Next Article in Journal
Variations in Sedimentation Rate and Corresponding Adjustments of Longitudinal Gradient in the Cascade Reservoirs of the Lower Jinsha River
Previous Article in Journal
Efficiency of Microalgae Employment in Nutrient Removal (Nitrogen and Phosphorous) from Municipal Wastewater
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effects of Drying Operation on the Mn2+ Removal Activity of MnOx: Performance and Mechanism

1
School of Urban Planning and Municipal Engineering, Xi’an Polytechnic University, Xi’an 710048, China
2
Key Laboratory of Membrane Separation of Shaanxi Province, Xi’an University of Architecture and Technology, Xi’an 710055, China
*
Author to whom correspondence should be addressed.
Water 2025, 17(2), 261; https://doi.org/10.3390/w17020261
Submission received: 2 December 2024 / Revised: 15 January 2025 / Accepted: 15 January 2025 / Published: 17 January 2025

Abstract

:
Mn2+ is a prevalent contaminant in groundwater. In this study, manganese oxides (MnOx) were prepared via a redox method to remove Mn2+ from water. The effects of the drying operation in the preparation process, including heat drying at 20–120 °C with different times and freeze drying methods, on the structural properties, manganese removal performance and mechanisms were investigated. The results indicate that the drying conditions can significantly affect the removal performance and stability of MnOx. The MnOx dried at 50 °C for 12 h exhibited the best Mn2+ removal efficiency and stability, with an adsorption capacity of 125.7 mg/g and removal efficacy of 95.1% after six reuse cycles. The removal pathway experiments revealed that the sample dried at 50 °C for 12 h had superior catalytic oxidation abilities for Mn2+, while other samples removed Mn2+ by primarily relying on the adsorption process. The investigation of the structure revealed that excessive heat drying led to the shrinkage of the oxide particles, a reduction in the surface voids, and a decrease in the hydroxyl groups. Conversely, insufficient drying time or temperatures resulted in high water content in MnOx, which occupied the surface active sites. The XPS analysis indicated that the catalytic oxidation of Mn2+ primarily relied on Mn(III) and adsorbed oxygen on the surface of MnOx. With freeze drying or inadequate heat drying, a large amount of Mn(II) remained on the oxide surface, and the over-drying operation resulted in excessive conversion from Mn(II) to Mn(IV), reducing the catalytic activity and resulting in low removal stability.

1. Introduction

Groundwater is an important drinking water source in China [1]. Manganese contamination is a prevalent issue in groundwater, leading to various sensory and operational challenges, such as discoloration, laundry stains, and pipe clogging [2]. Moreover, it also poses significant health risks [3], including chronic poisoning, abnormal neurological function, and endocrine dysfunction [4]. In nature, manganese is widely found in igneous and sedimentary rocks, and it can be dissolved into groundwater [5]. When groundwater flows through the soil and rocks, the concentration of manganese will be naturally elevated, due to the weathering and leaching of manganese-containing minerals [6]. Additionally, manganese can also come from human activities, such as industrial emissions, mining and mineral processing, and waste landfills [7]. Industrial wastewater containing manganese may be directly discharged into water bodies, resulting in serious manganese pollution accidents in aquatic environments [5]. Moreover, manganese can enter aquifers via the leaching solutions and runoff from mining, solid waste, and landfills [8].
At present, manganese contamination is becoming a growing issue due to the increasing demand for materials containing manganese, such as steel products and batteries [7]. Moreover, self-purification and restoration are almost impossible once contamination occurs, due to the lengthy natural renewal cycle of groundwater, making the issue more serious [9]. Therefore, the effective removal of manganese from groundwater remains a critical challenge in drinking water treatment [10].
Several methods are employed for the removal of manganese from drinking water, including chemical reagent oxidation, adsorption methods, biological processes, and chemical catalytic oxidation [11,12]. Among these, the utilization of manganese oxides as adsorbents or catalysts is a particularly promising approach [13]. Manganese oxides are abundantly present in the Earth’s crust [13]. They possess a strong affinity for metal ions and can efficiently remove Mn2+ from water due to their large specific surface area, abundant micropores, and high reactivity [14].
MnOx can be utilized either as powder materials or as coatings on various filter media, such as quartz sand, zeolite, and activated carbon [3,6], and it also plays crucial roles in chemical reagent oxidation and biological processes [14,15]. The removal of Mn2+ by manganese oxides occurs primarily through adsorption and catalytic oxidation pathways simultaneously. The removal efficiency and mechanisms of Mn2+ by MnOx are commonly influenced by the surface morphology, the valence state of Mn, and the species of reactive oxygen [16]. It has been proven that Mn2+ can react with the surface hydroxyl groups to form a hydroxyl complex connected by a coordination bond [17]. Additionally, structural defects are also an effective way to improve the adsorption capacity [18]. For example, oxygen vacancies are found to be capture centers, which are conducive to the adsorption of inorganic ions [19]. It has been known that Mn(III) has high reactivity and can promote the oxidation of Mn2+ [20]. According to the theory of catalytic oxidation, the removal process involves rapid adsorption followed by relatively slow catalytic oxidation, as described in Equations (1) and (2) [21].
Mn 2 + + Mn O 2 · x H 2 O s Mn O 2 · MnO · x 1 H 2 O s + 2 H +
Mn O 2 · MnO · ( x 1 ) H 2 O ( s ) + 1 2 O 2 + H 2 O 2 Mn O 2 · x H 2 O s
Initially, Mn2+ is adsorbed onto the active surface sites of MnOx; subsequently, MnOx facilitates the efficient utilization of dissolved oxygen to oxidize the adsorbed Mn2+, converting it back into MnOx and regenerating the active sites in the process [22]. Our previous studies have investigated the impact of structural cations on the activity of MnOx, revealing that Ca2+ doping can enhance Mn2+ removal more effectively than Na+, K+, and Mg2+.
Additionally, manganese oxides commonly contain large amounts of surface-adsorbed water and structural water [23]. The amount and behavior of this water are closely related to the ambient temperature and the valence state of manganese, significantly impacting the Mn2⁺ removal performance and mechanism [24]. Drying operations are commonly employed to control the water content and species in the preparation process of MnOx, and it can also induce some changes in the structure [25]. It was observed that the catalytic activity and stability of electrodeposited manganese oxide were slightly improved following heat treatment at 90 °C for 2 h [26]. Conversely, the loss of interlayer water molecules during natural drying was found to negatively affect the catalytic activity of Fe-Mn co-oxides [27]. Although drying operations are crucial in the preparation of MnOx, their effects on the Mn2+ removal activity of MnOx and the corresponding underlying mechanisms are still unclear, and there is little research that addresses this problem.
The aim of this study was to investigate the effects of the drying operation on the water properties, structural changes, and Mn2+ removal performance of MnOx. Moreover, the corresponding influence mechanism of the drying operation was investigated. The results will help to select appropriate drying operations in the preparation of high-reactivity MnOx and further elaborate on the removal mechanism of Mn2+.

2. Materials and Methods

2.1. Preparation of MnOx

First, 10.08 mL of 0.3 mol/L KMnO4 solution was slowly added to 10.2 mL of 0.1 mol/L CaCl2 and 6% MnCl2 solution. After the reaction was finished, the suspension was filtrated to obtain MnOx particles. The prepared particles were washed with deionized water several times until the pH of the cleaning water was approximately 7.0. Then, the obtained MnOx was placed in an oven or a vacuum freeze dryer and dried with different drying operations. Finally, the dried MnOx was kept in a sealed container before use.

