Next Article in Journal
Single Crystals of EuScCuSe3: Synthesis, Experimental and DFT Investigations
Previous Article in Journal
Variation Pattern of the Compressive Strength of Concrete under Combined Heat and Moisture Conditions
Previous Article in Special Issue
Formation and Inhibition Mechanism of Na8SnSi6O18 during the Soda Roasting Process for Preparing Na2SnO3
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Mn-Doped Spinel for Removing Cr(VI) from Aqueous Solutions: Adsorption Characteristics and Mechanisms

1
School of Resources and Safety Engineering, Wuhan Institute of Technology, Wuhan 430205, China
2
School of Minerals Processing and Bioengineering, Central South University, Changsha 410083, China
*
Authors to whom correspondence should be addressed.
Materials 2023, 16(4), 1553; https://doi.org/10.3390/ma16041553
Submission received: 20 January 2023 / Revised: 8 February 2023 / Accepted: 10 February 2023 / Published: 13 February 2023
(This article belongs to the Special Issue Advances in Processing and Characterization of Mineral Materials)

Abstract

:
In this study, the manganese (Mn) was doped in the MnFe2O4 crystal by the solid-phase synthesis method. Under the optimum conditions (pH = 3), the max removal rate and adsorption quantity of Cr(VI) on MnFe2O4 adsorbent obtain under pH = 3 were 92.54% and 5.813 mg/g, respectively. The DFT calculation results indicated that the adsorption energy (Eads) between HCrO4 and MnFe2O4 is −215.2 KJ/mol. The Cr(VI) is mainly adsorbed on the Mn atoms via chemical bonds in the form of HCrO4. The adsorption of Mn on the MnFe2O4 surface belonged to chemisorption and conformed to the Pseudo-second-order equation. The mechanism investigation indicated that the Mn in MnFe2O4 has an excellent enhancement effect on the Cr(VI) removal process. The roles of Mn in the Cr(VI) removal process included two parts, providing adsorbing sites and being reductant. Firstly, the Cr(VI) is adsorbed onto the MnFe2O4 via chemisorption. The Mn in MnFe2O4 can form ionic bonds with the O atoms of HCrO4/CrO42−, thus providing the firm adsorbing sites for the Cr(VI). Subsequently, the dissolved Mn(II) can reduce Cr(VI) to Cr(III). The disproportionation of oxidized Mn(III) produced Mn(II), causing Mn(II) to continue to participate in the Cr(VI) reduction. Finally, the reduced Cr(III) is deposited on the MnFe2O4 surface in the form of Cr(OH)3 colloids, which can be separated by magnetic separation.

1. Introduction

Chromium (Cr) widely exists in the surface and underground water as a common poisonous heavy metal. The Cr in water mainly comes from industrial discharges, such as leather tanning, electroplating, metal plating, and chemical manufacturing [1,2,3]. There are two primary species of Cr, trivalent (Cr (III)) and hexavalent (Cr (VI)) ions in the aqueous system [4], and the toxicity and mobility of Cr are highly dependent on its existing species [5]. It is widely believed that the toxicity of Cr (VI) is much higher than that of Cr (III), and Cr(VI) is highly carcinogenic and teratogenic [6]. The hypertoxic Cr(VI) in the natural environment could cause a range of diseases and pose a major threat to humans and ecology [7,8]. Therefore, many present researchers have focused on the field of Cr(VI) reduction [9,10]. Among this research, the adsorption-reduction method for Cr(VI) removal has received the most extensive research [11]. Furthermore, some biosorbents have been developed for heavy-metal removal [12].
The current Cr(VI) reduction methods primarily include chemical, electrochemical, photocatalytic, and biological reduction processes [13,14,15,16]. Most of the literature has shown that adsorption is a prerequisite for Cr(VI) reduction [17]. Many adsorbents have been prepared to achieve the adsorption-reduction reaction of Cr(VI) on their surfaces [18,19,20]. Among these adsorbents, metallic oxide, especially spinel-type materials, has been most extensively investigated due to its good magnetic properties, easy synthesis, low cost, and chemical stability. Many published literature studies have verified that Cr(VI) can be reduced by certain low-valence metal ions such as Fe(II) and Mn(II) in spinel materials [21,22]. Mu et al. [23] verified that Mn(II) could promote the Cr(VI) reduction via oxalic acid. In the Cr(VI) reduction process, Cr(VI) oxidized Mn(II) to Mn(III). Then, the oxidized Mn(III) acted as an electron bridge, transferring the electron from oxalic acid to Cr(VI). Another recent paper adopted a synthetic Mn3O4@ZnO/Mn3O4 composite to reduce Cr(VI) [24]. As an electron carrier, Mn acted in the role of a catalyst during the Cr(VI) reduction. Another research study elaborated on the adsorption mechanisms of Cr(VI) on the surface of manganese oxide in detail [25]. The researchers found that Cr(VI) adsorption belongs to more than one type of reaction. The Cr(VI) can bind to functional groups by the inner-and outer-sphere chromate complexes under different pH values.
Based on the previous research on the effects of Mn on the Cr(VI) reduction, we synthesized a novel Mn-doped spinel to remove the Cr(VI) from the Cr-bearing solution. The appropriate amount of natural ferromanganese and hematite ore powder was mixed and subsequently roasted under suitable conditions, and Mn replaced the Fe atom at the A site in the Fe3O4 spinel, forming MnFe2O4 [26]. Some reports have manifested that the MnFe2O4 has good adsorbing ability on Cr(VI) [27,28,29,30,31,32]. However, these researchers primarily focused on the adsorption characteristics of Cr(VI) on MnFe2O4, and the mechanism of Cr(VI) removal had been neglected. In this research, DFT calculation was introduced to reveal the adsorption mechanism of Cr(VI)on the MnFe2O4 surface. A molecular model was built to study the interaction type between Cr(VI) and the MnFe2O4 surface. What is more, the MnFe2O4 material used in the previous research is synthesized from pure chemical raw materials, and its crystallinity, morphology, and impurity content are different from the MnFe2O4 prepared from natural minerals. Indeed, the MnFe2O4 is more special as an adsorbent compared with other kinds of spinel because of the non-negligible role of Mn in the Cr(VI) removal process. Some investigations have already manifested that Mn ions in MnFe2O4 have higher reaction activity compared to the Fe ions in Fe3O4 [33,34].
The critical point of this study is to deeply analyze the roles of Mn in the Cr(VI) removal process on the MnFe2O4 surface by means of the XPS test and DFT calculation. Furthermore, the adsorption characteristics of Cr(VI) adsorption on the MnFe2O4 adsorbent were investigated by the batch adsorption experiments. This study attempts to explain the adsorption mechanisms of the Cr(VI) adsorption on the MnFe2O4 adsorbent.

