You are on page 1of 37

Accepted Manuscript

Title: Sorption mechanisms of chromate with coprecipitated


ferrihydrite in aqueous solution

Authors: Abdullah Al Mamun, Masao Morita, Mitsuaki


Matsuoka, Chiharu Tokoro

PII: S0304-3894(17)30224-8
DOI: http://dx.doi.org/doi:10.1016/j.jhazmat.2017.03.058
Reference: HAZMAT 18474

To appear in: Journal of Hazardous Materials

Received date: 3-12-2016


Revised date: 12-3-2017
Accepted date: 25-3-2017

Please cite this article as: Abdullah Al Mamun, Masao Morita, Mitsuaki
Matsuoka, Chiharu Tokoro, Sorption mechanisms of chromate with
coprecipitated ferrihydrite in aqueous solution, Journal of Hazardous
Materialshttp://dx.doi.org/10.1016/j.jhazmat.2017.03.058

This is a PDF file of an unedited manuscript that has been accepted for publication.
As a service to our customers we are providing this early version of the manuscript.
The manuscript will undergo copyediting, typesetting, and review of the resulting proof
before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that
apply to the journal pertain.
Sorption mechanisms of chromate with coprecipitated

ferrihydrite in aqueous solution

Abdullah Al Mamuna, Masao Moritaa, Mitsuaki Matsuokab, Chiharu Tokorob*

a
Graduate School of Creative Science and Engineering, Waseda University, 3-4-1

Okubo, Shinjuku-ku, Tokyo 169-8555, Japan

b
Faculty of Science and Engineering, Waseda University, 3-4-1 Okubo, Shinjuku-ku,

Tokyo 169-8555, Japan

*
Corresponding author: Tel.: +81-3-5286-3320, fax: +81-3-5286-3491.

E-mail addresses:

mamun@toki.waseda.jp (A. Mamun)

masao.swimer@akane.waseda.jp (M. Morita)

m-matsuoka@aoni.waseda.jp (M. Matsuoka)

tokoro@waseda.jp (C. Tokoro)

1
Abbreviations: BET: Brunauer–Emmett–Teller, EXAFS: Extended X-ray absorption fine structure,
ICP-AES: Inductively coupled plasma atomic emission spectroscopy, XRD: X-ray diffraction
XAFS: X-ray absorption fine structure, XANES: X-ray absorption near edge structure
Highlights

 Coprecipitation showed twice the sorption density of simple adsorption

at pH 5.

 Mechanism shift from outer- to inner-sphere surface complexation at

high Cr/Fe

 In coprecipitation the mechanism shift occurs at lower Cr/Fe ratios than

adsorption

 Higher-molar-ratio bidentate binuclear Cr-Fe bonds; yielded ferrihydrite

expansion.

2
ABSTRACT

Hexavalent chromium (Cr(VI)) attracted researchers’ interest for its toxicity, natural

availability and removal difficulty. Nevertheless, its sorption mechanism is not clearly

understood yet. In this work, we elucidated the sorption mechanism of the

co-precipitation of chromates with ferrihydrite through quantitative analysis. The

influence of Cr/Fe molar ratio on sorption was investigated by zeta potential

measurements, X-ray diffraction (XRD) and X-ray adsorption fine-structure analysis

(XAFS). Coprecipitation at pH 5 showed almost twice the sorption density of

adsorption at pH 5. In co-precipitation, a shift of the XRD peak due to inner-sphere

sorption of chromate was observed at Cr/Fe molar ratio 0.5. For adsorption, the same

peak shift was confirmed at Cr/Fe molar ratio of 1. Zeta potential at pH 5 suggested that

the sorption mechanism changed at Cr/Fe molar ratio 0.25 for coprecipitation and at

Cr/Fe molar ratio of 1 for adsorption. Fitting of Cr and Fe K-edge extended X-ray

adsorption fine-structure suggested that ferrihydrite immobilized Cr(VI) via outer

sphere surface complexation for lower Cr/Fe ratios and via inner-sphere surface

complexation for higher molar ratios. At higher molar ratios, bidentate binuclear Cr–Fe

bonds were well established, thus resulting in the expansion of the ferrihydrite structure.

Keywords: Chromate, Ferrihydrite, Coprecipitation, Inner sphere, Extended X-ray

absorption fine structure

3
1. Introduction

Chromium (VI) (Cr(VI)) is well known toxic species to ecosystems and its

disposal to near rivers could result in the alteration of the food chain and in a direct

hazard for human health[1]. In addition, highly concentrated Cr is toxic, mutagenic,

carcinogenic and teratogenic [2]. Due to it’s high mobility, removal of Cr(VI) through

conventional practice is more difficult compared to Cr(III) [3]. For the same reason,

Cr(VI) can be leached rapidly and it can easily reaches groundwater and subsoil [4]

although, the mobility of Cr release is limited in the last stage because of ferric-oxide

coatings on soil and sediments [5].

Worldwide, ninety-five sites have been listed for Cr pollution that has resulted

from conventional tannery practice and almost 16 million people have been exposed

directly and indirectly to the potential threat of Cr. Moreover, 25% of tap water has been

reported to be affected near contaminated sites in Greece [6].

Adsorption is considered to be one of the most attractive methods for Cr(VI)

removal because of its low cost, ease of implementation and high safety [7,8]. The

selection of a perfect adsorbent plays a significant role in wastewater treatment.

Previous systems have some limitations such as slow kinetics and low removal

efficiencies at low Cr(VI) concentrations [9].

Ferrihydrite [(Fe3+)2O3∙0.5H2O] is one of the widespread minerals that occurs

at the earth’s surface and it can form extraterrestrial materials [10,11] due to it’s

nanocrystal and nanoporous structure resulting in a large surface area. Hydrous ferric

oxide (2-line ferrihydrite) is able to adsorb toxic species such as chromate and arsenate.