2.2. Characterization of MnOx

The morphology of MnOx was characterized by scanning electron microscopy (SEM, Quanta FEG 250, FEI, Hillsborough, OR, USA). XRD patterns were recorded by an X-ray diffractometer (XRD, UltimalV, Rigaku, Tokyo, Japan) with Cu Kα radiation (λ = 0.1542 nm) and a tube voltage of 40 kV. Continuowere taken in a 2θ range of 5–80° with a scan rate of 5° min−1 and a step size of 0.02°. X-ray photoelectron spectroscopy (XPS) was performed using a photoelectron spectrometer (Thermo Scientific K-Alpha, Waltham, MA, USA) equipped with a monochromatic Al Kα excitation source. The binding energies were calibrated using the C1s binding energy at 284.6 eV. FTIR spectra were detected by a Fourier transform infrared spectrometer (Thermo Fisher Nicolet IS5, Waltham, MA, USA); the FTIR spectra of MnOx under different drying conditions were detected in the range of 4000–400 cm−1 with a resolution of 0.16 cm−1. Thermogravimetric–differential thermal analysis (HTC-4, Hengjiu, Bejing, China) was determined in the temperature range of 25 to 600 °C with a heating rate of 10 °C/min under flowing dry air.

2.3. Evaluation of Removal Capacity and Mechanism

2.3.1. Adsorption Models

Experiments on the isothermal adsorption and adsorption kinetics were carried out on a magnetic stirrer. The temperature was 25 °C, and the amount of added manganese oxides was 0.2 g/L. The reaction solution was a mixture of MnCl2 and NaHCO3. As the pH can significantly influence the adsorption behavior of MnOx, NaHCO3 of 200 mg/L was added into the reaction solution to maintain the pH within the range of 7.9–8.3 in the Mn2+ removal process. The concentration of Mn2+ in the isothermal adsorption and adsorption kinetics experiments was 15–90 mg/L and 25 mg/L, respectively. Samples of 5 mL were taken at 5, 10, 20, 30, 60, 90, 150, and 300 min and were filtered through 0.22 μm filter membranes to determine the concentration of Mn2+.
Pseudo-first- and second-order kinetics were used to analyze the adsorption process of Mn2+ on MnOx. The adsorption kinetics models and their equations are shown in Equations (3) and (4). The pseudo-first-order kinetics is as follows:
log q e q t = log q e k 1 2.303   t
where qe is the adsorption capacity at equilibrium, mg/g; qt is the adsorption capacity at time t, mg/g; t is the adsorption time, h; and k1 is the equilibrium constant, 1/min. The pseudo-second-order kinetics is as follows:
t q t = 1 k 2 q e 2 + 1 q e t
where k2 is the relative equilibrium constant, 1/min.
The adsorption isotherm model is a crucial tool in evaluating the adsorption capacities and characteristics of adsorbents. Commonly used theoretical models include the Langmuir and Freundlich isotherm models. The Langmuir isotherm model is based on several key assumptions, such as that the adsorbent surface is homogeneous, all adsorption sites have identical energies, and adsorption occurs in a monomolecular layer, etc. [28]. The model is described as in Equation (5):
q e = q max K L C e 1 + K L C e
where qmax is the maximum adsorption capacity, mg/g; KL represents the adsorption equilibrium constant; and Ce corresponds to the fluoride ion concentration after equilibrium, mg/L.
The Freundlich isothermal model applies to both monolayer adsorption (chemisorption) and multilayer adsorption (physisorption); it is based on the fact that the adsorption process occurs on the inhomogeneous surface of the adsorbent [29]. The equation is shown in Equation (6):
q e = K F · C e 1 n
where n is the adsorption constant; KF is the adsorption equilibrium constant.
The adsorption constant n is affected by the temperature, and, when n < 0.5, the adsorption process is considered to be difficult. When n = 2–10, the adsorption process can proceed easily [30]. The value of 1/n is related to the adsorption kinetic force and adsorption potential energy, and the smaller the value, the stronger the adsorption strength and the heterogeneity of the adsorbent surface [31].

2.3.2. Stability and Applicability Experiments

Manganese oxide of 0.2 g/L was added into a beaker and dispersed with pure water by ultrasonication. Then, the suspension was mixed with a certain amount of MnCl2/NaHCO3 solution. The concentrations of Mn2+ and NaHCO3 were 5 mg/L and 200 mg/L, respectively, and the volume of the reaction solution was 1 L. After this, the beaker was placed on a magnetic stirrer to carry out the reaction at a temperature of 25 °C. One reaction cycle was 300 min. Samples of 5 mL were taken at 5, 10, 20, 30, 60, 90, 150, and 300 min and were filtered through 0.22 μm filters to determine the concentration of Mn2+. After one cycle was completed, the reaction solution was filtered to separate MnOx and then the separated MnOx was cleaned several times using deionized water. Finally, the MnOx was used again in the next cycle. A total of 6–8 cycles were processed in this experiment.
To elaborate on the applicability of MnOx, tap water from groundwater sources was used to simulate real conditions. The MnOx treated at 50 °C for 12 h was coated on quartz sand and a dynamic filtration test was conducted, as reported in a previous study, to evaluate the applicability of MnOx [32]. Firstly, quartz sand with a particle size of 0.5–2.0 mm was soaked in 1 mol/L HCl for 24 h and then washed with deionized water to neutral and dried at 105 °C for backup use. Secondly, 500 g of acid-modified quartz sand was thoroughly mixed with a solution of 51 mL containing 0.1 mol/L CaCl2 and 6% MnCl2, and then 0.3 mol/L KMnO4 of 50.4 mL was slowly added. After the reaction was completed, the sands were dried at 50 °C for 12 h to obtain MnOx filter material. Finally, the filter material was filled into a plexiglass column with an inner diameter of 50 mm. The depth of the filter material was 1000 mm. Sampling ports were set at every 200 mm along the height of the filter material. The filtration rates and the Mn2+ concentration in the influent water of the filter were in the range of 6–12 m/h and 0.3–1.0 mg/L, respectively. Besides the concentration of Mn2+, all water parameters are shown in Table 1.

2.3.3. Effects of Oxygen

The effects of oxygen on Mn2+ removal were investigated by the method reported in a previous study, with a minor modification [33]. A MnCl2/NaHCO3 solution of 1 L was added into a flat-bottomed flask. The concentrations of NaHCO3 and Mn2+ in the solution were 200 mg/L and 0.2 g/L, respectively. Then, 0.2 g of MnOx was added to the above solution to carry out the reaction. In the experiment, air or N2 was continuously poured into the reactor with a certain flow rate, and the pH of the solution was adjusted with 1 M HCl and NaOH to keep the pH maintained at about 8.5. Samples of 5 mL were taken out at certain time intervals (0.5 h, 1.0 h, 2.0 h, 5.0 h, 8.0 h, 11.0 h, 14.0 h, 24.0 h) and filtrated with a 0.22 μm filter membrane, and the Mn2+ concentration was determined.