2. Materials and Methods

2.1. Materials

In the previous work, our group invented a novel high-temperature solid-phase synthesis method to prepare the MnFe2O4 adsorbent [26]. To begin with, the ferromanganese ore and hematite ore powders were evenly mixed, and the stoichiometric ratio of Mn and Fe in mixed raw materials was 1:2. The roasting temperature and time were controlled at 1200 °C and 1 h. The mixture briquettes were broken and ground to −0.074 mm, accounting for 70 wt% after roasting. Then the MnFe2O4 particles and the gangue minerals were separated by magnetic separation. The obtained MnFe2O4 particles continued to be nano-ground to the size of 200 nm (d0.5 = 200 nm). The nano-MnFe2O4 particles were used as a magnetic adsorbent for the Cr(VI) removal.
All chemicals used in this research were of analytical grade. The simulated Cr(VI)-bearing solution was prepared using K2Cr2O7 (Sinopharm Group Co., Ltd., Beijing, China). The pH was adjusted with 0.1 M HCl and 0.1 M NaOH solutions.

2.2. Adsorption Characteristics

Adsorption experiments were conducted under isothermal conditions using the batch method. The isothermal adsorption experiments were conducted to investigate the Cr(VI) adsorption characteristics of the MnFe2O4 adsorbent. A certain weight of the MnFe2O4 adsorbent was added to an Erlenmeyer flask with a 50 mL Cr(VI) solution to conduct the adsorption experiments. The mixed solutions were oscillated for 24 h to ensure the Cr(VI) achieved equilibrium. Then the filtrates were obtained by filtering through a 0.45 μm membrane filter. The equilibrium concentration of Cr(VI) was measured by ICP-MS analysis. The adsorbing amount of Cr(VI) on the MnFe2O4 adsorbent under various experimental conditions was calculated according to the following equation:
q e = C o C e · V m
where qe is the adsorbing amount of Cr(VI) on the MnFe2O4 adsorbent (mg·g−1); Co and Ce are the initial and equilibrium concentrations of the Cr(VI) solution (mg·L−1), respectively; V is the volume of the Cr(VI)-bearing solution; and m is the mass of the MnFe2O4 adsorbent.
The pH effect tests were conducted within the pH range of 1–10. The pH of the solution was adjusted by 0.1 M HCl and 0.1 M NaOH. The initial Cr(VI) concentration and MnFe2O4 dosage were controlled at 10 mg/L and 3 g/L, respectively. The adsorption kinetics of Cr(VI) on the MnFe2O4 were investigated by changing the oscillation time of the Cr(VI) adsorption. The oscillation times were changed from 5 to 720 min in this research, and the initial Cr(VI) concentration was controlled at 10 mg/L. The Cr(VI) adsorption data at the various oscillation time were analyzed according to the following two models, the pseudo-first-order equation and the pseudo-second-order equation. The linear forms of these two models are listed below [30]:
ln q e q t = ln q e k 1 t
t q t = 1 k 2 q e 2 + 1 q e t
where qe and qt are the adsorption amounts of Cr(VI) at equilibrium and a certain time t (mg·g−1), respectively; k1 and k2 are the rate constants of the pseudo-first-order equation and the pseudo-second-order equation, respectively.
The isothermal adsorption experiments were conducted with the initial Cr(VI) concentration range of 1–100 mg/L. The obtained equilibrium concentrations were fitted via the Langmuir and Freundlich models. The linearized Langmuir and Freundlich isotherm models are expressed as the following equations [31]:
C e / q e = C e / q m + 1 / k L q m
l o g q e = l o g k F + 1 / n l o g C e
where qe (mg·g−1) is the amount of Cr(VI) adsorbed per weight unit of the MnFe2O4 adsorbent after equilibrium; Ce (mg·L−1) is the equilibrium concentration of Cr(VI) in the solution; qm (mg·g−1) is the maximum adsorption capacity of the adsorption; KL is a constant related to the heat of adsorption (L·mg−1); KF (mg1−n·Ln·g−1) is a constant related to the extent of adsorption; and 1/n is related to the intensity of adsorption.

2.3. Characterization

The X-ray powder diffraction (XRD) patterns of the synthetic MnFe2O4 before and after nano-scale grinding were detected by a diffractometer with Cu Kα radiation (RIGAKU D/Max 2500, Tokyo, Japan). The tube current and voltage were 250 mA and 40 kV, the scanning range was 10–70°, and the step size was 0.02°. A vibrating sample magnetometer was adopted to measure the magnetism properties of the MnFe2O4 adsorbent (PPMS Dynacool, Quantum Design, San Diego, CA, USA). The element composition and binding energies of atomic orbital on the MnFe2O4 surface before and after adsorbing Cr(VI) were investigated by the XPS test (XPS, Mono 650 μm, 200 W, Al Kα, pass energy 20 eV). The scanning step size was 1.0 eV for the full scan and 0.1 eV for the narrow scan. Zeta potential measurements were conducted using a zetasizer under a wide pH range from 3 to 10 at 25 °C, and the pH value was adjusted using 0.1 M HCl and NaOH. The particle size distribution was characterized by a laser particle analyzer (Malvern Mastersizer 2000, Malvern, UK). The specific surface area and pore volume of the MnFe2O4 adsorbent were measured by nitrogen adsorption-desorption on a specific surface area with a pore analyzer.
The slab supercell that consists of a (2 × 2) unit cell of MnFe2O4 (100) was constructed as the substrate. A vacuum of 20 Å was added to eliminate the interaction between two periodic slabs and H atoms were attached to O on the surface in acidic conditions (MnFe2O4-H (100)). The calculations were conducted under density functional theory, and the Vienna ab initio simulation (VASP) code 1, 2 was used to implement. The initial adsorption site was predicted by the simulated annealing method for locating a good approximation to the global minimum. Projector Augmented Wave (PAW) pseudopotentials are used with a cutoff energy of 400 eV for plane wave expansions. The GGA-PBE5 was used as an exchange-correlation function. The atomic relaxation was performed with a k-mesh of 1 × 1 × 1 until the force on each atom is less than 0.02 eV/Å. Moreover, van der Waals forces between the adsorbate and substrate were also corrected by the DFT-D26 dispersion corrections. After relaxation, a gamma-center k-mesh 2 × 2 × 1 was adopted for the ground state total energy calculation. The adsorption energy (Eads) was evaluated by:
Eads = E(total) − E(sub) − E(ads)
where E(total), E(sub), and E(ads) are the total energy of the MnFe2O4-H (100)-adsorbate complex, MnFe2O4-H (100), and adsorbates (HCrO4, CrO42− and Cr2O72−) in the same cell, respectively. According to this definition, a negative value is energetically favorable.
The transfer charge (Δchg) is the net charge on adsorbates:
Δ chg = i = 1 n A i A i 0
where Ai and Ai0 are the self-consistent charge and the atomic charge of atom i, respectively. n is the total number of studied adsorbates. The negative value represents electron dissipation to the substrate and positive Δchg represents electron accumulation on the adsorbate.