Usually, metal ions are removed from solutions of low concentration by sorption onto
4
ferrihydrite. On the other hand, they can be released in high concentrations at low pH.

Thus making, ferrihydrite becomes available for reuse [12]. Therefore, ferrihydrite can

play a great role in the removal of chromate from wastewater. But inserted anions

irrespective of mononuclear bidentate edge sharing or binuclear bidentate corner sharing

might not release easily without any external influences, which is disadvantage for reuse

but advantageous environmentally for the highly polluted sites to prevent re-elution.

The interaction between ferrihydrite and chromate can occur in two different

ways known as coprecipitation and adsorption. In simple adsorption, chromate could

adsorb on the surface of previously formed/prepared ferrihydrite. On the other hand, in

coprecipitation, ferrihydrite was allowed to form and precipitate with the presence of

chromate, as a result chromate can directly adsorbed on the fresh ferrihydrite surface.

Unlike simple adsorption, where the competition between chromate and other ions

might hinder the removal performances, in coprecipitation there is a higher selectivity

towards the chromate ion, thereby leading to higher removal performances.

Coprecipitation performances depends on pH, ionic strength, presence of

competing ions and surface conditions [13], thus requiring for a more difficult

optimization of the operating condition. Moreover, despite of the better removal

efficiency and potentiality of applications with different metal such as arsenate, lead and

zinc [14,15] , the coprecipitation sorption mechanism has not been completely clarified.

It is still unclear whether the adsorption occurs via inner-sphere coordination and

whether a monodentate or a multidentate complexation is involved.

Hsia et al. (1993) revealed the inner-sphere coordination of chromate to iron

oxyhydroxide, but they could not provide detailed information on any specific
5
mechanism [16]. Johnston and Chrysochoou (2012) highlighted the monodentate and

bidentate inner-sphere surface complexation at pH ˂ 6, which indicates the occurrence

of a modified inner-sphere sorption [17]. They used an N2-flushed environment, which

might leads to inner-sphere coordination at lower Cr/Fe ratios; however, washing may
-
promote the formation of chemisorbed OH polar ions, which prefer the monodentate

relationship. By extended X-ray absorption fine-structure (EXAFS) fitting and Cr–Fe

second shell fitting at pH 4, Yoshio et al. (2015) found a very small coordination

number (0.4), which suggests mostly outer-sphere complexation with the inclusion of a

small amount of inner-sphere complex [18]. However, the EXAFS fitting was biased as

they used fixed Debye waller factor value while no detail about experimental conditions

were given. Surface complex modeling for ferrihydrite by Dzomback and Morel (1990)

showed the preference of chromate to form inner-sphere complexes on ferrihydrite

[15,19]. Therefore, given the ambiguous and sometimes conflicting results reported by

previous researchers, by this work we wanted to examine more in depth the sorption

mechanism of coprecipitation. Moreover, we provided a comparative analysis between

adsorption and coprecipitation. Since most researchers have tried to identify the pH

characteristics for ferrihydrite sorption [20], and it has been proven that acidic

conditions (pH 5) are good for chromate adsorption. For this purpose, coprecipitation at

pH 5 was investigated by means of zeta potential measurements, XRD and XAFS

analysis. Novelty aspects of the present study includes an identification of the sorption

mechanism with cross checks to compare adsorption and coprecipitation processes and

a more detail EXAFS fitting compared to previous researches.

6
2. Materials and methods

2.1 Standards and reagents

All chemicals were of analytical grade from Wako Pure Chemical Industries,

Japan. Cr(VI) and Fe(III) solutions were prepared from K2CrO4 and Fe(NO3)3∙9H2O,

respectively. KOH and HNO3 were used to adjust pH and ionic strength. All

experiments were under ambient atmosphere at 25°C and fixed ionic strength of 0.05M

in three replicates. Wastewater compositions vary in practice and many ingredients in

wastewater will influence Cr(VI) sorption. In this study, we examined the essential

mechanisms with the presence of only single metal adsorbent and adsorbate.

2.2 Coprecipitation experiments

Coprecipitation of ferrihydrite in the presence of Cr(VI) [13,21–23] was

investigated for different Cr/Fe molar ratio (0.05, 0.1, 0.125, 0.5, 1, 5, 10) by fixing

Cr(VI) concentration at 10 mg dm-3 (0.19 mmol dm-3) and by changing the

concentration of Fe(III) solutions. The pH was maintained at the target value (pH 5 or

pH 7) within ± 0.05 by the addition of 0.1 M HNO3 and KOH throughout the reaction

time. After 1 hour under stirring, the suspensions were filtered by using a 0.1-μm

membrane filter and the filtrates were analyzed by inductively coupled plasma-atomic

emission spectroscopy (ICP-AES) using an SPS-7800 atomic emission spectrometer

(Seiko Instruments Inc., Chiba, Japan).

7
2.3 Adsorption experiments

In adsorption experiment, ferrihydrite (0.1 dm3 solutions) and chromate (0.1

dm3 solutions) solution prepared separately at the desired pH (5 and 7) and ionic

strength (0.05M) by using 0.1 M HNO3 and KOH [13,21–23].and mixed it later for

adsorption experiment. Initially, ferrihydrite and chromate concentration were double so

that when it mixed together, the total volume became 0.2 dm3. Thus the Cr/Fe molar

ratio and initial chromate concentration became same as coprecipitation (0.1 dm3

solutions) experiment. Ferrihydrite suspension were prepared according to Cornell and

Schwertmann [10] procedure, with only differences that the ionic strength was adjusted

for a better comparison with other samples, and solid/liquid separation and drying were

not conducted before the adsorption experiments. The remaining procedure of

solid-liquid separation and the filtrate analyses were the same as coprecipitation

method.