2.4. Analytical Methods

The chemical reagents used in the experiments were purchased from Sinopharm Chemical Reagent Co., Ltd. (Shanghai, China). The reagents were all of analytical grade and used without further purification. The pH and dissolved oxygen (DO) were measured using a precision acidity meter (PHS-3 C, Leici Co., Ltd., Shanghai, China) and a dissolved oxygen meter (JPB-607A, Leici Co., Ltd., Shanghai, China), respectively. The concentrations of Mn2+ were measured by potassium periodate oxidation spectrophotometry according to the guidelines of the Ministry of Environmental Protection of China.

3. Results and Discussion

3.1. Removal Performance of Mn2+

The removal performance of Mn2+ by MnOx treated under different drying conditions is shown in Figure 1. It can be seen that when the drying time increased from 1 h to 12 h, the Mn2+ removal efficiency of MnOx increased obviously, and the maximum removal rate of Mn2+ at the reaction time of 300 min was up to 99.6% (Figure 1a). However, extending the drying time to 24 h resulted in a reduction in the Mn2+ removal efficiency. The effect of the drying temperature on the Mn2+ removal is shown in Figure 1b. When the drying temperature increased from 20 °C to 50 °C, the Mn2+ removal efficiency increased. As we continued to increase the temperature, the Mn2+ removal efficiency began to decline, and the removal efficiency of the sample dried at 120 °C was the lowest. Additionally, the Mn2+ removal rate of the 12 h freeze-dried MnOx was comparable to that of the MnOx dried at 20 °C or 80 °C. When the drying temperature is low or the drying time is short, more adsorbed water will remain on the surface of MnOx. Excessive adsorbed water can lead to a decrease in the number of reactive sites [34], resulting in a relatively low Mn2+ removal rate by MnOx. However, when the drying temperature is too high or the drying time is too long, there may be a change in the structure of MnOx [35]. Thus, there exist optimal drying conditions for MnOx to obtain the maximum Mn2+ removal efficiency.

3.2. Adsorption Kinetics

The kinetic fitting results are shown in Figure 2a,b and Table 2. Both pseudo-first-order and pseudo-second-order kinetic models can effectively describe the adsorption process. The correlation coefficients (R2) for both models are greater than 0.99, with the pseudo-second-order model providing a superior fit. The pseudo-second-order model assumes that the adsorption process is governed by chemical adsorption [36], suggesting that the adsorption of Mn2+ involves a chemical adsorption process. In addition, the highest adsorption capacity was obtained by 50 °C-12 h-MnOx, with a qe of 125.7 mg/g.

3.3. Adsorption Isotherms

The adsorption isotherms of Mn2+ by MnOx are shown in Figure 3, and the corresponding isothermal adsorption fitting results are shown in Table 3. As shown in Figure 3a, the Langmuir model cannot fit the adsorption process well. Therefore, the relative parameters were not calculated.
The regression coefficient R2 of the Freundlich isothermal model is higher than 0.9, suggesting that the Freundlich isothermal model can accurately describe the Mn2+ adsorption process on MnOx. The parameter 1/n, which reflects the adsorption kinetic force and potential energy, is inversely related to the adsorption performance, and a smaller 1/n value signifies superior adsorption capacity [31,37]. Thus, it can be concluded that 50 °C-12 h-MnOx had the best adsorption capabilities for Mn2+.

3.4. Stability and Applicability Analysis of MnOx for Mn2+ Removal

The removal efficiency at 150 min in different reuse cycles was calculated, and the results are shown in Figure 4a. It indicates that, with an increase of the number of reaction cycles, the Mn2+ removal efficiency decreased. The highest Mn2+ removal efficiency in the sixth cycle was 95.1%. It was obtained for 50 °C-12 h-MnOx, followed by 50 °C-3 h-MnOx, 80 °C-12 h-MnOx, and freeze-dried-12 h-MnOx, with removal efficiency of 87.8%, 87.3%, and 88.9%, respectively. This suggests that the MnOx dried at 50 °C for 12 h had the optimal Mn2+ removal stability. The Mn2+ removal performance of MnOx in different reuse cycles with the reaction time was also investigated, and the results are shown in Figure S1. It also suggests that 50 °C-12 h-MnOx has the superior stability in Mn2+ removal.
The 50 °C-12 h-MnOx was coated on quartz sand and a dynamic filtration test was conducted to evaluate the applicability of the MnOx [32]. The results are shown in Figure 4b–d. This indicated that the Mn2+ concentration in the effluent water of the filter column was stably lower than 0.1 mg/L within the operation period of 22 days. When the influent concentration of Mn2+ was up to 1.0 mg/L or the filtration rate was up to 12 m/h, the concentration of Mn2+ in the effluent water could also meet the drinking water standards. Therefore, it can be concluded that MnOx has good applicability for groundwater treatment.

3.5. Effects of Drying Conditions on the Structural Characterization of MnOx

3.5.1. Effects on Morphology

The morphology and the corresponding particle size distribution of the MnOx treated with different drying operations are shown in Figure 5. It shows that all the samples were composed of nano-sized lamellar particles. The average diameters of the particles on the surface of 50 °C-12 h-MnOx and freeze-dried-12 h-MnOx were close, which were 156.6 and 162.2 nm, respectively. Conversely, for 50 °C-24 h-MnOx and 120 °C-12 h-MnOx, the size of the particles was smaller. The corresponding average diameters of the particles were 118.9 and 117.1 nm, respectively. Shrinkage deformation was observed, which resulted in a denser surface. These observations suggest that a long drying time or high drying temperature can lead to particle compaction, which may negatively affect the removal of Mn2+.

3.5.2. Thermogravimetric Analysis

The TG and DTG curves of MnOx dried under different drying conditions are presented in Figure 6. The TG curves demonstrate that MnOx underwent noticeable weight loss in the drying process. Water in manganese oxides primarily exists in three forms, namely adsorbed water, structural water, and hydroxyl groups [27]. The complete removal of the water molecules in manganese oxides commonly occurs at about 400 °C [35]. Therefore, the weight loss of MnOx at a temperature lower than 400 °C can be attributed to the decrease in the water content in the oxide. The results reveal that the lowest water content was observed in the freeze-dried-12 h-MnOx (17.48%), followed by 120 °C-12 h-MnOx, with water content of 18.55%; the highest water content occurred in 20 °C-12 h-MnOx (23.06%). This indicates that freeze drying is the most effective method for water removal, followed by 120 °C-12 h-MnOx. Notably, the water content in 50 °C-24 h-MnOx was slightly higher than that in 50 °C-12 h-MnOx, suggesting that extending the drying time at 50 °C has little effect on water removal.
Two significant peaks are observed in the DTG curves. The peak observed in the range of 35–300 °C corresponds to the loss of water molecules from MnOx. The asymmetry of this peak may be caused by the desorption and reabsorption of the water molecules in the heating process, as well as the overlapping temperature ranges associated with the different water species [24]. Additionally, the right-side tailings of the peak for 120 °C-12 h-MnOx and freeze-dried-12 h-MnOx are more obvious, suggesting higher proportions of structural water and hydroxyl groups in these oxides than in the others. The peak observed at temperatures higher than 500 °C is likely due to the loss of lattice oxygen, as well as the phase transition from MnO2 to Mn3O4 [38]. The higher temperature position of the peak corresponds to the lower oxygen activity in MnOx, which requires more energy for the phase transformation [39]. It shows that when the drying temperature or time increased, the peak position gradually shifted to the right. The most significant shift was observed for the peak at 544 °C in 120 °C-12 h-MnOx, indicating that the activity of the lattice oxygen in 120 °C-12 h-MnOx was the lowest.