3. Results

3.1. Characterization of MnFe2O4 Adsorbent

The effects of nano-scale grinding on the crystallinity and crystal integrity of MnFe2O4 were investigated by X-ray diffraction patterns. Figure 1a shows that the diffraction peaks at 29.95°, 35.33°, 56.42°, and 61.94° were considered to belong to the 220, 311, 511, and 440 crystal faces of MnFe2O4 [35], respectively. After grinding, the crystallinity of MnFe2O4 worsened and certain diffraction peaks disappeared. This result indicated that the strong mechanical grinding would damage the MnFe2O4 crystal to some extent. The magnetization shown in Figure 1d exhibited that the magnetization intensity of MnFe2O4 decreased sharply from 63.38 to 23.80 emu/g after grinding. The main reason is that the magnetism of MnFe2O4 is the macroscopic manifestation of the magnetic properties of many tiny magnetic domains inside it, and as the particle size decreases, the number of internal magnetic domains decreases, resulting in a weakening in the magnetism of nano-size MnFe2O4 particles. Fortunately, the magnetic saturation curves displayed in Figure 1d manifested that the saturation magnetic of nano-MnFe2O4 still retained 20 emu/g, illustrating that the MnFe2O4 adsorbent still had good magnetism and could be recovered by magnetic separation. The BET test manifested that the specific surface area of the MnFe2O4 adsorbent could achieve 31.997 m2/g after nano-scale grinding. The zeta potential results exhibited in Figure 1c illustrated that the pHpzc (zero-point charge pH value) of the MnFe2O4 is 7.5. The MnFe2O4 adsorbent was positively charged within the pH range of 3–7.5. Therefore, there is an electrostatic attraction between electropositive MnFe2O4 particles and chromate (HCrO4, CrO42−, and Cr2O72−).

3.2. Adsorption Characteristics

3.2.1. Effect of the Initial pH

In this study, the Cr(VI) adsorption tests were conducted under pH values of 1 to 10. As shown in Figure 2a, the Cr(VI) removal first increased within the pH range of 1–3 and then continued to decline with an increasing pH. The Cr(VI) reduction in the acidic system conformed to the following equations [36,37]:
HCrO4 + 3e + 7H+ → Cr3+ + 4H2O
Cr2O72−+ 6e + 14H+ → 2Cr3+ + 7H2O
The hydrogen ions participated in the Cr(VI) reduction, and this indicated that the acidic system is favorable for Cr(VI) reduction. Furthermore, according to the zeta potential tests of MnFe2O4, the electrostatic attraction between the MnFe2O4 and chromate is more reliable in a low-pH environment. However, when in a very strong acidic system (pH = 1 and 2), the reduction product Cr(III) could not be deposited on the MnFe2O4 surface by forming the Cr(OH)x compound due to the neutralization. This is the main reason that the Cr(VI) removal under pH = 1 and 2 is lower than that under pH = 3.

3.2.2. Adsorption Kinetics

The time needed to reach the adsorption equilibrium was a significant value that decided the Cr(VI) removal efficiency. In this experiment, the Cr(VI) adsorption kinetics was determined by measuring the adsorption amount of Cr(VI) at various times. The adsorption kinetic curve of Cr(VI) is displayed in Figure 3a, and the fitting curves of the pseudo-first-order and pseudo-second-order equations are displayed in Figure 3b,c, respectively. The fitting parameters of the pseudo-first-order equation and pseudo-second-order models are given in Table 1.
The adsorption kinetic curve of Cr(VI) is displayed in Figure 3a, and the fitting curves of the pseudo-first-order and pseudo-second-order equations are displayed in Figure 3b,c, respectively. The fitting parameters of the pseudo-first-order equation and pseudo-second-order models are given in Table 1. The correlation coefficient (R2) of these two models indicated that the pseudo-second-order model was more suitable to describe the Cr(VI) adsorption kinetics on MnFe2O4 adsorbent. The result showed that the Cr(VI) adsorption achieved equilibrium in 480 min, and further extending the adsorption time could hardly affect the adsorption amount of Cr(VI). The above results manifested that Cr(VI) adsorption on the MnFe2O4 adsorbent belonged to chemisorption [38].

3.2.3. Adsorption Isotherm

As many papers reported, the adsorption isotherm model fitting was a common tool to understand the adsorption type between the adsorbent and adsorbate [39]. The Langmuir and Freundlich isotherm models fitting lines and parameters of Cr(VI) adsorption on the MnFe2O4 surface are shown in Figure 4 and Table 2, respectively.
The results illustrated that the max adsorption quantity of Cr(VI) is 5.813 mg/g. The correlation coefficient (R2) of the Langmuir model is higher than that of the Freundlich model, showing that the Cr(VI) adsorption on the MnFe2O4 surface conformed to the Langmuir model. It was likely that the Cr(VI) adsorption tended to be monolayer adsorption [40]. Compared with other adsorbents, the adsorption capacity of the adsorbent in this study is lower. The main reason is that there are some impurities when using natural minerals for synthesis. What is more, the solid-phase synthetic MnFe2O4 particles are coarse, while after grinding, their specific surface area is still much smaller than that of the liquid-phase synthetic MnFe2O4 particles.