2.4 Sorption Isotherm

Sorption isotherm experiments were performed similarly to the coprecipitation

and adsorption experiment. Wide range of Cr/Fe molar ratios were examined (0, 0.05,

0.1, 0.5, 1, 5 and 10) for better understanding the sorption trend. Sorption density (SD)

represents the removal efficiency of Cr by each Fe ion and it was measured by the

removal amount of Cr by per unit amount of added Fe adsorbent.

𝐶𝑟(𝑉𝐼)𝑖 −𝐶𝑟(𝑉𝐼)𝑟
SD = (1)
𝐹ℎ
8
Where Cr(VI)i is the initial molar concentration of chromate, Cr(VI)r is residual molar

concentration of chromate and Fh is added molar concentration of Fe.

2.5 XRD analysis

Solid samples after filtration were analyzed by XRD (Geiger flex RAD-IX,

Rigaku Corp. Japan). We used ferrihydrite as our reference material and that’s why

initial Fe concentration was maintained at 20 mg dm-3 and the initial Cr(VI)

concentration was varied to the desired Cr/Fe molar ratio. Filter residues that were

obtained from the coprecipitation and adsorption experiments were freeze-dried at

−45°C and 10 Pa for at least 24 h to avoid crystallization or mineralogical

transformation. XRD patterns were obtained using a copper target (Cu–Kα), a crystal

graphite monochromator and a scintillation detector. The X-ray source was operated at

40 kV and 30 mA with step scanning from 2θ values of 2 to 80°, sequential increments

of 0.02° and a scan speed of 2°/min. A crystal sample holder was used and the derived

pattern was not corrected for background diffraction.

2.6 Zeta potential measurements

The zeta potentials and electrophoretic measurement of the precipitates in the

suspensions were analyzed using an electrophoresis light-scattering spectrophotometer

(Zetasizer Nano, Malvern, Worcestershire, UK). For the measurements, the initial Fe

concentration was fixed at 10 mg dm-3 while the Cr concentration was changed to adjust

the desired Cr/Fe molar ratio. Suspensions produced in the coprecipitation and
9
adsorption experiments were dispersed in an ultrasonic bath for 5 min and rapidly, the

suspensions were poured into a capillary cell to measure the electrophoretic movement.

2.7 XAFS analysis

XANES and EXAFS analysis were performed at the beamline BL5S1 in the

Aichi Synchrotron Radiation Center, Aichi, Japan. Samples from the 1-h coprecipitation

experiments were freeze-dried at −45°C and 10 Pa for at least 24 h to avoid

crystallization or mineralogical alteration. For the transmission method, samples were

blended with boron nitride powder and pressed at 20 kN to create XAFS thin

tablet-shaped samples [24].

For the analysis, the ionization chamber was filled with N2 gas (30%) and He

gas (70%) to monitor the incident beam. N2 gas (75%) and Ar (25%) was used for the

transmission beam. The beam was untuned by 30% to avoid higher-order harmonics

generation in the K-edge analysis of Fe and Cr. The transmission method was used for

the Fe K-edge and Cr K-edge analysis because the Fe and Cr concentrations in the

residues were sufficient. Ferrihydrite (Cr/Fe=0) was analyzed as a reference material in

the Fe K-edge.

3. Results and Discussion

3.1 Chromate removal and sorption isotherm

10
Experimental results showed that coprecipitation yielded to a higher removal of

Cr(VI) compared with the adsorption methods (figure 1). This result suggested that

other mechanisms than adsorption might be involved during coprecipitation and

enhance the removal of Cr(VI) by ferrihydrite. The residual Cr(VI) concentration after

coprecipitation complies with Japanese effluent standards (0.5 mg dm-3) at 0.05 of the

initial Cr/Fe molar ratio (3.84 mmol dm-3 of Fe dosage).

As we previously reported [24,25],the removal of Cr(V) at pH 5 was higher

than the one obtained at pH 7. This result can be explained by considering that, more

anions can form surface complexes in an acidic pH versus in hydroxide because of

favorable surface complexation by protons (H+). Two main chromate species include

HCrO4- and CrO42-, and HCrO42- is more dominant at pH 6.5 because of the following

reaction:

HCrO4- ⇋ CrO42- + H+ log K = −6.51 (2)

where K is the equilibrium constant and this value is reported by Ball et al. (1980) [26].

Many researchers have calculated that hydrogen chromate (HCrO4-) is more strongly

adsorbed and is the dominant surface species for iron or aluminum oxide [27], which

explains why more chromate forms surface complexes at pH 5 than pH 7.

By looking at the sorption isotherms (figure 2), it appears clear that sorption

density for coprecipitation at pH 5 was almost double that the one for adsorption at

same pH (1.56 mol Cr/mol Fe versus 0.85 mol Cr/mol Fe). Regardless of different the
11
sorption density, experimental points fitted Brunauer–Emmett–Teller (BET)-like

isotherms instead of saturated monolayer isotherms. Thus suggesting that Cr(VI)

sorption on ferrihydrite is not only simple outer-sphere adsorption, but it may occur by

special surface complexation in inner-sphere [28–34] or multi-layer adsorbate formation

through polymerization [35–37]. The sorption isotherm moved upward very slowly

initially for very lower Cr/Fe molar ratios and then increased significantly for higher

Cr/Fe molar ratio. Specific Cr(VI) removal mechanisms might exist at a higher molar

ratio.