3.5.3. FTIR and XRD Analysis

The FTIR spectra of MnOx are shown in Figure S2a. The broad absorption peaks observed between 3000 and 3600 cm⁻1 correspond to the stretching vibrations of adsorbed water molecules on the surface, while the peak at 1630 cm⁻1 is attributed to the bending vibrations of structural water molecules or Mn-OH groups [27]. The absorption peaks in the range of 400–700 cm⁻1 correspond to Mn-O vibrations in MnO2, which are associated with the deformation of [MnO6] octahedra [40]. Moreover, these peaks are commonly significant in oxides containing Mn3+ and Mn4+ and vary among different MnOx [41]. The peak at 480 cm⁻1 was the most significant in 50 °C-1 h-MnOx, followed by 20 °C-12 h-MnOx. Its intensity decreased with a higher drying temperature or longer drying time. The most obvious reductions were observed in 120 °C-12 h-MnOx and freeze-dried-12 h-MnOx. Additionally, the intensity of the peak at 1630 cm⁻1, associated with structural water or Mn-OH groups, was the lowest in 120 °C-12 h-MnOx compared with the other oxides. The results indicate that the water molecules were the most effectively removed from 120 °C-12 h-MnOx, followed by freeze-dried-12 h-MnOx.
The XRD patterns of MnOx are shown in Figure S2b. Two diffraction peaks of 37.12° and 67.76° were observed across all five samples, with no significant difference. The patterns correspond to Akhtensite-type MnO2 (ε-MnO2) [42]. Moreover, the intensity of the peaks was weak, indicating that the MnOx prepared in this study had low crystallinity, which could provide more adsorption sites and enhance the Mn2+ removal [42]. Additionally, a peak at about 24.94° was observed in the patterns of the freeze-dried-12 h-MnOx and 50 °C-24 h-MnOx. This can be attributed to α-MnO2 [25], indicating that a small quantity of α-MnO2 formed in the 50 °C-24 h-MnOx and freeze-dried-12 h-MnOx during the drying process.

3.5.4. XPS Spectra Analysis

The XPS spectra of the Mn 2p and O1s of the manganese oxides are shown in Figure 7. They indicated that the Mn 2p spectra can be fitted into three valence states of Mn(II), Mn(III), and Mn(IV). The corresponding percentages of different species of Mn are shown in Table 4. It can be seen that the percentage of Mn(II) and Mn(III) decreased with the increase in the drying temperature or drying time. In addition, the Mn(III) proportion of 50 °C-12 h-MnOx was the highest and was up to 60.9%. Since Mn(III) has high reactivity and plays an important role in the oxidation of Mn2+ [43], the high removal efficiency of 50 °C-12 h-MnOx could be attributed to the high Mn(III) proportion. The highest Mn(II) proportion of 46.50% was found in the freeze-dried-12 h-MnOx. When operating in a high-vacuum environment, the freeze drying process has little effect on the valence of Mn. Therefore, the composition of the freeze-dried-12 h-MnOx can be regarded as the initial state of the MnOx. Thus, it can be found that, in the order of freeze-dried-12 h-MnOx→50 °C-12 h-MnOx→50 °C-24 h-MnOx→120 °C-12 h-MnOx, the proportion of Mn(II) decreased continually, the proportion of Mn(III) increased firstly and then decreased, and the proportion of Mn(IV) remained unchanged initially and then gradually increased. Consequently, it can be concluded that Mn(II) was first transformed to Mn(III), and then Mn(III) was transformed to Mn(IV) with the increased drying time or drying temperature.
The molar ratios of Mn(III)/Mn(IV) were also calculated (Table 4). The highest molar ratio of 3.12 was found in 50 °C-12 h-MnOx and the lowest ratio of 0.68 was observed in 120 °C-12 h-MnOx. It is known that when Mn(III) is presented on the surfaces of oxides, more oxygen vacancies (VO) will be generated [44]. The concentration of VO is positively correlated with the amount of Mn(III) on the surface of MnOx [45]. Therefore, it can be inferred that the amount of VO was 50 °C-12 h-MnOx > freeze-dried-12 h-MnOx > 50 °C-24 h-MnOx > 120 °C-12 h-MnOx. Moreover, the appropriate value of VO can significantly enhance the reactivity of O in oxides [44], which in turn facilitates the catalytic activity of MnOx [46]. This provides a good explanation for the high manganese removal efficiency of 50 °C-12 h-MnOx.
The XPS spectra of O1s in the manganese oxides are shown in Figure 7b. In the spectra, the peaks at the binding energy of around 530.0 eV can be attributed to the lattice oxygen (Olatt) bonded to Mn-O-Mn [47]. The peaks at about 531.5 eV correspond to the surface-adsorbed oxygen (Oads), including O2 or O, etc. [48], and the peaks at about 532.7 eV are associated with the oxygen in surface-adsorbed water (Owater). The content of the surface-adsorbed oxygen was 50 °C-12 h-MnOx > freeze-dried-12 h-MnOx > 120 °C-12 h-MnOx > 50 °C-24 h-MnOx. The surface-adsorbed oxygen, which is related to the VO on the surface of MnOx, plays an important role in the oxidation of Mn2+ due to its good mobility [48]. It can also be used to explain the activity change in the oxides.

3.6. Analysis of the Mechanism

3.6.1. The Removal Pathway of Mn2+

The removal performance of Mn2+ by different MnOx with nitrogen or oxygen is shown in Figure 8. In the experiment, the dissolved oxygen (DO) concentration of the reaction solution with O2 was about 7.53 mg/L, while the DO concentration of the reaction solution with N2 was lower than 0.15 mg/L. Therefore, the solution with N2 can be regarded as an anaerobic environment [33]. Mn2+ is mainly removed through the adsorption process under the conditions of N2, and the removal of Mn2+ with O2 occurs via both adsorption and catalytic oxidation. The results showed that the removal rate of Mn2+ by MnOx in an aerobic environment was higher than that in an anaerobic environment, indicating that all oxides had the ability to catalyze the oxidation of Mn2+.
However, there was no significant difference in the removal efficiency of Mn2+ between 50 °C-24 h-MnOx, 120 °C-12 h-MnOx, and freeze-dried-12 h-MnOx in aerobic and anaerobic environments, indicating that the removal process was dominated by the adsorption process. The removal rate of Mn2+ by 50 °C-12 h-MnOx in an aerobic environment was 99.3%. It was much higher than that in an anaerobic environment (69.1%). This indicates that catalytic oxidation plays an important role in the removal of Mn2+ by 50 °C-12 h-MnOx, which can explain the excellent stability of 50 °C-12 h-MnOx observed in Figure 4.