3.3. Enhancement Mechanism of Mn on Cr(VI) Adsorption and Reduction

3.3.1. Cr(VI) Adsorption Analyses

The Cr(VI) removal process on the MnFe2O4 surface primarily included the adsorption and reduction process. The Cr(VI) could only be reduced after being adsorbed on the MnFe2O4 surface. As much of the literature has reported, the DFT calculation is a powerful tool to investigate the interaction reaction between two kinds of molecules [41,42]. In this study, the MnFe2O4 (100) surface, a typical low-index surface, was used to investigate the role of Mn in the Cr(VI) adsorption process.
Figure 5a shows the optimum structure of spinel-type MnFe2O4, the bond distances of Fe-Fe, Fe-Mn, Fe-O, and Mn-O are 3.011 Å, 3.530 Å, 4.760 Å, and 3.687 Å, respectively. However, according to the stern double-layer theory [43], when MnFe2O4 is placed in an aqueous solution, a layer of hydrogen or hydroxyl groups adsorbed on the surface of MnFe2O4 molecules exists, which is called the “positioning ion” [44]. In this study, the Cr(VI) adsorption process primarily occurred in the acidic solution, and the MnFe2O4-H molecule was dominant compared with MnFe2O4-OH. We built up a new structure of MnFe2O4-H by causing the hydrogen ion to adsorb on the surface of the MnFe2O4 molecule. The optimum structure of MnFe2O4-H is displayed in Figure 5b. The bond distances of Fe-H, Mn-H, and O-H were 1.500 Å, 1.610 Å, and 1.100 Å, respectively.
Next, the groups of HcrO4, CrO42−, and Cr2O72− were selected to conduct DFT calculations. The optimum structures of MnFe2O4−H−HcrO4, MnFe2O4−H−CrO42−, and MnFe2O4−H−Cr2O72− are shown in Figure 6. Moreover, the adsorption energies and the bond distances of HcrO4, CrO42−, and Cr2O72− on MnFe2O4-H (100) are listed in Table 3.
As shown in Table 4, the adsorption energies of HcrO4, CrO42−, and Cr2O72− on the MnFe2O4-H surface were −215.2, −200.7, and −116.7 kJ/mol, respectively. The adsorption energy is an indicator of system stability calculated by subtracting the system’s energy before and after adsorption [45]. In general, the higher the adsorption energy is, the more stable the system after adsorption is [46]. Therefore, it can be easily observed that HcrO4 was the most easily adsorbed on the MnFe2O4 adsorbent. Moreover, it can also be observed that from the bond distances displayed in Table 3, in the MnFe2O4−H−HcrO4 system, the O atom of HcrO4 formed a chemical bond with the Mn atoms of MnFe2O4, providing high adsorption energy. As we know, the adsorption process is an essential link to the subsequent reduction [47,48]. The DFT calculation indicated that the HcrO4 could be efficiently adsorbed by the Mn atoms of the MnFe2O4 via the chemical bonds. This efficient adsorption of Cr(VI) created a favorable environment for the subsequent Cr(VI) reduction by the Mn(II).
Subsequently, the charge density distribution was calculated by VASP software to clarify the chemical bond type between the chromate and the MnFe2O4-H.
In general, the charges would be distributed at both ends of the bond in the electrostatic interaction or ionic bond [49]. In contrast, for the covalent bond, the charge would be distributed in the middle of the bond [50]. The charge density distribution shown in Figure 7 indicates that the charges of the Mn/Fe-O bonds in the MnFe2O4−H− HcrO4 system were distributed at the ends of the bonds. This implied that the Mn/Fe-O bond belonged to the ionic bond. Similarly, CrO42− was also adsorbed on the Mn atom of the MnFe2O4 adsorbent via the ionic bond, while during the Cr2O72− adsorption process, the bonding had little effect on the charge distribution of the two molecules. Therefore, it could be suggested that the adsorption of Cr2O72− on the MnFe2O4 surface is a type of physical adsorption via electrostatic attraction.

3.3.2. Cr(VI) Reduction Analyses

The chemical states on the MnFe2O4 surface before and after the Cr(VI) adsorption were investigated by the XPS test to explain the reduction mechanisms of Cr(VI). The XPS full-scanning spectra of the original MnFe2O4 adsorbent and MnFe2O4 adsorbed Cr(VI) under pH = 3 and 7 are shown in Figure 8. The main atom contents on the MnFe2O4 adsorbent surface before and after adsorbing Cr(VI) are displayed in Table 5.
As shown in Figure 8, there were four primary XPS peaks, 284.8, 532.8, 638.7, and 710.9 eV, belonging to the C1s, O1s, Mn2p, and Fe2p atomic orbitals, respectively. After the Cr(VI) adsorption, the Cr2p peak at 573.6 eV emerged. The results of atom contents showed that the Cr atom content of MnFe2O4-Cr at pH = 3 was significantly higher than that of MnFe2O4-Cr at pH = 7 due to the high adsorption quantity of the Cr(VI) under an acidic environment. It is worth noting that the Mn2p atom content declined after Cr(VI) adsorption, implying that the Mn atom in MnFe2O4 particles was involved in the Cr(VI) reduction reaction. It was inferred that the divalent Mn was oxidized to trivalent Mn by Cr(VI) and entered the solution, resulting in a decrease in the atom content of Mn on the surface of the MnFe2O4 adsorbent.
To further investigate the adsorption mechanisms of Cr(VI), the high-resolution spectra of Mn2p and Cr2p were conducted. From the Mn2p spectrum of the original MnFe2O4 adsorbent shown in Figure 9a, there is only one main peak at 641.5 eV of Mn-O considered to belong to the MnFe2O4. However, after the Cr(VI) adsorption, the peak of MnO2 emerged, indicating that the Mn(II) on the MnFe2O4 surface might be oxidized to tetravalence Mn by Cr(VI). In fact, Mn(II) was oxidized to Mn(III) first by Cr(VI). However, the Mn(III) was extremely unstable in an acidic solution and was prone to the disproportionation reaction to form Mn(II) and MnO2. The resulting Mn(II) could continue to participate in the Cr(VI) reduction, and MnO2 was deposited on the MnFe2O4 surface. The XPS results proved that Mn(II) acted as a reductant in the Cr(VI) reduction process.
The high-resolution Cr2p spectra of MnFe2O4-Cr depicted in Figure 10 could be resolved into three individual peaks: The groups of HCrO4/CrO42− (577.5 and 587.2 eV), Cr3+ (576.2 and 585.9 eV), and Cr2O72− (578.1 and 587.8 eV) [51]. Under the pH = 3 condition, the XPS spectrum showed that the main peak of Cr2p on the MnFe2O4-Cr surface is Cr(III). When the pH value increased to 7, the dominant species of Cr is Cr(VI). It could be seen that from the XPS results, the Cr adsorbed on the MnFe2O4 surface, mainly in the form of Cr(III) under pH = 3. This result confirmed that most of the Cr(VI) was reduced to Cr(III) by Mn(II) and formed the Cr(OH)3 compound on the MnFe2O4 surface. The rest of the Cr adsorbed onto the MnFe2O4 surface in the form of Cr(VI) via electrostatic adsorption. According to the XPS results, we could speculate that the Cr(VI) was adsorbed on the surface of the MnFe2O4 adsorbent in the form of chromate. Subsequently, most of the chromate was reduced to Cr(III) by Mn(II).

3.4. The Roles of Mn in Cr(VI) Adsorption and Reduction Process

According to the above results, it could be inferred that the roles of Mn in the Cr(VI) removal process could be divided into two parts, providing the adsorbing sites and being a reductant. The Cr(VI) removal process is explained in Figure 11.
In the Cr(VI) adsorption process, the Mn atoms of the MnFe2O4 could form a firm ionic bond with the O atoms of HCrO4 and CrO42−, thus providing the adsorbing sites for the Cr(VI) adsorption. Then, the adsorbed Cr(VI) was reduced by the Mn(II) dissolved on the MnFe2O4 surface. In the reduction process, the dissolved Mn(II) acted as a reductant. The Mn(II) first oxidized to Mn(III) by Cr(VI). The unstable Mn(III) would cause the disproportionated reaction and formed Mn(II) and MnO2 in the aqueous system [52]. The reaction link of the Cr(VI) reduction by Mn(II) can be described by the following equations:
3 M n 2 + + H C r O 4 + 7 H +     3 M n 3 + + C r 3 + + 4 H 2 O
C r 3 + + 3 O H   C r O H 3
2 M n 3 + + 2 H 2 O   M n 2 + + M n O 2 + 4 H +
After Cr(VI) reduction, the obtained Cr(III) was complexed with the hydroxy groups to form the inner-sphere complex [53]. The Cr(OH)3 precipitate was deposited on the MnFe2O4 surface and separated by magnetic separation.