Previous researchers reported that the Cr(VI) sorption isotherm with α-Fe2O3

and α-FeOOH to be of the Langmuir type [38,39]. In both cases, the experimental

condition was simple adsorption using ferrihydrite precipitate material prepared in

advance. In our case, the BET type isotherm can be explained by considering two

factors. In the first place, coprecipitation and adsorption experiments were carried out

with fresh ferrihydrite, which has a great activity towards Cr(VI). Moreover, at lower

Cr/Fe molar ratio, ferrihydrite might link with another octahedral ferrihydrite, thereby

facilitate a greater absorbable surface area and promote adsorption of Cr. The Fe

EXAFS fitting results confirmed that for lower molar ratio, the Fe-Fe correlations in the

ferrihydrite were progressively disrupted as the Cr/Fe ratio had increased. The

differences in the polymeric structure at higher Cr/Fe ratios might be due to strong

chromate bidentate adsorption that affects the surface of ferrihydrite in the presence of
12
substantial chromate. That could hinder Fe-Fe normal crystallization via oxygen, and

inhibit additional Fe-O-Fe polymerization. Strong bidentate Cr-Fe relation could be

augmented by the substantial presence of chromate like arsenate [37]. Moreover,

chromate could polymerize like silicates and be adsorbed on the ferrihydrite surface

[36]. Isolated aqueous ferric species, probably hexaquo-coordinated, link edges to form

small chains with Fe octahedral like arsenate sorption, thereby acting as second Fe pool

to ensure the maximum loading. As a result, higher sorption density at higher molar

ratio determined BET sorption isotherm. Interestingly, in the adsorption at Cr/Fe higher

molar ratio, chromate might not fully successful unzip the Fe-O-Fe crystallization like

coprecipitation [37]. That’s why sorption density was comparatively lower than

coprecipitation. In both cases immobilization is based on adsorption but the sequence is

different. Nevertheless, more Cr(VI) could condense at the ferrihydrite solid/liquid

interface because of the attraction of counter ions by electrostatic action [38,40].

3.2 XRD measurements

As showed in figure 3, in the coprecipitation, at low Cr/Fe molar ratio (0.1), the

XRD pattern is similar to ferrihydrite. By increasing the Cr/Fe molar ratio, a broad peak

was observed and it shifted from 2θ = 34° to a lower angle of 2θ = 23° (figure 3a). The

peak shifting suggested the mineralogical changes and/or mechanistic changes of Cr(VI)

uptake to ferrihydrite.

On the other hand, there were no changes occurred in XRD pattern until Cr/Fe
13
molar ratio 0.5. A little peak rise (2θ = 34°) and very tiny peak shift (2θ = 23°) was

observed at a higher molar ratio of 1 (figure 3b). This changing ratio of Cr/Fe is almost

identical with the increase in sorption density in the BET-type isotherm in figure 2.

3.3 Zeta potential measurement

In the coprecipitation experiment, a break-even point with sharp decrease of

zeta potential was observed for Cr/Fe = 0.25. On the other hand, in the adsorption

experiment, break-even point was observed until Cr/Fe=1. This difference highlights the

occurrence of the two different sorption style [41]. Indeed, a sharp decrease in zeta

potential can be associated to surface complexation at the interface while the slow

decrease is the result of the Cr(VI) uptake into the inner side of ferrihydrite. The zeta

potential decrease for higher sorption density at higher Cr/Fe molar ratio showed same

behavior for adsorption and coprecipitation, thus indicates the same mechanisms.

From the isotherms, XRD patterns and zeta potential measurements, it is

obvious that the sorption mechanism of Cr(VI) to ferrihydrite changes at higher sorption

densities, higher initial Cr/Fe molar ratio. However in the coprecipitation process, the

sorption density increased at a Cr/Fe ratio of ~0.5, the XRD patterns shifted at this ratio

and the of zeta potential swung at a 0.25 Cr/Fe molar ratio. On the other hand, in

adsorption, the sorption density increased at a Cr/Fe ratio of ~1, the XRD patterns

shifted at this value and the slope of zeta potential changed at this value. Therefore, it

can be reasonable remarked that coprecipitation enhances the removal by promoting a

change of the sorption mechanism, which in turn, determines a higher sorption density

14
than in the adsorption process.

3.4 XAFS measurement

Fe K-edge and Cr K-edge spectra were studied with a changing Cr/Fe molar

ratio to evaluate their interdependency. Fe K-edge spectra were compared with

ferrihydrite as reference material. Because XANES spectra did not show any significant

difference in all ratios (Figure S2 in supplementary information), we analyzed EXAFS

spectra to reveal the sorption nature of Cr (VI).

3.4.1 Fe K-edge EXAFS analysis

Fe K-edge EXAFS spectra of Cr(VI)-coprecipitated ferrihydrite and fresh

ferrihydrite did not show any visible difference (figure S3(a) in supplementary

information). In order to highlight the difference, K3-weighted EXAFS fitting was done

using goethite (α-FeOOH) to obtain the initial values of the fitting parameters [42]. The

radial distribution function that is obtained from the fourier-transformed EXAFS fitting

spectra is shown in table 1 and figure 5.

As it can be observed in table 1, Fe–O has an almost octahedral geometry.

However, for Cr/Fe molar ratio of 0.5, the Fe–O, and Fe–Fe interatomic distance

slightly decreased compared to ferrihydrite. Besides by, increasing the Cr/Fe molar ratio,

the interatomic distances gradually increased. This is another notable point regarding

the mechanism shift or modification of the local microenvironment between Cr/Fe ratio

0 to 0.5, and this result matches with previously described zeta potential results in

15
coprecipitation where mechanism shift was suggested for Cr/Fe 0.25.

EXAFS results also suggest that ferrihydrite may contain 20%–30% of

tetrahedral coordinated Fe(III) as it has a Fe–O interatomic distance of less than 1.99 Å

[43]. Although the ionic strength might affect the nature of the crystal structure, the

gradual increase of the interatomic distance between the Fe–O from the Cr/Fe ratio of 0

to a Cr/Fe ratio of 1 suggests that a small amount of tetrahedral Fe(III) may exist in the

ferrihydrite chain. The tetrahedral Fe(III) rearranges to pure octahedral because of the

inhibition of a significant amount of chromate. Nevertheless, like Manceau and

Gates(1997), we also feel that a large amount of tetrahedral Fe(III) in the ferrihydrite

chain could result from drying [44].