3.6.2. XPS Spectra After the Reaction

The XPS spectra of Mn 2p and O1s in the oxides after the reaction are shown in Figure S3. The proportion of Mn(II) on the surface of 50 °C-12 h-MnOx after reaction in an aerobic environment is much lower than that after reaction in anaerobic environment, indicating that Mn2+ can be effectively oxidized by O2. In addition, the proportion of Mn(II) in 50 °C-12 h-MnOx after reaction in an aerobic environment is also the lowest among all oxides. This suggests that 50 °C-12 h-MnOx had the highest catalytic activity for Mn2+ oxidation. This result is consistent with that in Figure 8. Compared with the spectra before the reaction (Figure 7), the proportion of Mn(III) in 50 °C-12 h-MnOx decreased, and the corresponding proportion of Mn(IV) increased, suggesting that the Mn(III) and oxygen vacancies were consumed in the reaction. As shown by the XPS spectra of O1s (Figure S3b), the content of Owater on the surfaces of the oxides increased significantly after the reaction, and the corresponding content of 50 °C-12 h-MnOx was significantly lower than that in the others. It is known that water molecules are preferentially adsorbed at Mn2+ sites with an adsorption energy of −16.5 kcal/mol [49]. Thus, the increase in the Owater can be attributed to the adsorption of Mn2+ on the surface of MnOx. Additionally, the content of Oads decreased after the reaction, indicating that Oads was involved in the removal of Mn2+.

3.6.3. Summary of the Mechanism

The results indicate that the drying operations had no obvious influence on the adsorption process of MnOx in a single adsorption experiment, with the adsorption capacity in the range of 110–125.7 mg/g (Table 2). However, it had significant effects on the Mn2+ removal efficiency and stability (Figure 1 and Figure 4). Based on the results of the TG-DTG curves (Figure 6) and FTIR spectra (Figure S2a), it can be seen that drying at 120 °C for 12 h and the freeze drying method had the most significant water removal performance. Theoretically, the removal of water molecules can release more active sites on the surface of MnOx, enhancing the removal of Mn2+ [26,34]. However, the removal efficiency and stability of MnOx dried at 120 °C for 12 h were lower than for that dried at 50 °C for 12 h (Figure 4). Additionally, the structural changes in MnOx with heat drying and freeze drying were significantly different. The XPS spectra (Figure 7) suggest that, upon increasing the drying temperature or time, the Mn(II) on the surfaces of the oxides will first transform into Mn(III) and then transform into Mn(IV) in the drying process, while the freeze-dried-12 h-MnOx maintained the highest Mn(II) ratio. The investigation of the removal pathway (Figure 8) confirmed that Mn2+ was removed by both adsorption and catalytic oxidation. The XPS spectra of MnOx after the reaction (Figure S3) suggest that Mn(III) and Oabs are critical for the oxidation of Mn2+. Excessive drying reduced the content of Mn(III) and Oabs and freeze drying maintained an excessive amount of Mn(II) in MnOx (Figure 7a), thus resulting in a decrease in the efficiency and stability of Mn2+ removal. It can be concluded that the manganese removal capability of MnOx cannot be assessed merely by the adsorption process; catalytic oxidation is more critical for the Mn2+ removal stability. The drying operation can decrease the water content and alter the species of Mn and O, thus affecting the catalytic oxidation of Mn2+ by MnOx.
The process is summarized in Equations (7)–(10). When the MnOx samples were dried, the adsorption and structural water molecules on the surface could be removed and they released more active sites, enhancing the removal of Mn2+ theoretically (Equation (7)). However, the drying operation could also influence the chemical valence of Mn in MnOx, which would result in different catalytic activity for Mn2+ removal. Appropriate drying operations, such as at 50 °C for 12 h, could promote the conversion from Mn(II) to Mn(III) (Equation (8)), and the abundant Mn(III) would oxidize Mn2+ effectively (Equation (9)). However, excessive drying operations, such as at 120 °C for 12 h, would lead to the transformation from Mn(III) to Mn(IV), which negatively affects the catalytic oxidation of Mn2+. As catalytic oxidation is more critical for the Mn2+ removal stability, appropriate drying operations could enhance the removal performance of Mn2+.
MnO x · nH 2 O   drying MnO x + nH 2 O
Mn ( II )   appropriate   drying Mn ( III ) excessive   drying Mn ( IV )
Mn(III) + Mn2+ → MnOx + Mn(II)
Mn(II) + Oabs → Mn(III)

4. Conclusions

In this study, the effects of drying operations on the water properties, structural characteristics, manganese removal performance, and mechanisms of MnOx were investigated. The results demonstrated that excessive drying time or temperatures led to the shrinkage of the manganese oxide particles, a reduction in surface voids, and a decrease in manganese hydroxyl groups. Conversely, inadequate drying operations resulted in high water content in the oxide, which occupied active sites for Mn2+ removal. Additionally, the drying conditions can significantly influence the stability of Mn2+ removal by MnOx. Reuse tests indicated that the sample dried at 50 °C for 12 h exhibited the highest manganese removal stability after six reuse cycles. The removal pathway experiments showed that Mn2+ was removed through both adsorption and catalytic processes and the sample dried at 50 °C for 12 h had the strongest catalytic oxidation ability compared with the other samples. The mechanistic analysis revealed that the catalytic oxidation of Mn2+ by MnOx primarily depended on the presence of Mn(III) and adsorbed oxygen on the surface of MnOx. Higher content of Mn(III) and adsorbed oxygen significantly enhanced the catalytic oxidation of Mn2+. Moreover, freeze drying or inadequate heat drying operations resulted in a large proportion of Mn(II) on the oxide surface, while excessive heat drying operations caused excessive conversion from Mn(II) to Mn(IV), reducing the catalytic oxidation capabilities of the oxides.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/w17020261/s1, Figure S1: Mn2+ removal performance of MnOx with reuse cycles; Figure S2: (a) FTIR spectra and (b) XRD patterns of MnOx under different drying conditions; Figure S3: XPS spectra of (a) Mn 2p and (b) O1s of MnOx after the reaction.

Author Contributions

Conceptualization, R.Z.; methodology, L.Y. and J.Y.; investigation, L.Y. and J.Y.; data curation, L.Y. and J.Y.; writing—original draft preparation, R.Z.; writing—review and editing, R.Z.; project administration, Q.N. and B.Z.; funding acquisition, R.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Nature Science Basic Research Plan in Shaanxi Province of China (2024JC-YBQN-0548), the Key Laboratory of Membrane Separation of Shaanxi Province (2022MFL01), the Key Research and Development Projects of Shaanxi Province, China [2024SF-YBXM-573], and the Xi’an Municipal Science and Technology Project, China [23GXFW0023].

Data Availability Statement

The data supporting this study’s findings are available from the corresponding authors upon reasonable request.

Conflicts of Interest

No potential conflict of interest is reported by the authors.