4. Conclusions

The Cr(VI) could be removed efficiently by the MnFe2O4 adsorbent. The adsorption process belonged to chemisorption. The mechanism research showed that the roles of Mn in the Cr(VI) removal process could be divided into two parts, providing adsorbing sites and being a reductant. The adsorption mechanisms of the Cr(VI) on the MnFe2O4 adsorbent was the formation of ionic bonds between O atoms of HCrO4/CrO42− and the Mn/Fe atoms of the MnFe2O4 adsorbent, which provided firm adsorbing sites for the Cr(VI). Subsequently, the adsorbed Cr(VI) was reduced by Mn(II) dissolved from the MnFe2O4 surface. The reduced Cr(III) would form the Cr(OH)3 colloids deposited on the MnFe2O4 surface, then be removed by magnetic separation. The present work clarified the roles of Mn in the Cr(VI) removal process and provided the removal mechanisms of Cr(VI) on the MnFe2O4 adsorbent.

Author Contributions

Conceptualization and methodology, Z.S., Y.Z., and Q.L.; methodology, investigation, and writing, M.L.; methodology and supervision, H.Z.; investigation and data curation, J.W.; conceptualization and supervision, T.J. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China, grant numbers 52204276, 52074361, and 51904273.

Data Availability Statement

The data is unavailable due to privacy.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Mortazavian, S.; Hunyadi Murph, S.E.; Moon, J. Biochar nanocomposite as an inexpensive and highly efficient carbonaceous adsorbent for hexavalent Chromium removal. Materials 2022, 15, 6055. [Google Scholar] [CrossRef] [PubMed]
  2. Chen, Y.; Zeb, S.; Peng, J.; Zhao, Y.M.; Xu, C.J.; Li, L.; Cui, Y.; Sun, G.X. Enhanced adsorption of Cr(VI) under neutral conditions using a novel adsorbent with preorganized diquaternary ammonium structure. J. Mol. Liq. 2021, 322, 114905. [Google Scholar] [CrossRef]
  3. Liang, H.X.; Song, B.; Peng, P.; Jiao, G.J.; Yan, X.; She, D. Preparation of there-dimensional honeycomb carbon materials and their adsorption of Cr(VI). Chem. Eng. J. 2019, 367, 9–16. [Google Scholar] [CrossRef]
  4. Yang, S.Y.; Li, Q.; Chen, L.; Chen, Z.S.; Hu, B.W.; Wang, H.H.; Wang, X.K. Synergistic removal and reduction of U(VI) and Cr(VI) by Fe3S4 micro-crystal. Chem. Eng. J. 2020, 385, 123909. [Google Scholar] [CrossRef]
  5. Liu, Y.Q.; Shan, H.M.; Zeng, C.Y.; Zhan, H.B.; Pang, Y.Y. Removal of Cr(VI) from Wastewater Using Graphene Oxide Chitosan Microspheres Modified with α–FeO(OH). Materials 2022, 15, 4909. [Google Scholar] [CrossRef]
  6. Xu, X.Y.; Huang, H.; Zhang, Y.; Xu, Z.B.; Cao, X.D. Biochar as both electron donor and electron shuttle for the reduction transformation of Cr(VI) during its sorption. Environ. Pollut. 2019, 244, 423–430. [Google Scholar] [CrossRef] [PubMed]
  7. Karthik, C.; Barathi, S.; Pugazhendhi, A.; Ramkumar, V.S.; Thi, N.D.; Arulselvi, P.I. Evaluation of Cr(VI) reduction mechanism and removal by Cellilosimicrobium funkei strain AR8, a novel haloalkaliphilic bacterium. J. Hazard. Mater. 2017, 333, 42–53. [Google Scholar] [CrossRef]
  8. Han, C.Y.; Yang, L.; Yu, H.L.; Luo, Y.M.; Shan, X. The adsorption behavior and mechanism of Cr(VI) on facile synthesized mesoporous NH-SiO2. Environ. Sci. Pollut. R. 2020, 27, 2455–2463. [Google Scholar] [CrossRef]
  9. Arslan, H.; Eskikaya, O.; Bilici, Z.; Dizge, N.; Balakrishnan, D. Comparison of Cr(VI) adsorption and photocatalytic reduction efficiency using leonardite powder. Chemosphere 2022, 300, 134492. [Google Scholar] [CrossRef]
  10. Zhu, J.L.; Lei, P.; Liu, M.F.; He, P.; Chen, Y.Z.; Gan, M.; Zhu, J.Y. The Interplay of Iron Minerals and Microflora to Accelerate Cr(VI) Reduction. Materials 2022, 12, 460. [Google Scholar] [CrossRef]
  11. Benettayeb, A.; Ghosh, S.; Usman, M.; Seihoub, F.Z.; Sohoo, I.; Chia, C.H.; Sillanpää, M. Some Well-Known Alginate and Chitosan Modifications Used in Adsorption: A Review. Water 2022, 14, 1353. [Google Scholar] [CrossRef]
  12. Benettayeb, A.; Morsli, A.; Guibal, E.; Kessas, R. New derivatives of urea-grafted alginate for improving the sorption of mercury ions in aqueous solutions. Mater. Res. Express 2021, 8, 035303. [Google Scholar]
  13. Vilardi, G.; Palma, L.D.; Verdone, N. A physical-based interpretation of mechanism and kinetics of Cr(VI) reduction in aqueous solution by zero-valent iron nanoparticles. Chemosphere 2019, 220, 590–599. [Google Scholar] [CrossRef] [PubMed]
  14. Yang, Y.T.; Li, J.; Yan, T.; Zhu, R.X.; Yan, L.G.; Pei, Z.G. Adsorption and photocatalytic reduction of aqueous Cr(VI) by Fe3O4-ZnAl-layered double hydroxide/TiO2 composites. J. Colloid. Interf. Sci. 2020, 562, 492–504. [Google Scholar] [CrossRef] [PubMed]
  15. Peng, W.J.; Li, H.Q.; Liu, Y.Y.; Song, S.X. A review on heavy metal ions adsorption from water by graphene oxide and its composites. J. Mol. Liq. 2017, 230, 496–504. [Google Scholar] [CrossRef]
