted to be bound mostly to HA
Fig. 2. Breakthrough curves comparing bromide tracer ( ) and indium in the
presence ( ) and absence ( ) of 10 mg/L humic acid (HA), and fits of the (∼90%) in the inlet solution, as shown in Fig. S4B. Therefore, in the
attachment/detachment (red and blue solid lines) and chemical nonequilibrium presence of HA, indium can be expected to react and form indium-HA
(red and blue dashed lines) models to each curve (For interpretation of the complexes that stabilize the dissolved indium fraction in the aqueous
references to colour in this figure legend, the reader is referred to the web solution. Moreover, in the column experiments, indium complexes must
version of this article). compete with HA molecules for binding sites on the quartz sand sur-
faces. This effective reduction in available binding sites for indium can
cause a slowdown in the sorption kinetics, which would also be in ac-
cord with the observation of a reduced fraction of retained vs. injected
mass during the retention phase (Fig. 2).
If the binding affinity of HA to the sand is weaker than that of the
indium complexes the binding sites that are blocked by HA become free
after some time. These binding sites are then occupied by indium. As
this means that the retention of indium effectively proceeds more
slowly than in the absence of humic acid, the equilibrium is reached at
higher pore volume values. The result is a situation wherein the total
amounts of binding sites occupied by indium are eventually similar
under both experimental conditions. The plateau phase of the BTC in-
dicates that either all sorption sites are occupied or that the system has
reached equilibrium. The presence of HA stabilizes the dissolved in-
dium in the solution; as a result, a smaller fraction of dissolved indium
undergoes transformations such as precipitation, which leads to a
higher steady state solution concentration and higher plateau level
(Fig. 2).

3.2.2. Effect of humic acid on gallium transport


In contrast, the presence of HA has a much smaller impact on gal-
Fig. 3. Breakthrough curves comparing bromide tracer ( ) and gallium in the lium mass retention and release, relative to indium. The shape of the
presence ( ) and absence ( ) of 10 mg/L humic acid (HA), and fits of the
BTCs does not change notably with the addition of HA (Fig. 3). How-
attachment/detachment (red and blue solid lines) and chemical nonequilibrium
ever, in the presence of HA, retardation increases slightly and a higher
(red and blue dashed lines) models to each curve (For interpretation of the
references to colour in this figure legend, the reader is referred to the web
fraction of gallium injected mass is retained during the retention phase.
version of this article). The difference between transport of gallium in the presence and ab-
sence of HA in the range between 6 and 12 PVs, where the two curves
split, is statistically and experimentally significant; see Fig. S10 and
(Rf = 18.7 and 19.4) is much lower than without (Fig. 2; Rf = 31.2 and
Section 6 in the SI. However, when considering the overall behavior of
30.5). In contrast, the BTCs for gallium in the presence of HA are quite
gallium and the environmental implications, the effect of HA in the case
similar to those for gallium without addition of HA (see Fig. 3). How-
of gallium is minimal. Note also that the retardation factors in this case
ever, comparison of the Rf values indicates greater retardation in ex-
are calculated (by definition) exactly in the zone of the largest gap
periments with HA (Rf = 8.1, 7.0) than in experiments without it (Rf =
between the curves, so that Rf emphasizes the effect of HA which for
6.0). This finding is somewhat surprising, given that the presence of HA
most parts of the BTC is not significant. Much of this retained mass is
generally increases metal mobility and reduces sorption [e.g., [42]]. It
subsequently released during the flushing phase, regardless of the
is noted that throughout the experiment, the pH of indium with and
presence of HA. Clearly, the mechanisms discussed above for indium
without addition of humic acid was similar, with plateau at ∼6.5 as
retention onto quartz, regarding the role of HA, do not apply for gal-
shown in Fig. S4.
lium. For this case too, throughout the experiment, the pH values of
The adsorption of HA onto quartz sand and/or the competition
gallium with and without addition of humic acid overlap almost com-
between HA and citrate, both acting as ligands, are expected to alter the
pletely, with a plateau at ∼6.4 as shown in Fig. S5.