References

  1. Cheng, Y.; Xiong, W.; Huang, T.; Wen, G. Study on the preparation of manganese oxide filter media for catalytic oxidation removal of ammonium and manganese in high alkalinity groundwater: The effect of copper and cobalt doping. J. Clean. Prod. 2022, 366, 132815. [Google Scholar] [CrossRef]
  2. Fang, K.; Wang, X.; Peng, Z.; Dong, J.; Du, X.; Luo, Y. Gravity driven ceramic membrane loaded birnessite functional layer for manganese removal from groundwater: The significance of disinfection on biofilm. Sep. Purif. Technol. 2024, 332, 125735. [Google Scholar] [CrossRef]
  3. Shrestha, A.M.; Kazama, S.; Sawangjang, B.; Takizawa, S. Improvement of Removal Rates for Iron and Manganese in Groundwater Using Dual-Media Filters Filled with Manganese-Oxide-Coated Sand and Ceramic in Nepal. Water 2024, 16, 2450. [Google Scholar] [CrossRef]
  4. Yang, H.; Yan, Z.; Du, X.; Bai, L.; Yu, H.; Ding, A.; Li, G.; Liang, H.; Aminabhavi, T.M. Removal of manganese from groundwater in the ripened sand filtration: Biological oxidation versus chemical auto-catalytic oxidation. Chem. Eng. J. 2020, 382, 123033. [Google Scholar] [CrossRef]
  5. Ghosh, S.; Mohanty, S.; Akcil, A.; Sukla, L.B.; Das, A.P. A greener approach for resource recycling: Manganese bioleaching. Chemosphere 2016, 154, 628–639. [Google Scholar] [CrossRef] [PubMed]
  6. Jerroumi, S.; Amarine, M.; Gourich, B. Technological trends in manganese removal from groundwater: A review. J. Water Process Eng. 2023, 56, 104365. [Google Scholar] [CrossRef]
  7. Dey, S.; Tripathy, B.; Kumar, M.S.; Das, A.P. Ecotoxicological consequences of manganese mining pollutants and their biological remediation. Environ. Chem. Ecotoxicol. 2023, 5, 55–61. [Google Scholar] [CrossRef]
  8. Mishra, P.; Kaur, L.; Patel, R.; Mishra, R.K.; Verma, D.K.; Daoudi, W. Synthesis and characterization of halloysite-cerium nanocomposite for removal of manganese. J. Environ. Chem. Eng. 2024, 12, 114611. [Google Scholar] [CrossRef]
  9. Cheng, Y.; Zhang, Y.; Xiong, W.; Huang, T. Simultaneous removal of tetracycline and manganese (II) ions from groundwater using manganese oxide filters: Efficiency and mechanisms. J. Water Process Eng. 2021, 42, 102158. [Google Scholar] [CrossRef]
  10. Yang, H.; Tang, X.; Luo, X.; Li, G.; Liang, H.; Snyder, S. Oxidants-assisted sand filter to enhance the simultaneous removals of manganese, iron and ammonia from groundwater: Formation of active MnOx and involved mechanisms. J. Hazard. Mater. 2021, 415, 125707. [Google Scholar] [CrossRef] [PubMed]
  11. Luo, J.; Zhang, Y.; Chang, H.; Lin, C.; Hu, Y.; Wang, H.; Wang, Y.; Tang, X. Manganese Oxide Enhanced Gravity-Driven Membrane (GDM) Filtration in Treating Iron- and Manganese-Containing Surface Water. Water 2024, 16, 2374. [Google Scholar] [CrossRef]
  12. Lee, W.S.; Aziz, H.A.; Akbar, N.A.; Wang, M.-H.S.; Wang, L.K. Removal of Fe and Mn from Groundwater. In Industrial Waste Engineering; Springer International Publishing: New York, NY, USA, 2023; pp. 135–170. [Google Scholar]
  13. Ducret, J.; Barbeau, B. Modeling Mn(II) autocatalytic sorption on MnOx-coated filtration media. J. Water Process Eng. 2024, 62, 105408. [Google Scholar] [CrossRef]
  14. Taffarel, S.R.; Rubio, J. Removal of Mn2+ from aqueous solution by manganese oxide coated zeolite. Miner. Eng. 2010, 23, 1131–1138. [Google Scholar] [CrossRef]
  15. Jin, X.; Fu, J.; Yu, P.; Luo, D. Characterization and properties of manganese oxide film coated clinoptilolite as filter material in fixed-bed columns for removal of Mn(II) from aqueous solution. Sci. Rep. 2023, 13, 17440. [Google Scholar] [CrossRef] [PubMed]
  16. Zeng, J.; Xie, H.; Zhang, H.; Huang, M.; Liu, X.; Zhou, G.; Jiang, Y. Insight into the effects of oxygen vacancy on the toluene oxidation over α-MnO2 catalyst. Chemosphere 2022, 291, 132890. [Google Scholar] [CrossRef] [PubMed]
  17. Li, F.; Yin, H.; Zhu, T.; Zhuang, W. Understanding the role of manganese oxides in retaining harmful metals: Insights into oxidation and adsorption mechanisms at microstructure level. Eco-Environ. Health 2024, 3, 89–106. [Google Scholar] [CrossRef]
  18. Yang, Y.; Wang, Y.; Li, X.; Xue, C.; Dang, Z.; Zhang, L.; Yi, X. Effects of synthesis temperature on ε-MnO2 microstructures and performance: Selective adsorption of heavy metals and the mechanism onto (100) facet compared with (001). Environ. Pollut. 2022, 315, 120218. [Google Scholar] [CrossRef]
  19. Cheng, Y.; Huang, T.L.; Sun, Y.K.; Shi, X.X. Catalytic oxidation removal of ammonium from groundwater by manganese oxides filter: Performance and mechanisms. Chem. Eng. J. 2017, 322, 82–89. [Google Scholar] [CrossRef]
  20. Xie, X.; Lu, C.; Xu, R.; Yang, X.; Yan, L.; Su, C. Arsenic removal by manganese-doped mesoporous iron oxides from groundwater: Performance and mechanism. Sci. Total Environ. 2022, 806, 150615. [Google Scholar] [CrossRef] [PubMed]
  21. Yu, P.; Song, Y.; Jin, X.; Fu, J.; Zhang, S. Study on the efficiency of manganese oxide-bearing manganese sand for removing Mn2+ from aqueous solution. Microporous Mesoporous Mater. 2024, 364, 112859. [Google Scholar] [CrossRef]
  22. Junta, J.L.; Hochella, M.F. Manganese (II) oxidation at mineral surfaces: A microscopic and spectroscopic study. Geochim. Cosmochim. Acta 1994, 58, 4985–4999. [Google Scholar] [CrossRef]
  23. Scheitenberger, P.; Euchner, H.; Lindén, M. The hidden impact of structural water—How interlayer water largely controls the Raman spectroscopic response of birnessite-type manganese oxide. J. Mater. Chem. A 2021, 9, 18466–18476. [Google Scholar] [CrossRef]
  24. Feng, X.; Cox, D.F. Oxidation of MnO(100) and NaMnO2 formation: Characterization of Mn2+ and Mn3+ surfaces via XPS and water TPD. Surf. Sci. 2018, 675, 47–53. [Google Scholar] [CrossRef]
  25. Cai, T.; Liu, Z.; Yuan, J.; Xu, P.; Zhao, K.; Tong, Q.; Lu, W.; He, D. The structural evolution of MnOx with calcination temperature and their catalytic performance for propane total oxidation. Appl. Surf. Sci. 2021, 565, 150596. [Google Scholar] [CrossRef]