  16. He, F.; Lu, Z.Y.; Song, M.S.; Liu, X.L.; Tang, H.; Huo, P.W.; Fan, W.Q.; Dong, H.J.; Wu, X.Y.; Xing, G.L. Construction of ion imprinted layer modified ZnFe2O4 for selective Cr(VI) reduction with simultaneous organic pollutants degradation based on different reaction channels. Appl. Surf. Sci. 2019, 483, 453–462. [Google Scholar] [CrossRef]
  17. Mallik, A.K.; Moktadir, A.M.; Rahman, A.M.; Shahruzzaman, M.; Rahman, M.M. Progress in surface-modified silicas for Cr(VI) adsorption: A review. J. Hazard. Mater. 2022, 423, 127041. [Google Scholar] [CrossRef]
  18. Lu, M.M.; Zhang, Y.B.; Su, Z.J.; Tu, Y.K.; Wang, J.; Liu, S.; Liu, J.C.; Jiang, T. The comprehensive investigation on removal mechanism of Cr(VI) by humic acid-Fe(II) system structured on V, Ti-bearing magnetite surface. Adv. Powder. Technol. 2021, 32, 37–51. [Google Scholar] [CrossRef]
  19. Cao, F.M.; Sun, Y.Q.; Zhang, L.; Sun, J. High efficient adsorption accompanied by in-situ reduction of Cr(VI) removal by rice straw fiber ball coated with polypyrrole. Appl. Surf. Sci. 2022, 575, 151583. [Google Scholar] [CrossRef]
  20. Peng, X.; Liu, S.J.; Luo, Z.J.; Yu, X.W.; Liang, W.W. Selective removal of hexavalent chromium by novel nitrogen and sulfur containing cellulose composite: Role of counter anions. Materials 2023, 16, 184. [Google Scholar] [CrossRef]
  21. Liang, C.W.; Tang, B.; Zhang, X.D.; Fu, F.L. Mobility and transformation of Cr(VI) on the surface of goethite in the presence of oxalic acid and Mn(II). Environ. Sci. Pollut. R. 2020, 27, 26115–26124. [Google Scholar] [CrossRef] [PubMed]
  22. Wang, Y.; Chen, S.Y.; Yang, X.; Wu, Y.X.; Huang, X.F.; He, E.K.; Qiu, R.L.; Wang, S.Z. Enhanced removal of Cr(VI) in the Fe(III)/natural polyphenols system: Role of the in situ generated Fe(II). J. Hazard. Mater. 2019, 377, 321–329. [Google Scholar] [CrossRef]
  23. Mu, Y.; Jiang, Z.; Ai, Z.H.; Jia, F.L.; Zhang, L.Z. Mn2+ promoted Cr(VI) reduction with oxalic acid: The indispensable role of In-situ generated Mn3+. J. Hazard. Mater. 2018, 343, 356–363. [Google Scholar] [CrossRef]
  24. Li, N.; Tian, Y.; Zhao, J.H.; Zhang, J.; Zhang, J.; Zuo, W.; Ding, Y. Efficient removal of chromium from water by Mn3O4@ZnO/Mn3O4 composite under simulated sunlight irradiation: Synergy of photocatalytic reduction and adsorption. Appl. Catal. B-Environ. 2017, 214, 126–136. [Google Scholar] [CrossRef]
  25. Islam, M.A.; Angove, M.J.; Morton, D.W.; Pramanik, B.K.; Awual, M.R. A mechanistic approach of chromium (VI) adsorption onto manganese oxides and boehmite. J. Environ. Chem. Eng. 2020, 8, 103515. [Google Scholar] [CrossRef]
  26. Liu, B.B.; Zhang, Y.B.; Su, Z.J.; Lu, M.M.; Peng, Z.W.; Li, G.H.; Jiang, T. Formation mechanism of MnxFe3−xO4 by solid-state reaction of MnO2 and Fe2O3 in air atmosphere: Morphologies and properties evolution. Powder. Technol. 2017, 313, 201–209. [Google Scholar] [CrossRef]
  27. Eyvazi, B.; Zanjani, A.J.; Darban, A.K. Synthesis of nano-magnetic MnFe2O4 to remove Cr(III) and Cr(VI) from aqueous solution: A comprehensive study. Environ. Pollut. 2019, 29, 113685. [Google Scholar] [CrossRef]
  28. He, J.; Lo, I.M.C.; Chen, G.H. Fast removal and recovery of Cr(VI) using surface-modified jacobsite (MnFe2O4) nanoparticles. Langmuir 2005, 21, 11173–11179. [Google Scholar]
  29. Ahmadi, A.; Foroutan, R.; Esmaeili, H.; Tamjidi, S. The role of bentonite clay and bentonite clay@MnFe2O4 composite and their physico-chemical properties on the removal of Cr(III) and Cr(VI) from aqueous media. Environ. Sci. Pollut. R. 2020, 27, 14044–14057. [Google Scholar] [CrossRef]
  30. Pang, Y.X.; Kong, L.J.; Chen, D.Y.; Yuvaraja, G. Rapid Cr(VI) reduction in aqueous solution using a novel microwave-based treatment with MoS2-MnFe2O4 composite. Appl. Surf. Sci. 2019, 471, 408–416. [Google Scholar] [CrossRef]
  31. Wang, J.; Xu, Q.Y.; Hou, J.H.; Wang, S.S.; Wang, X.Z. Mechanism analysis of MnFe2O4/FeSX for removal of Cr(VI) from aqueous phase. Ecotox. Environ. Safe. 2021, 217, 112209. [Google Scholar] [CrossRef]
  32. Tuyen, T.V.; Chi, N.K.; Tien, D.T.; Tu, N.; Quang, N.V.; Huong, P.T.L. Carbon-encapsulated MnFe2O4 nanoparticles: Effects of carbon on structure, magnetic properties and Cr(VI) removal efficiency. Appl. Phys. A 2020, 126, 577. [Google Scholar] [CrossRef]
  33. Lin, Y.L.; Chen, J.J.; Cao, W.Z.; Persson, K.M.; Ouyang, T.; Zhang, L.; Xie, X.D.; Liu, F.; Li, J.; Chang, C.T. Novel materials for Cr(VI) adsorption by magnetic titanium nanotubes coated phosphorene. J. Mol. Liq. 2019, 287, 110826. [Google Scholar] [CrossRef]
  34. Lu, M.M.; Zhang, Y.B.; Zhou, Y.L.; Su, Z.J.; Liu, B.B.; Li, G.H.; Jiang, T. Adsorption-desorption characteristics and mechanisms of Pb(II) on natural vanadium, titanium-bearing magnetite-humic acid magnetic adsorbent. Powder. Technol. 2019, 344, 947–958. [Google Scholar] [CrossRef]
  35. Liu, S.Y.; Wang, L.J.; Chou, K. Selective chlorinated extraction of iron and manganese from vanadium slag and their application to hydrothermal synthesis of MnFe2O4. ACS Sustain. Chem. Eng. 2017, 5, 10588–10596. [Google Scholar] [CrossRef]