retention dynamics of both indium and gallium onto quartz. However,
The retention of gallium, both in the presence and in the absence of

397
K. Ringering et al. Journal of Hazardous Materials 363 (2019) 394–400

HA, is reversible, implying a weaker binding between gallium com- the equilibrium sorbed concentration, were fixed separately throughout
plexes and the sand surface regardless of the exact type of complex each set of indium and gallium transport simulations (see Smax, KF and n
formed by gallium in each case. It was shown previously that gallium in Table S1). Moreover, the detachment rate was set to zero for all in-
forms strong complexes with small organic ligands [27]. Hagvall et al. dium simulations because no desorption was observed.
[51] studied the coordination of gallium(III) in the presence of natural The plateau phase of the BTC represents an equilibrium or steady-
organic matter (Suwannee River Natural Organic Matter and Suwannee state phase of the retention process. A plateau at C/C0 < 1 is governed
River Fulvic Acid) using EXAFS and infrared spectroscopy. These au- by additional processes represented by (eq. (4), SI), which is de-
thors found that mononuclear complexes between gallium(III) and termined by the rate constant, μ (for both models). The fitted values for
carboxylate sites of natural organic matter are predominant for acidic μ indicate higher inactivation rate of indium than gallium, in the ab-
to neutral media. When the pH increases, gallium(III)-HA mononuclear sence of HA. This leads to a lower plateau level, consistent with the
complexes tend to break up and yield Ga(OH)3s, which can then interact observation shown in Fig. 2. The effect of HA on indium stabilization is
with carbohydrates. At the lowest studied concentration (0.3 mmol reflected by a decrease in the inactivation rate constant. Consequently,
gallium(III)/mol R-COOH), EXAFS showed that the contribution of the an elevated plateau level is established, as shown in Fig. 2 (compare μ
second-shell C atoms became more important. This finding can be in- values for indium in the presence and absence of HA in Table S4). In
terpreted as the formation of gallium(III)-NOM cage-like structures, contrast, the presence of HA had an insignificant effect on gallium
although further study is needed to describe these structures. The case transport behavior. The similar plateau levels in all gallium experi-
of the coordination of gallium in a ligand shell might not allow for ments, regardless of the presence of HA (Figs. 3, S10, and S11), de-
direct interaction with the sorbent, resulting in relatively weak bonding monstrate comparable values of the inactivation rate constants.
[52]. Obviously, the presence of HA does not have an inhibitory effect The fitted attachment rate constants are larger for indium than for
on the retention of gallium onto sand, which can be seen in its slightly gallium, ka-indium > ka-gallium, accounting for the higher retention of in-
higher retardation factor as compared to experiments without HA. dium in the column. Similarly, ka indium > ka indium, HA, indicating the
Gallium, being a strong complexation center, forms complexes not effect of HA on indium retention.
only with citrate, but also with HA [e.g., [53]]. However, Takahashi The chemical nonequilibrium sorption model involves three reten-
et al. [54] reported that the behavior of gallium sorption to kaolinite is tion parameters, k , KF and n (Eqs. (2), (7), SI). The model results re-
largely unaffected by the presence HA. In these experiments, gallium turned k indium < k gallium , which is counter-intuitive because indium
was preferentially sorbed on the mineral surface. Moreover, when was retained in the column to a significantly greater degree than gal-
gallium was sorbed on the mineral surface in a gallium(III)-kaolinite lium. However, in (Eq. (7), SI), the parameter k controls both the
binary system, these authors reported that ∼10% of the gallium(III) sorption and desorption rates; because desorption of indium was neg-
was partitioned to the dissolved phase in the pH range of 5–7, in the ligible and desorption of gallium was substantial, the k value for in-
presence of HA. It is further noted that at the beginning of the experi- dium was minimized. To compensate, the retention parameter, KF, was
ments, the gallium(III) in solution is complexed to the carboxylic sites larger for indium as compared to gallium.