Ferrihydrite usually shows two Fe–Fe shell configurations: one is the edge

sharing and the other one is the double-corner sharing. According to the literature

review, the edge sharing of Fe(O, OH)6 octahedra showed an interatomic distance that

started from 2.99 Å to 3.07 Å with a wide range of coordination numbers (1.4 to 3.5)

[44]. Xie et al. (2015) showed that the Fe–Fe distance in the double-corner sharing was

3.12–3.15 Å, whereas most researchers agreed on a range of 3.4–3.5 Å [45]. In our case,

the interatomic distance is slightly smaller (~0.15 Å) than the previous Fe–Fe second

shell distance range of 3.4–3.5 Å. The reason may be ionic strength adjustment to 0.05

M. Some researchers have found that a face-sharing (2.91 Å) and single-corner-sharing

(3.85 Å) shell could yield good fitting results [46]. We did not obtain face-sharing and
16
single-corner-sharing fitting results. Like most researchers, we obtained two Fe–Fe shell

fittings with an interatomic distance of 3.05 ± 0.01 Å (edge sharing) and 3.35 ± 0.003 Å

(double-corner sharing).

The EXAFS fitting results (Table 1) indicate that the Fe–Fe coordination

number showed irregularities or unevenness and that the interatomic distances increased

gradually by increasing the Cr/Fe molar ratio (Figure 7a). This trend suggests that more

chromate was inhibited in the inner sphere of the ferrihydrite [36]. Moreover, the

presence of a high concentration of Cr was favorable for the formation of octahedral

ferrihydrite.

3.4.2 Cr K-edge EXAFS analysis

Cr K-edge EXAFS results (figure S3(b) in supplied as supplementary

information) showed no visible differences by changing the Cr/Fe molar ratio. A radial

distribution function as from Fourier-transformed EXAFS fitting results (Table 2 and

Figure 6) was produced for further evaluation. The crystal structure of Fe2(CrO4)3∙3H2O

by Bonin [47] was used to provide the initial fitting parameters for K3-weighted EXAFS

fitting. Table 2 shows that Cr is coordinated tetrahedrally with an oxygen atom and the

Cr–O distance was ~1.63 Å. These results match with Bonin (1978) as tetrahedral Cr

was attached as Cr–O with a distance that ranged from 1.566 Å to 1.747 Å [47]. The

Cr–Fe interatomic distance slightly increased from a 0.1 to a 0.5 Cr/Fe molar ratio

whilst it decreased above 0.5 (figure 7b). For the Cr/Fe between 0.1 and0.5, Cr–Fe
17
might be coordinated via oxygen atoms [47] and the direct coordination was increased

gradually by decreasing the Cr–Fe interatomic distance at Cr/Fe molar ratio above 0.5.

This result suggests that more inner-sphere surface complexation exists for higher Cr/Fe

molar ratios. For high surface coverage at higher Cr/Fe molar ratio below pH 6, the

bidentate Cr–Fe relationship is more likely to be established [17]. For comparison,

goethite is used as a proxy for ferrihydrite. Like in our result, chromate showed an

inner-sphere complexation mechanism with bidentate dominance at a lower pH, and

monodentate species become more pronounced at a higher pH in goethite adsorption

[45]. Scott et al. showed that goethite forms a bidentate binuclear (3.29 Å) and bidentate

mononuclear complex (2.91 Å) with Cr in the lower surface coverage (pH 5) from

EXAFS results [41]. Our findings suggest that Cr–Fe surface complexation is close to

the bidentate binuclear complex (~3.378 Å) and comprises tetrahedral Cr and octahedral

Fe bonding [48] ( graphical abstract). We did not obtain Cr–Fe bidentate mononuclear

complexes. This may be because goethite contains more protonation sites than

ferrihydrite [17]. Protonation sites are more likely to displace water molecules over

hydroxyl groups, and thus they promote closer precipitation.

4. Conclusions

Coprecipitation determined a higher chromate removal compared with

adsorption. A mineralogical transformation revealed by XRD along with zeta potential


18
break-even point for the same Cr/Fe molar ratio as a proof of mechanism shift in

coprecipitation. The reverse trend for interatomic distance obtained from EXAFS

indicate inner-sphere surface complexation, which supports the mechanism shift and

confirmed the bidentate binuclear Cr–Fe bonding. Overall results accord and prove that

inner-sphere surface complexation occurs in coprecipitation comparatively at a lower

molar ratio than adsorption. Furthermore, the shift of sorption mechanism to polymer

formation for higher molar ratios suggested by the BET-type sorption isotherm.

Acknowledgments

The study was supported partially by the Ministry of Education, Culture, Sports,

Science and Technology (MEXT), Japan (MEXT no. 24561006/2012-2015). We would

like to thank Aichi Synchrotron Radiation Center, Aichi Science, and Technology, Aichi,

Japan, for the XAFS analysis.

19
References

[1] Y.M. Tzou, M.K. Wang, R.H. Loeppert, Sorption of phosphate and Cr(VI) by

Fe(III) and Cr(III) hydroxides, Arch. Environ. Contam. Toxicol. 44 (2003) 445–

453.

[2] A.A. Belay, Impacts of Chromium from Tannery Effluent and Evaluation of

Alternative Treatment Options, J. Environ. Prot. (Irvine,. Calif). 1 (2010) 53–58.

[3] US Environmental Protection agency, Review of the Environmental Effect of

Pollutants, Washington DC, 1978.

[4] D. Dermatas, T. Mpouras, M. Chrysochoou, I. Panagiotakis, C. Vatseris, N.

Linardos, et al., Origin and concentration profile of chromium in a Greek aquifer,

J. Hazard. Mater. 281 (2015) 35–46.

[5] M.A. Lilli, D. Moraetis, N.P. Nikolaidis, G.P. Karatzas, N. Kalogerakis,

Characterization and mobility of geogenic chromium in soils and river bed

sediments of Asopos basin, J. Hazard. Mater. 281 (2015) 12–19.