  36. Yang, X.; Liu, L.H.; Zhang, M.Z.; Tan, W.F.; Qiu, G.H.; Zheng, L.R. Improved removal capacity of magnetite for Cr(VI) by electrochemical reduction. J. Hazard. Mater. 2019, 374, 26–34. [Google Scholar] [CrossRef]
  37. Jiang, B.; Gong, Y.F.; Gao, J.N.; Sun, T.; Liu, Y.J.; Oturan, N.; Oturan, M.A. The reduction of Cr(VI) to Cr(III) mediated by environmentally relevant carboxylic acids: State-of-the-art and perspectives. J. Hazard. Mater. 2019, 365, 205–225. [Google Scholar] [CrossRef]
  38. Guo, X.; Wang, J.L. A general kinetic model for adsorption: Theoretical analysis and modeling. J. Mol. Liq. 2019, 288, 11100. [Google Scholar] [CrossRef]
  39. Huang, X.P.; Hou, X.J.; Song, F.H.; Zhao, J.C.; Zhang, L.Z. Facet-dependent Cr(VI) adsorption of hematite nanocrystals. Environ. Sci. Technol. 2016, 50, 1964–1972. [Google Scholar] [CrossRef]
  40. Yazidi, A.; Sellaoui, L.; Dotto, G.L.; Bonilla-Petriciolet, A.; Frohlich, A.C.; Lamine, A.B. Monolayer and multilayer adsorption of pharmaceuticals on activated carbon: Application of advanced statistical physics models. J. Mol. Liq. 2019, 283, 276–286. [Google Scholar] [CrossRef]
  41. Ma, R.Q.; Wang, T.; Huang, T.B.; Sun, W.L.; Qiao, S.; Liu, W. Insights into the interactions of Cr(III) and organic matters during adsorption onto titanate nanotubes: Differential adsorbance and DFT study. J. Mol. Liq. 2020, 312, 113432. [Google Scholar] [CrossRef]
  42. Li, M.Q.; Mu, Y.; Shang, H.; Mao, C.L.; Cao, S.Y.; Ai, Z.H.; Zhang, L.Z. Phosphate modification enables high efficiency and electron selectivity of nZVI toward Cr(VI) removal. Appl. Catal. B-Environ. 2020, 263, 118364. [Google Scholar] [CrossRef]
  43. Robben, B.; Beunis, F.; Neyts, K.; Callens, M.; Johansson, T.; Beales, G.; Fleming, R.; Strubbe, F. Optical evidence for adsorption of charged inverse micelles in a Stern layer. Colloid. Surface. A 2020, 589, 124451. [Google Scholar] [CrossRef]
  44. Mei, L.J.; Chou, T.H.; Cheng, Y.S.; Huang, M.J.; Yeh, L.H.; Qian, S.Z. Electrophoresis of pH-regulated nanoparticles: Impact of the Stern layer. Phys. Chem. Chem. Phys. 2016, 18, 9927–9934. [Google Scholar] [CrossRef]
  45. Yin, S.X.; Ellis, D.E. DFT studies of Cr(VI) complex adsorption on hydroxylated hematite (1102) surfaces. Surf. Sci. 2009, 603, 736–746. [Google Scholar] [CrossRef]
  46. Li, Q.L.; Liu, Y.; Yu, X.F.; Li, L.L.; Zhang, X.H.; Lu, Z.M.; Lin, J.; Yang, X.J.; Huang, Y. Removal of Cr(III)/Cr(VI) from wastewater using defective porous boron nitride: A DFT study. Inorg. Chem. Front. 2018, 5, 1933–1940. [Google Scholar] [CrossRef]
  47. Zhang, H.L.; Li, P.; Wang, Z.M.; Cui, W.W.; Zhang, Y.; Zhang, Y.; Zheng, S.L.; Zhang, Y. Sustainable disposal of Cr(VI): Adsorption–reduction strategy for treating textile wastewaters with amino-functionalized boehmite hazardous solid wastes. ACS Sustain. Chem. Eng. 2018, 6, 6811–6819. [Google Scholar] [CrossRef]
  48. Chen, C.S.; Li, H.; Wen, Y.Z.; Zhao, F.; Zhang, Y.P.; Wu, D.L.; Liu, Y.B.; Li, F. Spherical Cu2O-Fe3O4@chitosan bifunctional catalyst for coupled Cr-organic complex oxidation and Cr(VI) capture-reduction. Chem. Eng. J. 2020, 383, 123105. [Google Scholar]
  49. Bader, R.F.W.; Henneker, W.H. The ionic bond. J. Am. Chem. Soc. 1965, 87, 3063–3068. [Google Scholar] [CrossRef]
  50. Howard, S.T.; Lamarche, O. Description of covalent bond orders using the charge density topology. J. Org. Chem. 2003, 16, 133–141. [Google Scholar] [CrossRef]
  51. Park, D.; Yun, Y.S.; Park, J.M. XAS and XPS studies on chromium-binding groups of biomaterial during Cr(VI) biosorption. J. Colloid. Interf. Sci. 2008, 317, 54–61. [Google Scholar] [CrossRef] [PubMed]
  52. Qian, A.; Zhang, W.; Shi, C.; Pan, C.; Giammar, D.E.; Yuan, S.; Zhang, H.L.; Wang, Z.M. Geochemical Stability of Dissolved Mn(III) in the Presence of Pyrophosphate as a Model Ligand: Complexation and Disproportionation. Environ. Sci. Technol. 2019, 53, 5768–5777. [Google Scholar] [CrossRef] [PubMed]
  53. Xie, J.Y.; Gu, X.Y.; Tong, F.; Zhao, Y.P.; Tan, Y.Y. Surface complexation modeling of Cr(VI) adsorption at the goethite–water interface. J. Colloid. Interf. Sci. 2015, 455, 55–62. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The XRD pattern of MnFe2O4 before and after nano-scale grinding (a); the BET adsorption−desorption curve of MnFe2O4 adsorbent (b); zeta potential of MnFe2O4 adsorbent (c); the magnetization hysteresis loops of MnFe2O4 before and after nano−scale grinding (d).
Figure 1. The XRD pattern of MnFe2O4 before and after nano-scale grinding (a); the BET adsorption−desorption curve of MnFe2O4 adsorbent (b); zeta potential of MnFe2O4 adsorbent (c); the magnetization hysteresis loops of MnFe2O4 before and after nano−scale grinding (d).
Materials 16 01553 g001
Figure 2. The Cr(VI) removal (a) and species (b) at various pH values.
Figure 2. The Cr(VI) removal (a) and species (b) at various pH values.
Materials 16 01553 g002