of the citrate. While the gallium citrate complex is strong, a small As discussed above, gallium-citrate complexes are relatively strong
amount of gallium can be released and react with the carboxylic sites of and stable in solution, so that gallium transport is governed mostly by
the HA, which are the major sites involved in the complexation of the direct interaction of gallium-citrate with the sand surface, regard-
gallium(III) in this pH range [51]. It is noted that in the inlet solution, less of the presence of HA. Both models captured the experimental
Ga(III) is predicted to be bound to citrate (∼35%) and to HA (∼63%) measurements with R2 > 0.88, as shown in Fig. 3 and Table S4.
at pH 5.5 (Fig. S2). When the pH increases to 6.5, Ga(III) is present as The quantitative analyses of indium BTCs are shown in Fig. 2. Un-
Ga(OH)4− and Ga-HA (respectively 45.7 and 53.7%) and Ga(OH)3s is like gallium, indium shows larger retention and is affected significantly
predicted to precipitate. In these BTC experiments, gallium can sorb to by the presence of HA. The attachment/detachment model exhibits
the sand surface either in form of a citrate complex and/or as a HA better performance, compared to the chemical nonequilibrium sorption
complex, but this will not cause a substantial change in the retention model, with regard to simulating indium transport. The latter mainly
pattern or magnitude. failed to capture the S shape of the retention phase of the BTC, and the
A comparison of the complexation behavior of indium and gallium, sharp decrease to background (zero elution) concentration level of the
both belonging to group 13 in the periodic table, to the most-studied BTC tail. On the other hand, in the presence of HA, the attachment/
element of the same group - aluminum - provides some interesting in- detachment model predicts a small sharp increase of indium con-
sights. A review of the fate of aluminum in aqueous environments centration at the initial breakthrough, which is not observed in the
highlights that the type of complexation can even differ for the same experiments.
species of metal ion under different conditions [29]. Aluminum forms As discussed in Section 3.1, indium forms relatively weak complexes
complexes with HA either in the form of outer-sphere complexes, or in with citrate, so that the total retention of indium may involve a larger
the form of inner-sphere complexes, depending on the chemical para- extent of metal species other than indium-citrate. In addition, deso-
meters such as the presence of competing ions, the ionic strength and rption of indium, following the retention phase, was not observed.
pH value of solutions. Whether inner- or outer-sphere complexes are Often, for metal species on soil surfaces, rates of desorption are con-
formed can in turn lead to fundamentally different interactions between siderably slower than rates of adsorption [e.g., [55]]. A possible ex-
the complexes and charged surfaces. planation is that the sorbed citrate complex is transformed, enabling
direct interaction of indium or indium hydroxide with the sand surface;
3.3. Modeling such interaction is strong, and practically irreversible for the conditions
of this experiment. This mechanism may be stronger for indium than for
Two kinetic sorption models were applied to simulate the transport gallium, due to the different stability of the metal-citrate complex. In
behavior of indium and gallium: attachment/detachment model (Eq. such a mechanism, the sorption of indium is: (1) limited by the max-
(5), SI) and chemical nonequilibrium sorption (Eq. (7), SI). Simulation imum sorption sites for indium ions (related to the cation exchange
results are shown in Figs. 2, S8 and S9 for indium experiments and capacity of the sand), rather than the equilibrium between sorbed and
Figs. 3, S10 and S11 for gallium experiments. To reduce the number of aqueous indium-citrate complex; and (2) described by independent
free parameters, the maximum number of retention sites, Smax, was sorption and desorption rates (desorption is practically minimized).
assumed to be unaffected by the experimental conditions, and its value Therefore, the attachment/detachment model is preferred here.
was fixed throughout the entire set of attachment/detachment simu- Overall, for indium, when the metal is stabilized, i.e., when HA is
lations. Similarly, the Freundlich parameters, KF and n, which describe added to the system, there is one major complex that is transported

398
K. Ringering et al. Journal of Hazardous Materials 363 (2019) 394–400

through the column. In such a case, both nonequilibrium and attach- Switzer-Foster and Mr. Bill Foster, and from the De Botton Center for
ment/detachment models describe the BTC adequately. When the Marine Science. B.B. holds the Sam Zuckerberg Professorial Chair in
system is less stable and multiple complexes can be formed and mobi- Hydrology.
lized, the attachment/detachment model better describes the transport.
This is due to the independence of the sorption and desorption pro- Appendix A. Supplementary data
cesses in the model. When there are more possibilities that the complex
will react and then be transformed – e.g., when a sorbed complex reacts Supplementary material related to this article can be found, in the
while being sorbed to form different species that show lower desorption online version, at doi:https://doi.org/10.1016/j.jhazmat.2018.09.079.
rates – breaking the link between sorption and desorption can be ad-
vantageous as each phase is independent of interactions of the other References
phase. For gallium, the citrate complex is relatively stable and is not
affected by the presence of HA; thus the system is more stable and both [1] S.A. Wood, I.M. Samson, The aqueous geochemistry of gallium, germanium, indium
models can describe the transport behavior. and scandium, Ore Geol. Rev. 28 (2006) 57–102.
[2] A. Kabata-Pendias, A.B. Mukherjee, Trace Elements from Soil to Human, Springer,
Berlin, 2007.
4. Conclusion [3] S.J.O. White, H.F. Hemond, The anthrobiogeochemical cycle of indium: a review of
the natural and anthropogenic cycling of indium in the environment, Crit. Rev.
Environ. Sci. Technol. 42 (2012) 155–186.
The objectives of this study were to provide insights into the re- [4] H.-W. Chen, Gallium, indium, and arsenic pollution of groundwater from a semi-
tention and transport behavior of indium and gallium organic com- conductor manufacturing area of Taiwan, Bull. Environ. Contam. Toxicol. 77
plexes in porous media that are representative of some groundwater (2006) 289–296.
[5] J.L. Yang, Comparative acute toxicity of gallium(III), antimony(III), indium(III),
environments. The use of both elements has grown almost ex-
cadmium(II), and copper(II) on freshwater swamp shrimp (Macrobrachium nippo-
ponentially since the early 2000s, resulting in their increased, un- nense), Biol. Res. 47 (2014) 1–4.
avoidable, release to the environment. Therefore, better constraints on [6] T. Butcher, T. Brown, Gallium, in: G. Gunn (Ed.), Critical Metals Handbook, 2014,
pp. 150–176 Chap.7.
the potential of indium and gallium to migrate in the subsurface and
[7] A. Cobelo-Garcia, M. Filella, P. Croot, C. Frazzoli, Du Laing G, N. Ospina-Alvarez,
contaminate water reservoirs are needed critically. S. Rauch, P. Salaun, J. Schäfer, S. Zimmermann, COST action TD1407: network on
Here it is shown that indium and gallium demonstrate similar technology-critical elements (NOTICE)-from environmental processes to human
sorption patterns, with the Freundlich isotherm (multilayer, equili- health threats, Environ. Sci. Pollut. Res. 22 (2015) 15188–15194.
[8] A. Ladenberger, A. Demetriades, C. Reimann, M. Birke, M. Sadeghi, J. Uhlback,
brium dependent) yielding somewhat better fits to the observations M. Andersson, E. Jonsson, The GEMAS project team, GEMAS: indium in agricultural
than the Langmuir isotherm (monolayer sorption). However, indium and grazing land soil of Europe - its source and geochemical distribution patterns, J.
shows much stronger adsorption parameters. The transport behaviors of Geochem. Explor. 154 (2015) 61–80.
[9] U. Schwarz‐Schampera, Indium, Critical Metals Handbook, (2014), pp. 204–229.
indium and gallium were found to be very different, however, with [10] X. Bi, P. Westerhoff, Adsorption of III/ ions (In(III), Ga(III) and As(V)) onto SiO2,
gallium being more mobile and less affected by the addition of HA. CeO2 and Al2O3 nanoparticles used in the semiconductor industry, Environ. Sci.
Therefore, indium appears to pose a lower risk of travelling through Nano 3 (2016) 1014–1026.
[11] T. Tokumaru, H. Ozaki, S. Onwona-Agyeman, J. Ofosu-Anim, I. Watanabe,
waterways and entering drinking water, relative to gallium. However, Determination of the extent of trace metals pollution in soils, sediments and human
indium is much more sensitive to the presence of HA in solution; even hair at e-waste recycling site in Ghana, Arch. Environ. Contam. Toxicol. 73 (2017)
levels of 10 mg/L can lead to very large changes in the pattern, rate and 377–390.
[12] F.-S. Zhang, S. Yamasaki, K. Kimura, Waste ashes for use in agricultural production:
amounts of indium transported in porous media. This may indicate the II. Contents of minor and trace metals, Sci. Total Environ. 286 (2002) 111–118.
potential that other components found in the aqueous solution may [13] J. Połedniok, Speciation of scandium and gallium in soil, Chemosphere 73 (2008)
affect the mobility of both elements in the porous medium, due to their 572–579.
[14] C.S. Ivanoff, A.E. Ivanoff, T.L. Hottel, Gallium poisoning: a rare case report, Food
different reactivity, competition or even different speciation. The sig-
Chem. Toxicol. 50 (2012) 212–215.
nificant differences in the behavior of the two metals are explained by [15] H. Chang, S. Wang, K. Yeh, Effect of gallium exposure in Arabidopsis thaliana is
the strength of the complexes they form, and in turn the potential in- similar to aluminum stress, Environ. Sci. Technol. 51 (2017) 1241–1248.
teractions of the complex with other substances in the system (such as [16] C. Syu, P. Chien, C. Huang, P. Jiang, K. Juang, D. Lee, The growth and uptake of Ga
and In of rice (Oryza sative L.) seedlings as affected by Ga and In concentrations in
other ligands or charged surfaces). hydroponic cultures, Ecotoxicol. Environ. Saf. 135 (2017) 32–39.
Modeling of the BTCs of both indium and gallium confirmed the [17] A. Al-Aoukaty, V.D. Appanna, H. Falter, Gallium toxicity and adaptation in
suggested complexation behavior, which explains the major differences Pseudomonas fluorescens, FEMS Microbiol. Lett. 92 (1992) 265–272.
[18] N. Onikura, A. Nakamura, K. Kishi, Acute toxicity of gallium and effects of salinity
observed for the mobility of the two elements and the effect of the on gallium toxicity to brackish and marine organisms, Bull. Environ. Contam.
presence of humic acid in solution on their transport. The modeling also Toxicol. 75 (2005) 356–360.
suggests that when the composition of solution is stable and the spe- [19] C.H. Lim, J.-H. Han, H.-W. Cho, M. Kang, Studies on the toxicity and distribution of
indium compounds according to particle size in Sprague-Dawley rats, Toxicol. Res.
ciation of the complexes remains constant, both nonequilibrium and 30 (2014) 55–63.
attachment/detachment models can describe the transport. When the [20] N. Onikura, A. Nakamura, K. Kishi, Acute toxicity of thallium and indium toward
solution composition in the system is changing and a substance can be brackish-water and marine organisms, J. Fac. Agric. Kyushu Univ. 53 (2008)
467–469.
transformed to different species, decoupling the sorption and deso- [21] C.J. Anderson, M.J. Welch, Radiometal-labeled agents (non-technetium) for diag-
rption processes is advantageous. This study provides initial informa- nostic imaging, Chem. Rev. 99 (1999) 2219–2234.
tion regarding the transport behavior of both elements, but further [22] M. Nakayama, H. Egawa, Recovery of gallium(III) from strongly alkaline media
using a Kelex-100-loaded ion-exchange resin, Ind. Eng. Chem. Res. 36 (1997)
research that considers more reactive porous media, more complex
4365–4368.
solution compositions, and different conditions (including effects of [23] P. Bermejo-Barrera, N. Martinez-Alfonso, A. Bermejo-Barrera, Separation of gallium
acidity and other ligands) should also be considered. and indium from ores matrix by sorption on Amberlite XAD-2 coated with PAN,
Fresenius J. Anal. Chem. 369 (2001) 191–194.