  26. Zaharieva, I.; Chernev, P.; Risch, M.; Klingan, K.; Kohlhoff, M.; Fischer, A.; Dau, H. Electrosynthesis, functional, and structural characterization of a water-oxidizing manganese oxide. Energy Environ. Sci. 2012, 5, 7081–7089. [Google Scholar] [CrossRef]
  27. Cheng, Y.; Xiong, W.; Huang, T. Mechanistic insights into effect of storage conditions of Fe-Mn co-oxide filter media on their catalytic properties in ammonium-nitrogen and manganese oxidative removal. Sep. Purif. Technol. 2021, 259, 118102. [Google Scholar] [CrossRef]
  28. Chaudhry, S.A.; Khan, T.A.; Ali, I. Adsorptive removal of Pb(II) and Zn(II) from water onto manganese oxide-coated sand: Isotherm, thermodynamic and kinetic studies. Egypt. J. Basic Appl. Sci. 2019, 3, 287–300. [Google Scholar] [CrossRef]
  29. Foo, K.Y.; Hameed, B.H. Insights into the modeling of adsorption isotherm systems. Chem. Eng. J. 2010, 156, 2–10. [Google Scholar] [CrossRef]
  30. Dutta, D.; Borah, J.P.; Puzari, A.; Valencia, S. Adsorption of Mn2+ from Aqueous Solution Using Manganese Oxide-Coated Hollow Polymethylmethacrylate Microspheres (MHPM). Adsorpt. Sci. Technol. 2021, 2021, 10. [Google Scholar] [CrossRef]
  31. Tang, N.; Niu, C.-G.; Li, X.-T.; Liang, C.; Guo, H.; Lin, L.-S.; Zheng, C.-W.; Zeng, G.-M. Efficient removal of Cd2+ and Pb2+ from aqueous solution with amino- and thiol-functionalized activated carbon: Isotherm and kinetics modeling. Sci. Total Environ. 2018, 635, 1331–1344. [Google Scholar] [CrossRef]
  32. Zhang, R.; Yang, S. Rapid start-up and pollutant removal mechanism of MnOx filter for simultaneous removal of manganese and ammonium. China Environ. Sci. 2023, 43, 197–205. [Google Scholar]
  33. Lan, S.; Wang, X.; Xiang, Q.; Yin, H.; Tan, W.; Qiu, G.; Liu, F.; Zhang, J.; Feng, X. Mechanisms of Mn(II) catalytic oxidation on ferrihydrite surfaces and the formation of manganese (oxyhydr)oxides. Geochim. Cosmochim. Acta 2017, 211, 79–96. [Google Scholar] [CrossRef]
  34. Zhou, F.; Izgorodin, A.; Hocking, R.K.; Armel, V.; Spiccia, L.; MacFarlane, D.R. Improvement of Catalytic Water Oxidation on MnOx Films by Heat Treatment. ChemSusChem 2013, 6, 643–651. [Google Scholar] [CrossRef] [PubMed]
  35. Dose, W.M.; Donne, S.W. Manganese dioxide structural effects on its thermal decomposition. Mater. Sci. Eng. B 2011, 176, 1169–1177. [Google Scholar] [CrossRef]
  36. Guo, X.; Wang, J. A general kinetic model for adsorption: Theoretical analysis and modeling. J. Mol. Liq. 2019, 288, 111100. [Google Scholar] [CrossRef]
  37. Peng, X.; Hu, F.; Huang, J.; Wang, Y.; Dai, H.; Liu, Z. Preparation of a graphitic ordered mesoporous carbon and its application in sorption of ciprofloxacin: Kinetics, isotherm, adsorption mechanisms studies. Microporous Mesoporous Mater. 2016, 228, 196–206. [Google Scholar] [CrossRef]
  38. Pickles, C.A.; Marzoughi, O. Thermodynamic modelling of decomposition processes in the Mn-O and Mn-O-H systems. Can. Metall. Q. 2022, 62, 151–170. [Google Scholar] [CrossRef]
  39. Duan, J.; Feng, S.; He, W.; Li, R.; Zhang, P.; Zhang, Y. TG-FTIR and Py-GC/MS combined with kinetic model to study the pyrolysis characteristics of electrolytic manganese residue. J. Anal. Appl. Pyrolysis 2021, 159, 105203. [Google Scholar] [CrossRef]
  40. Ramarajan, D.; Sivagurunathan, P.; Yan, Q. Synthesis and characterization of sol-processed α-MnO2 nanostructures. Mater. Sci. Semicond. Process. 2012, 15, 559–563. [Google Scholar] [CrossRef]
  41. Frey, C.E.; Kurz, P. Water Oxidation Catalysis by Synthetic Manganese Oxides with Different Structural Motifs: A Comparative Study. Chem.—A Eur. J. 2015, 21, 14958–14968. [Google Scholar] [CrossRef] [PubMed]
  42. Zhang, R.; Yang, S.; Dong, C.; Qiao, Y.; Zhang, J.; Guo, Y. Synthesized akhtenskites remove ammonium and manganese from aqueous solution: Removal mechanism and the effect of structural cations. RSC Adv. 2021, 11, 33798–33808. [Google Scholar] [CrossRef] [PubMed]
  43. Hu, E.; Zhang, Y.; Wu, S.; Wu, J.; Liang, L.; He, F. Role of dissolved Mn(III) in transformation of organic contaminants: Non-oxidative versus oxidative mechanisms. Water Res. 2017, 111, 234–243. [Google Scholar] [CrossRef]
  44. Jia, J.; Zhang, P.; Chen, L. Catalytic decomposition of gaseous ozone over manganese dioxides with different crystal structures. Appl. Catal. B: Environ. 2016, 189, 210–218. [Google Scholar] [CrossRef]
  45. Wang, Z.; Jia, H.; Zheng, T.; Dai, Y.; Zhang, C.; Guo, X.; Wang, T.; Zhu, L. Promoted catalytic transformation of polycyclic aromatic hydrocarbons by MnO2 polymorphs: Synergistic effects of Mn3+ and oxygen vacancies. Appl. Catal. B Environ. 2020, 272, 119030. [Google Scholar] [CrossRef]
  46. Dong, C.; Qu, Z.; Jiang, X.; Ren, Y. Tuning oxygen vacancy concentration of MnO2 through metal doping for improved toluene oxidation. J. Hazard. Mater. 2020, 391, 122181. [Google Scholar] [CrossRef] [PubMed]
  47. Huang, L.; Luo, X.; Chen, C.; Jiang, Q. A high specific capacity aqueous zinc-manganese battery with a ε-MnO2 cathode. Ionics 2021, 27, 3933–3941. [Google Scholar] [CrossRef]
  48. Wan, J.; Zhou, L.; Deng, H.; Zhan, F.; Zhang, R. Oxidative degradation of sulfamethoxazole by different MnO2 nanocrystals in aqueous solution. J. Mol. Catal. A Chem. 2015, 407, 67–74. [Google Scholar] [CrossRef]
  49. Garcês Gonçalves, P.R.; De Abreu, H.A.; Duarte, H.A. Stability, Structural, and Electronic Properties of Hausmannite (Mn3O4) Surfaces and Their Interaction with Water. J. Phys. Chem. C 2018, 122, 20841–20849. [Google Scholar] [CrossRef]
Figure 1. The removal performance of Mn2+ by MnOx dried at different (a) times and (b) temperatures.
Figure 1. The removal performance of Mn2+ by MnOx dried at different (a) times and (b) temperatures.
Water 17 00261 g001