[6] E. Kaprara, N. Kazakis, K. Simeonidis, S. Coles, A.I. Zouboulis, P. Samaras, et

al., Occurrence of Cr(VI) in drinking water of Greece and relation to the

geological background, J. Hazard. Mater. 281 (2015) 2–11.

[7] F. Mou, J. Guan, Z. Xiao, Z. Sun, W. Shi, X.-A. Fan, Solvent-mediated synthesis

of magnetic Fe2O3 chestnut-like amorphous-core/γ-phase-shell hierarchical

nanostructures with strong As(v) removal capability, J. Mater. Chem. 21 (2011)

5414.

[8] Z. Wu, W. Li, P.A. Webley, D. Zhao, General and controllable synthesis of novel

mesoporous magnetic iron oxide@carbon encapsulates for efficient arsenic


20
removal, Adv. Mater. 24 (2012) 485–491.

[9] T. Ren, P. He, W. Niu, Y. Wu, L. Ai, X. Gou, Synthesis of α-Fe2O3 nanofibers for

applications in removal and recovery of Cr(VI) from wastewater, Environ. Sci.

Pollut. Res. 20 (2013) 155–162.

[10] R.M. Cornell, U. Schwertmann, The Iron Oxides: Structure, Properties, Reactions,

Occurrences and Uses, John Wiley & Sons, 2006.

https://books.google.com/books?id=CCnNKzh4oKUC&pgis=1 (accessed March

14, 2016).

[11] J.L. Jambor, J.E. Dutrizac, Occurrence and Constitution of Natural and Synthetic

Ferrihydrite, a Widespread Iron Oxyhydroxide., Chem. Rev. 98 (1998) 2549–

2586.

[12] J.F.F. Matthew F. Schultr, Mark M. Benjamin, Adsorption and Desorption of

Metals on Ferrihydrite: Reversibility of the Reaction and Sorption Properties of

the Regenerated Solid, Environ. Sci. Technol. 21 (1987) 863–869.

[13] C. Tokoro, Y. Yatsugi, H. Sasaki, S. Owada, PROCESSING A Quantitative

Modeling of Co-precipitation Phenomena in Wastewater Containing Dilute

Anions with Ferrihydrite using a Surface Complexation Model, Resour. Process.

8 (2008) 3–8.

[14] C. Tokoro, Y. Yatsugi, H. Koga, S. Owada, Sorption mechanisms of arsenate

during coprecipitation with ferrihydrite in aqueous solution, Environ. Sci.

Technol. 44 (2010) 638–43.

[15] D.A. Dzombak and F.M.M. Morel, Surface Complexation Modeling: Hydrous

ferric oxide, John Wiley&Sons, New York City, NY, USA,, NY, USA, 1990.
21
[16] D.Y.L. Hsia, T. H., S. L. Lo, C. F. Lin, CHEMICAL AND SPECTROSCOPIC

ADSORPTION OF CHROMATE EVIDENCE FOR SPECIFIC ON HYDROUS

IRON OXIDE, Chemosphere. 26 (1993) 1897–1904.

[17] C.P. Johnston, M. Chrysochoou, Investigation of chromate coordination on

ferrihydrite by in situ ATR-FTIR spectroscopy and theoretical frequency

calculations, Environ. Sci. Technol. 46 (2012) 5851–5858.

[18] T.K. Yoshio Takahashi, Daisuke Ariga, Qiaohui Fan, Systematics of Distributions

of Various Elements Between Ferromanganese Oxides and Seawater from

Natural Observation, Thermodynamics, and Structures, in: J. Ishibashi, K. Okino,

M. Sunamura (Eds.), Subseafloor Biosph. Linked to Hydrothermal Syst. TAIGA

Concept, Springer Japan, Tokyo, 2015: pp. 1–666.

[19] S Goldberg, Surface complexation modeling, Riverside, CA, USA, 2013.

http://linkinghub.elsevier.com/retrieve/pii/S0016703797814676.

[20] I. Garcia-Sosa, A. Cabral-Prieto, N. Nava, J. Navarrete, M.T. Olguin, L. Escobar,

et al., Sorption of chromium (VI) by Mg/Fe hydrotalcite type compunds,

Hyperfine Interact. 232 (2015) 67–75.

[21] C. Tokoro, S. Suzuki, D. Haraguchi, S. Izawa, Silicate removal in aluminum

hydroxide co-precipitation process, Materials (Basel). 7 (2014) 1084–1096.

[22] C. Tokoro, T. Sakakibara, S. Suzuki, Mechanism investigation and surface

complexation modeling of zinc sorption on aluminum hydroxide in

adsorption/coprecipitation processes, Chem. Eng. J. 279 (2015) 86–92.

[23] Y.O. and Daisuke Haraguchi, Chiharu Tokoro, S. Owada, Sorption mechanisms

of arsenate in Aqueous solution during coprecipitation with aluminum hydroxide,


22
J. Chem. Eng. Japan. 46 (2013) 173–180.

[24] Y.T. Chiharu Tokoro, Hajime Koga, Yugi Oda, Shuji Owada, XAFS investigation

for As (V) Co-precipitation Mechanism with Ferrihydrite, J. MMIJ. 127 (2011)

213–218.

[25] S. Leckie, J. O. Benjamin, M. M., Hayes, K., Kaufman, G., Altman,

Adsorption/Coprecipitation of Trace Elements from Water with Iron

Oxyhydroxide, Palo Alto, CA, 1980.

[26] E.A. Ball, J. W., Nordstrom, D. K., and Jenne, Additional and revised

thermochemical data and computer code for WATEQ2-a computerized chemical

model for trace and major element speciation and mineral equilibria of natural

waters., Reston, VA, USA, 1980.

[27] J.M. Zachara, C.E. Cowan, R.L. Schmidt, C.C. Ainsworth, CHROMATE

ADSORPTION BY KAOLINITE, Clays Clay Miner. 36 (1988) 317–326.

[28] S. Das, M. Jim Hendry, J. Essilfie-Dughan, Adsorption of selenate onto

ferrihydrite, goethite, and lepidocrocite under neutral pH conditions, Appl.

Geochemistry. 28 (2013) 185–193.

[29] P. Trivedi, J.A. Dyer, D.L. Sparks, Lead sorption onto ferrihydrite. 1. A

macroscopic and spectroscopic assessment, Environ. Sci. Technol. 37 (2003)

908–914.

[30] C. Tiberg, J.P. Gustafsson, Phosphate effects on cadmium(II) sorption to

ferrihydrite, J. Colloid Interface Sci. 471 (2016) 103–111.

[31] G.A. Waychunas, C.C. Fuller, J.A. Davis, J.J. Rehr, Surface complexation and

precipitate geometry for aqueous Zn(II) sorption on ferrihydrite: II. XANES


23
analysis and simulation, Geochim. Cosmochim. Acta. 67 (2003) 1031–1043.

[32] T.D. Waite, J.A. Davis, T.E. Payne, G.A. Waychunas, N. Xu, Uranium(VI)

adsorption to ferrihydrite: Application of a surface complexation model,

Geochim. Cosmochim. Acta. (1994).

[33] K. Kameda, Y. Hashimoto, S. Wang, Y. Hirai, Simultaneous and continuous

stabilization of As and Pb in contaminated solution and soil by a

ferrihydrite-gypsum sorbent, J. Hazard. Mater. 327 (2017) 171–179.

[34] V. Veselska, R. Fajgar, S. Cihalova, R.M. Bolanz, J. Gottlicher, R. Steininger, et

al., Chromate adsorption on selected soil minerals: Surface complexation

modeling coupled with spectroscopic investigation, J. Hazard. Mater. 318 (2016)

433–442.

[35] A.C. Cismasu, F.M. Michel, A.P. Tcaciuc, G.E. Brown, Properties of

impurity-bearing ferrihydrite III. Effects of Si on the structure of 2-line

ferrihydrite, Geochim. Cosmochim. Acta. 133 (2014) 168–185.

[36] E. Doelsch, J. Rose, A. Masion, J.Y. Bottero, D. Nahon, P.M. Bertsch, Speciation

and crystal chemistry of iron(III) chloride hydrolyzed in the presence of SiO 4

ligands. 1. An Fe K-edge EXAFS study, Langmuir. 16 (2000) 4726–4731.

[37] G.A. Waychunas, B.A. Rea, C.C. Fuller, J.A. Davis, Surface chemistry of

ferrihydrite: Part 1.EXAFS studies ongeometrie of coprecipitated and adsorbed

arsenate, 57 (1992) 2251–2269.

[38] H.I. Adegoke, F. AmooAdekola, O.S. Fatoki, B.J. Ximba, Adsorption of Cr (VI)

on synthetic hematite (α-Fe2O3) nanoparticles of different morphologies, Korean

J. Chem. Eng. 31 (2013) 142–154.


24
[39] Y. Wang, J. Ma, K. Chen, Adsorptive removal of Cr(vi) from wastewater by

α-FeOOH hierarchical structure: kinetics, equilibrium and thermodynamics, Phys.

Chem. Chem. Phys. 15 (2013) 19415.

[40] C. Tokoro, Removal Mechanism in Anionic Co-precipitation with Hydroxides in

Acid Mine Drainage Treatment, Resour. Process. 62 (2015) 3–9.

[41] S. Fendorf, M.J. Eick, P. Grossl, D.L. Sparks, Arsenate and Chromate Retention

Mechanisms on Goethite. 1. Surface Structure, Environ. Sci. Technol. 31 (1997)

315–320.

[42] Y. Hexiong, L. Ren, R.T. Downs, G. Costin, Goethite, alfa- FeO(OH), from

single-crystal data, Acta Crystallogr. Sect. E Struct. Reports Online. 62 (2006).

[43] F. Maillot, G. Morin, Y. Wang, D. Bonnin, P. Ildefonse, C. Chaneac, et al., New

insight into the structure of nanocrystalline ferrihydrite: EXAFS evidence for

tetrahedrally coordinated iron(III), Geochim. Cosmochim. Acta. 75 (2011) 2708–

2720.

[44] A. Manceau, W.P. Gatest, Surface structural model for ferrihydrite, Clay Clay

Miner. 45 (1997) 448–460.

[45] J. Xie, X. Gu, F. Tong, Y. Zhao, Y. Tan, Surface complexation modeling of Cr(VI)

adsorption at the goethite–water interface, J. Colloid Interface Sci. 455 (2015)

55–62.

[46] a. Manceau, Local Structure of Ferrihydrite and Feroxyhite by EXAFS

Spectroscopy, Clay Miner. 28 (1993) 165–184.

[47] P.A.R.E.A. BONNIN, Structure CristalUne de Fe2(CrO4)3.3H20 alpha, Acta

Cryst. 2 (1978) 706–709.


25
[48] K. Momma, F. Izumi, VESTA 3 for three-dimensional visualization of crystal,

volumetric and morphology data, J. Appl. Crystallogr. 44 (2011) 1272–1276.

26
Table 1 Fe K-edge EXAFS fitting results with changing Cr/Fe molar ratio at 0.05 M ionic strength and pH 5.

sample Fe-O Fe-Fe1 Fe-Fe2 E0 (eV) R factor

N R (Å) σ2 (Å2) N1 R (Å) σ2 (Å2) N2 R (Å) σ2 (Å2)

ferrihydrite 6.14 1.9882 0.01 1.89 3.0578 0.02 1.63 3.3508 0.02 -0.77 0.01

Cr/Fe 0.5 6.27 1.9864 0.01 1.58 3.0560 0.01 1.00 3.3490 0.02 -0.82 0.02

Cr/Fe 1 6.10 1.9875 0.01 1.38 3.0571 0.01 0.93 3.3501 0.03 -0.71 0.02

Cr/Fe 5 6.17 1.9928 0.01 1.45 3.0624 0.01 1.26 3.3554 0.02 -0.42 0.01

Cr/Fe 10 5.74 1.9960 0.01 1.43 3.0656 0.01 1.00 3.3586 0.02 -0.43 0.01

N: coordination number, σ2: Debye–Waller factor, E0: threshold shift in electron volts, R: interatomic distance.
Table 2 Cr K-edge EXAFS fitting results with changing Cr/Fe molar ratio at 0.05 M ionic strength and pH 5.

Sample Cr-O Cr-Fe E0 (eV) R Factor

N R (Å) σ2 (Å2) N R (Å) σ2 (Å2)

Cr/Fe 0.1 4.03 1.6344 0.0012 1.40 3.3808 0.0477 8.72 0.01

Cr/Fe 0.5 4.04 1.6350 0.0015 1.47 3.3814 0.0483 8.59 0.01

Cr/Fe 1 4.04 1.6327 0.0018 1.60 3.3791 0.0502 8.22 0.01

Cr/Fe 5 3.99 1.6307 0.0011 1.61 3.3771 0.0487 8.26 0.01

Cr/Fe 10 4.08 1.6300 0.0017 1.68 3.3764 0.0501 8.02 0.01

N: coordination number, σ2: Debye–Waller factor, E0: threshold shift in electron volts, R: interatomic distance.

28
List of figures:

 Figure 1: Relationship between Cr(VI) removal with Fe(III) dosage at pH 5 and

7 using adsorption and coprecipitation with ferrihydrite; initial concentration of

Cr(VI) was fixed at 0.19 mmol dm−3 and ionic strength was maintained at 0.05

mol dm−3.

 Figure 2: Sorption isotherm at pH 5 and 7 using coprecipitation and adsorption

with ferrihydrite; initial concentration of Cr(VI) was fixed at 0.19 mmol dm−3

and the ionic strength was maintained at 0.05 mol dm−3. Numbers in the figure

represent the initial molar ratio of Cr/Fe.

 Figure 3: XRD patterns of ferrihydrite with/without Cr(VI) at pH 5 obtained

from (a) coprecipitation process and (b) adsorption process.

 Figure 4: Relationship between sorption density and zeta potential during

adsorption and coprecipitation at pH 5. Numbers in the figure represent the

initial molar ratio of Cr/Fe.

 Figure 5: Fourier-transformed unfiltered spectra and fitting results of Fe K-edge

with changing Cr/Fe molar ratio (coprecipitation at pH 5).

 Figure 6: Fourier-transformed unfiltered spectra and fitting results of Cr K-edge

with changing Cr/Fe molar ratio (coprecipitation at pH 5).

 Figure 7: Inter-atomic distance with changing Cr/Fe molar ratio obtained from

EXAFS fitting: (a) Fe K-edge Fe–O inter-atomic distance, (b) Cr K-edge Cr–O

inter-atomic distance.

1
0.2

0.18 pH 7 adsorption
pH 7 coprecipitation
Residual Cr (VI) Conc [mmol dm -3]

0.16
pH 5 adsorption
0.14 pH 5 coprecipitation
0.12

0.1

0.08

0.06

0.04

0.02

0
0 1 2 3 4
Fe Dosage [mmol dm-3]

Figure 1

2
1.8
pH 7 adsorption
1.6 pH 7 coprecipitation 10
pH 5 adsorption
Sorption Density [mol-Cr/mol-Fe]

1.4 pH 5 coprecipitation
1.2 5

1
10
0.8
5
0.6

0.4 10
0.5 1
1 5 10
0.2 0.1 0.5 5
0.1 0.05 0.5 1 1
0.05 0.05 0.1 0.1
0 0.5
0 0.05 0.1 0.15 0.2
Equilibrium Cr(VI) Conc. [mmol dm-3]

Figure 2

3
(a)
23° 34° 23° 34° (b)

Cr/Fe 10
Cr/Fe 10

Relative Intensity [-]


Relative Intensity [-]

Cr/Fe 5
Cr/Fe 5

Cr/Fe 1 Cr/Fe 1

Cr/Fe 0.5 Cr/Fe 0.5

Cr/Fe 0.1 Cr/Fe 0.1

Ferrihydrite Ferrihydrite

0 10 20 30 40 50 60 70 80 0 10 20 30 40 50 60 70 80
Cu-Ka 2ϴ [º] Cu-Ka 2ϴ [°]

Figure 3.

4
Figure 4

5
Fe-O
Fitting results
Actual results

Fe-Fe

Fe-Fe
Cr/Fe 10
FT Magnitude

Cr/Fe 5

Cr/Fe 1

Cr/Fe 0.5

Ferrihydrite

0 1 2 3 4 5 6
Distance [Å]

Figure 5

6
Cr-O
Fitting result
Actual result
FT magnitude

Cr-Fe

Cr/Fe 10

Cr/Fe 5

Cr/Fe 1

Cr/Fe 0.5

Cr/Fe 0.1

0.5 1.5 2.5 3.5 4.5 5.5


Distance [Å]

Figure 6

7
1.997
(a) Fe-O
1.995
Interatomic distance [Å]

1.993

1.991

1.989

1.987

1.985
0 2 4 6 8 10
Cr/Fe molar ratio [Cr-mmol/Fe-mmol]

1.6355
(b) Cr-O
1.6345
Interatomic distance [Å]

1.6335

1.6325

1.6315

1.6305

1.6295
0 2 4 6 8 10
Cr/Fe molar ratio [Cr-mmol/Fe-mmol]

Figure 7

You might also like