[24] S. Nusen, T. Chairuangsri, Z. Zhu, C.Y. Cheng, Recovery of indium and gallium from
Competing interests statement synthetic leach solution of zinc refinery residues using synergistic solvent extraction
with LIX 63 and Versatic 10 acid, Hydrometallurgy 160 (2016) 137–146.
The authors have no competing interests to declare. [25] Y. Sasaki, N. Matsuo, T. Oshima, Y. Baba, Selective extraction of In(III), Ga(III) and
Zn(II) using a novel extractant with phenylphosphinic acid, Chin. J. Chem. Eng. 24
(2016) 232–236.
Acknowledgements [26] A. Benedicto, C. Degueldre, T. Missana, Gallium sorption on montmorillonite and
illite colloids: experimental study and modelling by ionic exchange and surface
complexation, Appl. Geochem. 40 (2014) 43–50.
B.B. gratefully acknowledges the support of research grants from the [27] V.Y. Ivanova, V.V. Chevela, S.G. Bezryadin, Complex formation of indium (III) with
Crystal Family Foundation, from Rochelle Rubinstein and Ms. Darlene

399
K. Ringering et al. Journal of Hazardous Materials 363 (2019) 394–400

citratic acid in aqueous solutions, Russ. Chem. Bull. Interface Ed. 64 (2015) [43] Z. Guangzhong, D. Dongyun, L. Jianping, Adsorption model of humic acid with
1842–1849. gallium and indium, Chin. J. Rare Met. 23 (1999) 398–400.
[28] G. Eskenazy, Adsorption of gallium on peat and humic acids, Fuel 46 (1967) [44] A. Jada, R. Ait Akbour, J. Douch, Surface charge and adsorption from water onto
187–191. quartz sand of humic acid, Chemosphere 64 (2006) 1287–1295.
[29] G. Ritchie, Soluble aluminium in acidic soils: principles and practicalities, Plant Soil [45] A. Pitois, L.G. Abrahamsen, P.I. Ivanov, N.D. Bryan, Humic acid sorption onto a
171 (1995) 17–27. quartz sand surface: a kinetic study and insight into fractionation, J. Colloid
[30] P. Zhou, H. Yan, B. Gu, Competitive complexation of metal ions with humic sub- Interface Sci. 325 (2008) 93–100.
stances, Chemosphere 58 (2005) 1327–1337. [46] J. Lambert, J. Buddrus, P. Burba, Evaluation of conditional stability constants of
[31] D.B. Kleja, W. Standring, D.H. Oughton, J.P. Gustafsson, K. Fifield, A.R. Fraser, dissolved aluminum/humic substance complexes by means of 27Al nuclear mag-
Assessment of isotopically exchangeable Al in soil materials using 26Al tracer, netic resonance, Fresenius J. Anal. Chem. 351 (1995) 83–87.
Geochim. Cosmochim. Acta 69 (2005) 5263–5277. [47] B.A. Browne, C.T. Driscoll, pH-dependent binding of aluminum by a fulvic acid,
[32] J.C. Joo, C.D. Shackelford, K.F. Reardon, Association of humic acid with metal Environ. Sci. Technol. 27 (1993) 915–922.
(hydr)oxide-coated sands at solid-water interfaces, J. Colloid Interface Sci. 317 [48] N. Clarke, L.G. Danielsson, A. Sparén, Studies of aluminium complexation to humic
(2008) 424–433. and fulvic acids using a method for the determination of quickly reacting alumi-
[33] E. Oburger, D. Leitner, D.L. Jones, K.C. Zygalakis, A. Schnepf, T. Roose, Adsorption nium, Water Air Soil Pollut. 84 (1995) 103–116.
and desorption dynamics of citric acid anions in soil, Eur. J. Soil Sci. 62 (2011) [49] E. Tipping, C.A. Backes, M.A. Hurley, The complexation of protons, aluminium and
733–742. calcium by aquatic humic substances: a model incorporating binding-site hetero-
[34] R.G. McLaren, K.C. Cameron, Soil Science: Sustainable Production and geneity and macroionic effects, Water Res. 22 (1988) 597–611.
Environmental Protection, Oxford University Press Canada, Don Mills, 1996. [50] S.H. Sutheimer, S.E. Cabaniss, Aluminum binding to humic substances determined
[35] A. Gramlich, S. Tandy, E. Frossard, J. Eikenberg, R. Schulin, Availability of zinc and by high performance cation exchange chromatography, Geochim. Cosmochim. Acta
the ligands citrate and histidine to wheat: does uptake of entire complexes play a 61 (1997) 1–9.
role? J. Agric. Food Chem. 61 (2013) 10409–10417. [51] K. Hagvall, P. Persson, T. Karlsson, Spectroscopic characterization of the co-
[36] J. Šimůnek, M.T. van Genuchten, Modeling nonequilibrium flow and transport ordination chemistry and hydrolysis of gallium(III) in the presence of aquatic or-
processes using HYDRUS, Vadose Zone J. 7 (2008) 782–797. ganic matter, Geochim. Cosmochim. Acta 146 (2014) 76–89.
[37] S.K. Widmer, R.F. Spalding, J. Skopp, Nonlinear regression of breakthrough curves [52] K.S. Smith, Metal sorption on mineral surfaces: an overview with examples relating
to obtain retardation factors in a natural gradient field study, J. Environ. Qual. 24 to mineral deposits, in: G.S. Plumlee, M.J. Logsdon, L.F. Filipek (Eds.), The
(1995) 439–444. Environmental Geochemistry of Mineral Deposits: Part A: Processes, Techniques,
[38] R.D. Shannon, Revised effective ionic-radii and systematic studies of interatomic and Health Issues, Rev. Econ. Geol, Society of Economic Geologists, Inc. 6A,
distances in halides and chalcogenides, Acta Crystallogr. Sec. A 32 (1976) 751–767. Littleton, Colorado, 1999, pp. 161–182.
[39] A.P. Schwab, Y. He, M.K. Banks, The influence of organic ligands on the retention of [53] H. Lippold, A. Mansel, H. Kupsch, Influence of trivalent electrolytes on the humic
lead in soil, Chemosphere 61 (2005) 856–866. colloid-borne transport of contaminant metals: competition and flocculation effects,
[40] J. Gerke, Phosphate, aluminium and iron in the soil solution of three different soils J. Contam. Hydrol. 76 (2005) 337–352.
in relation to varying concentrations of citric acid, Zeitschrift für [54] Y. Takahashi, Y. Minai, S. Ambe, Y. Makide, F. Ambe, Comparison of adsorption
Pflanzenernährung und Bodenkunde 155 (1992) 339–343. behavior of multiple inorganic ions on kaolinite and silica in the presence of humic
[41] D.L. Jones, L.V. Kochian, Aluminium-organic acid interactions in acid soils: I. Effect acid using the multitracer technique, Geochim. Cosmochim. Acta 63 (1999)
of root-derived organic acids on the kinetics of Al dissolution, Plant Soil (1996) 815–836.
221–228. [55] H.M. Selim, M.C. Amacher, Sorption and release of heavy metals in soils: nonlinear
[42] M.N. Jones, N.D. Bryan, Colloidal properties of humic substances, Adv. Colloid kinetics, in: H.M. Selim, D.L. Sparks (Eds.), Heavy Metals Release in Soils, Lewis
Interface Sci. 78 (1998) 1–48. Publishers, Boca Raton, 2001, p. 249.

400

You might also like