Figure 2. Adsorption kinetics of Mn2+ by MnOx under different drying conditions: (a) pseudo-first order kinetic plots and (b) pseudo-second order kinetic plots (MnOx = 0.2 g/L, Mn2+ = 25 mg/L, T = 25 °C, t = 300 min).
Figure 2. Adsorption kinetics of Mn2+ by MnOx under different drying conditions: (a) pseudo-first order kinetic plots and (b) pseudo-second order kinetic plots (MnOx = 0.2 g/L, Mn2+ = 25 mg/L, T = 25 °C, t = 300 min).
Water 17 00261 g002
Figure 3. Adsorption isotherms of Mn2+ on MnOx under different drying conditions: (a) Langmuir isotherm and (b) Freundlich isotherm.
Figure 3. Adsorption isotherms of Mn2+ on MnOx under different drying conditions: (a) Langmuir isotherm and (b) Freundlich isotherm.
Water 17 00261 g003
Figure 4. The removal performance of Mn2+ by MnOx in the (a) reuse experiments and dynamic filtration test: (b) the influent and effluent Mn2+ concentrations with time, (c) the Mn2+ concentration along the filter depth with different influent concentrations, and (d) the Mn2+ concentration along the filter depth with different filtration rates.
Figure 4. The removal performance of Mn2+ by MnOx in the (a) reuse experiments and dynamic filtration test: (b) the influent and effluent Mn2+ concentrations with time, (c) the Mn2+ concentration along the filter depth with different influent concentrations, and (d) the Mn2+ concentration along the filter depth with different filtration rates.
Water 17 00261 g004
Figure 5. SEM images and the corresponding particle size distributions of MnOx under different drying conditions: (a,b) 50 °C-12 h-MnOx, (c,d) 50 °C-24 h-MnOx, (e,f) 120 °C-12 h-MnOx, (g,h) freeze-dried-12 h-MnOx.
Figure 5. SEM images and the corresponding particle size distributions of MnOx under different drying conditions: (a,b) 50 °C-12 h-MnOx, (c,d) 50 °C-24 h-MnOx, (e,f) 120 °C-12 h-MnOx, (g,h) freeze-dried-12 h-MnOx.
Water 17 00261 g005aWater 17 00261 g005b
Figure 6. TG-DTG curves of MnOx under different drying conditions: (a) 20 °C-12 h-MnOx, (b) 50 °C-12 h-MnOx, (c) 50 °C-24 h-MnOx, (d) 120 °C-12 h-MnOx, (e) freeze-dried-12 h-MnOx.
Figure 6. TG-DTG curves of MnOx under different drying conditions: (a) 20 °C-12 h-MnOx, (b) 50 °C-12 h-MnOx, (c) 50 °C-24 h-MnOx, (d) 120 °C-12 h-MnOx, (e) freeze-dried-12 h-MnOx.
Water 17 00261 g006aWater 17 00261 g006b
Figure 7. XPS spectra of (a) Mn 2p and (b) O1s in MnOx.
Figure 7. XPS spectra of (a) Mn 2p and (b) O1s in MnOx.
Water 17 00261 g007
Figure 8. Removal performance of Mn2+ by different MnOx with nitrogen and oxygen: (a) 50 °C-12 h-MnOx, (b) 50 °C-24 h-MnOx, (c) 120 °C-12 h-MnOx, (d) freeze-dried-12 h-MnOx. (MnOx = 0.2 g/L, Mn2+ = 200 mg/L, T = 25 °C, t = 24 h).
Figure 8. Removal performance of Mn2+ by different MnOx with nitrogen and oxygen: (a) 50 °C-12 h-MnOx, (b) 50 °C-24 h-MnOx, (c) 120 °C-12 h-MnOx, (d) freeze-dried-12 h-MnOx. (MnOx = 0.2 g/L, Mn2+ = 200 mg/L, T = 25 °C, t = 24 h).
Water 17 00261 g008
Table 1. Water quality of the influent water for the filter column.
Table 1. Water quality of the influent water for the filter column.
ParameterUnitValue
Temperature°C10.0–20.0
Dissolved oxygen concentrationmg/L7.40–8.10
pH value/7.61–8.23
TurbidityNTU0.12–0.34
CODMnmg/L0.90–1.50
TOCmg/L0.69–3.97
Alkalinity (as CaCO3)mg/L135.28–137.52
Table 2. Kinetic parameters of Mn2+ adsorption by MnOx under different drying conditions.
Table 2. Kinetic parameters of Mn2+ adsorption by MnOx under different drying conditions.
SamplePseudo-First-OrderPseudo-Second-Order
K1 (min−1)qe (mg/g)R12K2 (g·mg−1·min−1)qe (mg/g)R22
50 °C-1 h-MnOx0.003238.740.9690.0013110.50.996
50 °C-3 h-MnOx0.003537.590.9720.0014111.20.997
50 °C-12 h-MnOx0.014222.680.9490.0025125.70.999
50 °C-24 h-MnOx0.005935.990.9800.0012119.60.997
20 °C-12 h-MnOx0.007033.610.9600.0013121.50.998
80 °C-12 h-MnOx0.006831.390.9840.0014121.60.998
120 °C-12 h-MnOx0.007041.830.9840.0010115.20.995
Freeze-dried-12 h-MnOx0.007629.810.9720.0015122.50.999
Table 3. Adsorption isotherm parameters of Mn2+ by MnOx under different drying conditions.
Table 3. Adsorption isotherm parameters of Mn2+ by MnOx under different drying conditions.
SampleFreundlich Isotherm Model
KF (L/mg)1/nR2
50 °C-1 h-MnOx104.5 ± 6.700.152 ± 0.0160.948
50 °C-3 h-MnOx105.6 ± 8.090.161 ± 0.0200.940
50 °C-12 h-MnOx136.3 ± 9.650.134 ± 0.0190.968
50 °C-24 h-MnOx108.5 ± 10.80.172 ± 0.0260.943
20 °C-12 h-MnOx106.6 ± 6.490.160 ± 0.0160.973
80 °C-12 h-MnOx113.8 ± 10.20.154 ± 0.0230.940
120 °C-12 h-MnOx113.5 ± 11.30.153 ± 0.0270.959
Freeze-dried-12 h-MnOx110.4 ± 4.780.168 ± 0.0110.950
Table 4. The percentages of the components of Mn and O based on the XPS spectra.
Table 4. The percentages of the components of Mn and O based on the XPS spectra.
SampleMn 2p2/3O1s
Mn(II)
(%)
Mn(III)
(%)
Mn(IV)
(%)
Mn(III)
/Mn(IV)
Owater
(%)
Olatt
(%)
Oads
(%)
50 °C-12 h-MnOx19.6060.9019.493.1213.4440.5646.01
50 °C-24 h-MnOx10.9244.2244.860.9811.3954.9033.70
120 °C-12 h-MnOx8.7836.9554.270.6819.3345.4735.20
Freeze-dried-12 h-MnOx32.9446.5020.572.2612.4234.8135.60
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

Zhang, R.; Yang, L.; Yang, J.; Niu, Q.; Zhu, B. Effects of Drying Operation on the Mn2+ Removal Activity of MnOx: Performance and Mechanism. Water 2025, 17, 261. https://doi.org/10.3390/w17020261

AMA Style

Zhang R, Yang L, Yang J, Niu Q, Zhu B. Effects of Drying Operation on the Mn2+ Removal Activity of MnOx: Performance and Mechanism. Water. 2025; 17(2):261. https://doi.org/10.3390/w17020261

Chicago/Turabian Style

Zhang, Ruifeng, Lina Yang, Jing Yang, Qiuyan Niu, and Binrong Zhu. 2025. "Effects of Drying Operation on the Mn2+ Removal Activity of MnOx: Performance and Mechanism" Water 17, no. 2: 261. https://doi.org/10.3390/w17020261

APA Style

Zhang, R., Yang, L., Yang, J., Niu, Q., & Zhu, B. (2025). Effects of Drying Operation on the Mn2+ Removal Activity of MnOx: Performance and Mechanism. Water, 17(2), 261. https://doi.org/10.3390/w17020261

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