Figure 3. Adsorption kinetics analysis of Cr(VI) on MnFe2O4 adsorbent: (a) Adsorption amount of Cr(VI) at various times; (b) fitting line of the pseudo−first−order equation; (c) fitting line of the pseudo-second-order equation.
Figure 3. Adsorption kinetics analysis of Cr(VI) on MnFe2O4 adsorbent: (a) Adsorption amount of Cr(VI) at various times; (b) fitting line of the pseudo−first−order equation; (c) fitting line of the pseudo-second-order equation.
Materials 16 01553 g003
Figure 4. Langmuir (a) and Freundlich (b) isotherms for the adsorption of Cr(VI) onto the MnFe2O4 adsorbent.
Figure 4. Langmuir (a) and Freundlich (b) isotherms for the adsorption of Cr(VI) onto the MnFe2O4 adsorbent.
Materials 16 01553 g004
Figure 5. The optimum structure of (a) MnFe2O4 (100) surface and (b) MnFe2O4−H (100) surface.
Figure 5. The optimum structure of (a) MnFe2O4 (100) surface and (b) MnFe2O4−H (100) surface.
Materials 16 01553 g005
Figure 6. The optimum structure of MnFe2O4−H−HcrO4 (a), MnFe2O4−H−CrO42−, (b) and MnFe2O4−H−Cr2O72− (c).
Figure 6. The optimum structure of MnFe2O4−H−HcrO4 (a), MnFe2O4−H−CrO42−, (b) and MnFe2O4−H−Cr2O72− (c).
Materials 16 01553 g006
Figure 7. The charge density of MnFe2O4−H−HcrO4 (a), MnFe2O4−H−CrO42−, (b) and MnFe2O4−H−Cr2O72− (c).
Figure 7. The charge density of MnFe2O4−H−HcrO4 (a), MnFe2O4−H−CrO42−, (b) and MnFe2O4−H−Cr2O72− (c).
Materials 16 01553 g007
Figure 8. The XPS spectra of MnFe2O4 adsorbent, MnFe2O4−Cr (pH = 3), MnFe2O4-Cr (pH = 7).
Figure 8. The XPS spectra of MnFe2O4 adsorbent, MnFe2O4−Cr (pH = 3), MnFe2O4-Cr (pH = 7).
Materials 16 01553 g008
Figure 9. The Mn2p decomposition spectra of MnFe2O4 (a), MnFe2O4−Cr (pH = 3) (b) and MnFe2O4−Cr (pH = 7) (c).
Figure 9. The Mn2p decomposition spectra of MnFe2O4 (a), MnFe2O4−Cr (pH = 3) (b) and MnFe2O4−Cr (pH = 7) (c).
Materials 16 01553 g009
Figure 10. The Cr2p decomposition spectra of MnFe2O4−Cr at pH = 3 (a) and pH = 7 (b).
Figure 10. The Cr2p decomposition spectra of MnFe2O4−Cr at pH = 3 (a) and pH = 7 (b).
Materials 16 01553 g010
Figure 11. The diagram of the Cr(VI) adsorption and reduction processes.
Figure 11. The diagram of the Cr(VI) adsorption and reduction processes.
Materials 16 01553 g011
Table 1. Kinetic fitting parameters for the pseudo-first-order and pseudo-second-order equations for Cr(VI) adsorption onto the MnFe2O4 adsorbent.
Table 1. Kinetic fitting parameters for the pseudo-first-order and pseudo-second-order equations for Cr(VI) adsorption onto the MnFe2O4 adsorbent.
Pseudo-First-Order EquationPseudo-Second-Order Equation
Adsorbateqe (mg/g)qe,c (mg/g)k1 (min−1)R2qe,c (mg/g)K2 (g/mg·min−1)R2
Cr(VI)333526980.008340.95935380.005970.989
Table 2. The parameters for Langmuir and Freundlich models for Cr(VI) adsorption onto the MnFe2O4 adsorbent.
Table 2. The parameters for Langmuir and Freundlich models for Cr(VI) adsorption onto the MnFe2O4 adsorbent.
T (K)LangmuirFreundlich
qm (mg/g)kL (L·mg−1)R2kF (mg1−n·Ln·g−1)1/nR2
298581312820.999524480.2700.9121
Table 3. Comparison of the results of this research with other related research.
Table 3. Comparison of the results of this research with other related research.
AdsorbentsMetal IonsSynthetic Raw MaterialsAdsorption
Capacity/mg/g
Refs
MnFe2O4Cr(VI)Natural minerals5813This study
BCMFCCr(VI)Chemical reagents178.6[27]
Nano-MnFe2O4Cr(VI)Chemical reagents31.5[25]
MnFe2O4/FeSX−0.5Cr(VI)Chemical reagents43.36[29]
MFO/C NPsCr(VI)Chemical reagents73.26[30]
Table 4. The adsorption energy, the shortest distance between Mn/Fe and O, and the charge transfer between MnFe2O4−H (100) and adsorbates.
Table 4. The adsorption energy, the shortest distance between Mn/Fe and O, and the charge transfer between MnFe2O4−H (100) and adsorbates.
Eads/kJ/molD(Mn-O)/ÅD(Fe-O)/ÅΔchg
HcrO4−215.21.881.940.609
CrO42−−200.71.892.930.524
Cr2O72−−116.73.413.300.032
Table 5. Atom contents of the surfaces of the three samples.
Table 5. Atom contents of the surfaces of the three samples.
ElementsAtom Content of MnFe2O4/%Atom Content of MnFe2O4-Cr (pH3)/%Atom Content of MnFe2O4-Cr (pH7)/%
C1s24.0830.0229.76
O1s60.6556.0055.58
Fe2p11.7311.1711.35
Mn2p3.552.033.11
Cr2pNone0.780.19
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Lu, M.; Su, Z.; Zhang, Y.; Zhang, H.; Wang, J.; Li, Q.; Jiang, T. Mn-Doped Spinel for Removing Cr(VI) from Aqueous Solutions: Adsorption Characteristics and Mechanisms. Materials 2023, 16, 1553. https://doi.org/10.3390/ma16041553

AMA Style

Lu M, Su Z, Zhang Y, Zhang H, Wang J, Li Q, Jiang T. Mn-Doped Spinel for Removing Cr(VI) from Aqueous Solutions: Adsorption Characteristics and Mechanisms. Materials. 2023; 16(4):1553. https://doi.org/10.3390/ma16041553

Chicago/Turabian Style

Lu, Manman, Zijian Su, Yuanbo Zhang, Hanquan Zhang, Jia Wang, Qian Li, and Tao Jiang. 2023. "Mn-Doped Spinel for Removing Cr(VI) from Aqueous Solutions: Adsorption Characteristics and Mechanisms" Materials 16, no. 4: 1553. https://doi.org/10.3390/ma16041553

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop