Next Article in Journal
Multivariate Analysis of Magnetic Parameters and Trace Metals in Atmospheric Dustfall and Its Environmental Implications in Northern China
Previous Article in Journal
Grid-Characteristic Method on Overlapping Curvilinear Meshes for Modeling Elastic Waves Scattering on Geological Fractures
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Development of Nanoscale Hydrated Titanium Oxides Support Anion Exchange Resin for Efficient Phosphate Removal from Water

College of Environmental Science and Engineering, Nanjing Tech University, Nanjing 211816, China
*
Author to whom correspondence should be addressed.
Submission received: 1 November 2022 / Revised: 1 December 2022 / Accepted: 7 December 2022 / Published: 12 December 2022
(This article belongs to the Section Environmental Mineralogy and Biogeochemistry)

Abstract

:
In this work, a macroporous strongly basic anion exchange resin D201 was used as the matrix and loaded with nano hydrated titanium oxide (HTO) to fabricate a novel resin-based nano hydrated titanium oxide adsorbent (HTO-D201), which was characterized by scanning electron microscope-energy dispersion spectroscopy (SEM-EDS), transmission electron microscope (TEM), X-ray diffraction (XRD) and Brunauer-Emmett-Teller (BET) analysis. Adsorption isotherm, pH influence competitive adsorption and column adsorption experiments were conducted to investigation the adsorption behavior of HTO-D201 to phosphorus in water. The adsorption effect of adsorbent HTO-D201 on phosphorus in water, and the corresponding adsorption mechanism, are discussed. It was observed that HTO-D201 exhibited spontaneous adsorption behavior with Langmuir fitting maximum adsorption capacity of 34.08 mg∙g−1 under a pH of 6.8 and a temperature of 298 K. Adsorption isotherms confirmed that enhancing temperature could promote the adsorption process. SO42−, NO3 and Cl were used as competing ions in competitive adsorption, which confirmed better anti-interference ability of HTO-D201 compared with that of unmodified D201. The column adsorption experiment implied that HTO-D201 possessed a stable structure and good dynamic adsorption performance, with effective processing capacity of 420BV, which could be regenerated and recycled. The adsorption mechanism of HTO-D201 to phosphorus in water is discussed, which was ascribed to a quaternary amine group on the resin and a hydroxyl group on the HTO. This work shows that HTO-D201 is a promising adsorbent that a possesses excellent phosphorus-removing capacity from wastewater and the potential for practical application.

1. Introduction

Phosphorus is an important part of nucleic acids and various metabolites, and participates in the metabolic reactions of all living organisms. Therefore, phosphorus plays an important role in biological systems and is one of the main nutrients for biological growth [1]. Phosphorus mainly exists in soil and water bodies in nature, and the concentration of phosphorus in natural water bodies is low [2]. With the intensification of human activities, phosphorus is being widely used in agriculture, the chemical industry and medicine. Due to the direct discharge of sewage after the use of high-phosphorus washing powder and the limited utilization of phosphorus in the soil, excessive phosphorus enters the environment and natural water bodies The concentration of phosphorus is increasing day by day [3]. When the phosphorus content exceeds 0.02 mg∙L−1, water eutrophication occurs [4]. Eutrophication of the water body causes the excessive growth of autotrophic organisms (such as algae), inhibits photosynthesis of deep aquatic plants, thereby reducing the dissolved oxygen content and leading to hypoxia. The number of aquatic organisms is drastically reduced, and the ecological balance is destroyed. This especially poses a serious threat to drinking water safety [5]. Therefore, it is necessary to develop an optimal method for removing phosphorus from wastewater.
At present, various methods have been proposed to remove phosphorus, mainly including biological treatment [6], crystallization [7], chemical precipitation [8], an aquatic organism method [9], and adsorption [10]. Biological treatment is currently widely used in the decomposition of high-concentration organic phosphorus [11]. However, the large amount of sludge produced in the biological process must be treated, with a high cost [12]. Researchers have used crystallization and chemical precipitation methods to remove soluble phosphates in wastewater [13]. However, both the crystallization method and the chemical precipitation method require the addition of a large amount of chemical reagents, and the operating conditions of these two methods are demanding. They are mainly suitable for the treatment of high-concentration phosphorous wastewater [14]. The aquatic organism method is only suitable for the restoration of natural water bodies with a lower cost but a longer treatment time [15,16]. The adsorption method is simple to operate, easy to control, and has few influencing factors. Choosing a suitable adsorbent can achieve a greater phosphorus removal effect that the above several methods. Therefore, it has been recognized and reported that the adsorption method is one of the best methods to remove phosphorus from wastewater [17,18].
So far, a variety of adsorbents for phosphorus removal have been developed, including ion exchange resins, zirconia, activated alumina, and iron oxide particles [19,20]. However, from the perspective of long-term use, some adsorbents (such as activated alumina, zirconia and iron oxide particles) lack mechanical strength and lose high adsorption activity after multiple regeneration cycles [21]. Therefore, among these various adsorbents, ion exchange resins have more application prospects. At present, nanomaterials have become excellent adsorption materials due to the advantages of small particle size, large specific surface area, and large numbers of surface atoms [22]. Research has found that certain nano-particle metals and their compounds, such as nano-titanium oxide, nano-zirconia, nano-iron oxide, and nano-alumina, are highly effective on phosphorus removal from wastewater through the coordination and complexation of coordination bonds. Such nano-particles possess excellent adsorbent properties, which are not easily affected by coexisting ions [23,24]. However, nanomaterials have the disadvantages of being small particles, easy to agglomerate, easy to float, having low mechanical strength, and being difficult to recycle and regenerate [25]. An ion exchange resin has the characteristics of high mechanical strength and excellent regeneration performance. Therefore, composite adsorbents that load nanomaterials onto the resin have emerged. These adsorbents have the advantages of both nanomaterials and ion exchange resins, and have become new research hotspots. Among them, although nano-zirconia has a high specific surface area, it has a higher adsorption capacity for phosphorus compared with other adsorbents. However, since the raw materials required to prepare mesoporous zirconium are very expensive, its commercial application is limited. Therefore, people have studied titanium as an alternative material because titanium is stronger, denser, and has a more stable mesoporous structure network than zirconium [26].
The purpose of this research was to develop high-efficiency dephosphorization nano-adsorbents that can be practically applied. Such as high-efficiency dephosphorization nano-adsorbent should have the advantages of simple preparation method, low material price, no ecological toxicity, and no secondary pollution. The adsorbent developed made full use of HTO’s strong adsorption capacity for phosphorus in wastewater, and at the same time solved the problems of HFO’s low mechanical strength, large water flow head loss, and difficulty in direct application to fixed beds or other fluid adsorption systems [27,28].
In this study, using inorganic titanium (TiCl4) as the metal source and D201 resin as the carrier, a composite adsorbent (HTO-D201) was prepared. HTO-D201 was characterized by scanning electron microscope-energy dispersion spectroscopy (SEM-EDS), transmission electron microscope (TEM), X-ray diffraction (XRD) and Brunauer-Emmett-Teller (BET) analysis. The Ti content in HTO-D201 was measured by an inductively coupled plasma direct reading spectrometer (ICP). The dephosphorization performance of D201 and HTO-D201 was evaluated and compared by adsorption isotherm experimental analysis. In addition, the effects of pH and competing ions on the adsorption of phosphorus by HTO-D201 were studied. A column adsorption experiment was carried out to study the actual performance and stability of HTO-D201. The potential adsorption mechanism was studied using X-ray photoelectron spectroscopy (XPS).

2. Materials and Methods

2.1. Materials

Titanium tetrachloride, sodium dihydrogen phosphate, ascorbic acid, molybdate, antimony tartrate, perchloric acid and hydrofluoric acid were purchased from Aladdin reagent (Shanghai, China). All reagents were analytically pure and were used directly without purification. D201 resin, purchased from Zhengguang Resin Co., Ltd., Ningbo, Zhejiang Province, China, has a particle size of 0.5–1 mm, a quaternary ammonium content of 2–5 mmol·g−1, and a pore diameter of 2–100 nm. Before use, it was rinsed with hydrochloric acid (HCl), sodium hydroxide (NaOH), and deionized water in sequence, and passed through an ethanol permeation treatment. Finally, D-201 was dried until it reached a constant weight.

2.2. Synthesis of HTO-D201

HTO-D201 was prepared by ion exchange-in-situ deposition [29]. The fabrication process of HTO-D 201 is shown in Figure 1. Firstly, 15 mL TiCl4 and 15 g D201 were added to 250 mL HCl-absolute ethanol solution, and the mixture was stirred at room temperature for 24 h, then filtered and dried. The TiCl5-complex anion was formed by TiCl4 in a high-concentration of hydrochloric acid solution, and combined with D201 through electrostatic adsorption and ion exchange. Secondly, the dried resin was added to a mixture of 300 mL NaCl and NaOH (the mass fraction of both NaCl and NaOH was 5.00%), and the resin was precipitated with an alkaline solution and then filtered. The adsorbent was washed repeatedly with deionized water and ethanol to pH of 7, and cured at 60 °C. Finally, the titanium is left on D201 in the form of HTO to obtain HTO-D201 [30].

2.3. Characterization of HTO-D201

The surface and cross section of the material were scanned by SEM(S-3400N II, Hitachi, Marunouchi, Japan) to observe the surface and profile morphology of HTO-D201. EDS (Ex-250, Horiba, Kyoto, Japan) scanning was used to analyze the distribution of each element on the HTO-D201 cross section. TEM (Hitachi Model H-800, Hitachi, Japan) was used to observe the morphology of HTO dispersed in resin holes. The crystal morphology of HTO was analyzed by XRD (X’TRA, ARL, Ecublens, Switzerland). The pore size distribution and specific surface area of HTO-D201 were measured using nitrogen as an adsorption-desorption gas at 77 K using a BET surface analyzer (ASAP-2010C, Micromeritics, Norcross, GA, USA). After a certain amount of HTO-D201 was digested by reflux heating with nitric acid and hydrofluoric acid, and diluted, the content of Ti element in the solution was measured by ICP (Optima 5300DV, Perkinelmer, Shelton, CT, USA) using an inductively coupled plasma direct reading spectrometer. After conversion, the mass fraction of Ti on HTO-D201 was obtained. The properties of D201 and HTO-D201 after phosphorus adsorption were analyzed by XPS spectroscopy (Phi 5000 Versa Probe, Japan ULVAC-Phi Corporation, Toki, Japan).

2.4. Batch Adsorption Experiment and Analysis Method

HTO-D201 (25.00 mg) and 25.00 mg D201 were weighed and added into 50 mL of simulated phosphorus (we used orthophosphates) containing wastewater with mass concentrations of 2.00, 4.00, 10.00, 20.00, 30.00, and 40.00 mg·L−1, respectively. HTO-D201 was adsorbed in a constant-temperature shaking water bath at three different temperatures of 288 298 and 308 K for 24 h. D201 was adsorbed in a constant temperature water bath shaker for 24 h at a temperature of 298 K, and the mass concentration of phosphorus in the simulated wastewater after adsorption was measured. HTO-D201 (25.00 mg) was weighed and added into two groups of 50 mL simulated phosphor-containing wastewater with different pH values; the mass concentration of phosphorus was 15.00 mg·L−1. One of the groups of simulated wastewater was added to a certain concentration of SO42− as a competitive ion, and the pH value was adjusted with HNO3 and NaOH solutions. The mass concentration of phosphorus in the simulated wastewater was determined after 24 h adsorption in a constant-temperature water bath oscillator at 298 K. HTO-D201 (25.00 mg) and 25.00 mg D201 were weighed and added to 50 mL of simulated phosphorus-containing wastewater with different mass concentrations of SO42−, NO3 and Cl. The mass concentration of phosphorus was 10.00 mg·L−1, and the pH value of the solution was 5.80. The mass concentration of phosphorus in the simulated wastewater was determined after 24 h adsorption in a constant-temperature water bath oscillator at 298 K. The effects of SO42−, NO3 and Cl on the adsorption of phosphorus by HTO-D201 and D201 were compared. The column adsorption experiment was conducted as shown in Figure 2. A glass column with an inner diameter of 14 mm and a column length of 120 mm was filled with a certain volume of HTO-D201, and a peristaltic pump was used to ensure that the simulated wastewater containing phosphorus passed through the adsorption column at a constant rate. The mass concentration of phosphorus in the influent was 2.00 mg·L−1, and SO42−, NO3 and Cl, with mass concentrations of 50.00 mg·L−1, were contained in the influent as competing ions. The contact time of the empty bed of the adsorption column was an EBTC of 6 min. An automatic sampler was used for continuous automatic sampling, and the mass concentration of phosphorus was measured. When the adsorption material was saturated, the HTO-D201 was desorbed with a mixed solution of 5.00% NaCl + 5.00% NaOH. The desorbed HTO-D201 was rinsed with 1% NaCl solution and deionized water until the pH of the effluent was neutral and entered the next round of adsorption. At the same time, the carrier D201 was used as the adsorption material to do a comparative experiment under the same operating conditions. The detection of phosphorus in water adopted the national standard “Ammonium Molybdate Spectrophotometry” (GB11893-89). The detection limit of this method for phosphorus is 0.01 mg·L−1, and the detection upper limit is 0.60 mg·L−1 [31].

3. Results and Discussion

3.1. Characterization of HTO-D201

From Figure 3a, it can be seen that after loading Ti, the resin still maintained the spherical morphology of the D201 carrier, with a diameter of about 1 mm, thus ensuring its good hydraulic properties during use. HTO-D201 exhibited a dense porous structure inside (Figure 3b). Such structure possesses a large specific surface area, and can provide many adsorption active sites, thus enhancing the adsorption capability of HTO-D201 to phosphorus [32]. Figure 3c shows a distribution diagram of the Ti element in the radial direction of the HTO-D201 section. It can be seen from the figure that Ti elements occur along the inner diameter of the resin, indicating that Ti was successfully loaded on D201, and Ti was more distributed in the outer layer of the resin carrier, and relatively less inside, which is convenient for the contact reaction of HTO and pollutants [33]. Figure 3d shows the distribution of the P element in the radial direction of the section. After HTO-D201 adsorbed P, it was evenly distributed in the radial direction. It can be seen that the HTO-loaded configuration does not cause blockage of the resin pores. The macroporous structure of the resin carrier facilitates the entry of phosphate ions into the inside of the adsorbent. Transmission electron microscopy was used to observe the morphology of HTO in D201. The black particles in Figure 3e are HTO, which exhibit] irregular morphology and are uniformly dispersed on the surface of the D201 resin at nano size. To further investigate the particle size distribution of HTO nanoparticles, statistical analysis was performed for TEM images of HTO-D 201. As shown in Figure 3f, HTO nano-particles on HTO-D 201 had a maximum frequency particle size in the range 20–30 nm. Nano-sized particles have a large specific surface area that can expose many adsorption active sites to the outside [34,35].
XRD is a useful tool for investigation of the phase structure of materials [36]. Figure 4 shows the XRD pattern of D201 and HTO-D201. The broad peak at around 18.6° is ascribed to the crystal structure of polystyrene in the D201 matrix. Compared with the XRD pattern of the D201 carrier, there was no sharp diffraction peak in HTO-D201, indicating that Ti was loaded on D201 in a more active and unsteady state [37]. Table 1 compares the basic properties of D201 and HTO-D201. HTO-D201 used D201 as a carrier, so the basic skeletons of both were polystyrene. After loading Ti, the adsorption active site of HTO-D201 was not only the quaternary ammonium group of carrier D201, but also HTO [38]. The specific surface area of HTO-D201 measured by a specific surface area meter was 27.52 m2·g−1, which was higher than the 23.61 m2·g−1 of D201, consistent with TEM results. This phenomenon arise from the existence of Ti in the form of HTO, and can significantly enhance specific surface area of HTO-D 201. The ICP results confirmed that HTO-D201 was had a Ti content of 12.80% in HTO-D 201, compared with zero of D 201.

3.2. Adsorption Isotherm Experiment

The experimental results of adsorption isotherms are shown in Figure 5a,b. It can be seen that the equilibrium adsorption capacity of HTO-D201 increased with the increase of temperature. Compared with D201, the phosphorus adsorption capacity of HTO-D201 was greater. The fitting results of the Langmuir and Freundlich adsorption isotherm models are shown in Table 2. Both the K value of the Freundlich model and the qmax value of the Langmuir model increased with an increase of temperature. The reciprocal of the n value obtained by Freundlich model fitting was between 0.1 and 0.5, indicating that HTO-D201 has a strong adsorption capacity [39]. It can be seen that the K value, n value and b value of HTO-D201 were all higher than D201. Therefore, the adsorption capacity of HTO-D201 for phosphorus is stronger than that of D201 [40].
The calculated thermodynamic parameters are listed in Table 3. The Gibbs free energy change (∆G) of the adsorption reaction of HTO-D201 at three different temperatures was less than D201. This shows that the degree of spontaneous adsorption of HTO-D201 was higher than that of D201, and the higher the temperature, the greater the absolute value of free energy change, and the higher the degree of spontaneity [41] This phenomenon is probably attributable to enhanced diffusion of phosphate anions into the smaller micropores of HTO-D201 and thus accessing more binding sites. The adsorption enthalpy change (ΔH) of HTO-D201 was greater than zero, indicating that the process of adsorption of phosphorus is an endothermic process. The higher the temperature, the better the adsorption. This conclusion was also supported by the isotherm experiment results. The entropy change (ΔS) of the adsorption process was positive, indicating that the adsorption of phosphorus on HTO-D201 is an entropy-driven process, and the disorder of the solid-liquid contact surface increases [42].
Table 4 shows the comparison of the phosphorus adsorption performance of HTO-D201 and the supported adsorption materials that have been studied [43,44], It can be seen from the table that the adsorption capacity of HTO-D201 for phosphorus is greater than that of other adsorption materials. In addition, the magnesium oxyhydroxide modified resin material HMO-PN has a larger adsorption capacity, which shows the superiority of the resin as a carrier material.

3.3. The Effect of pH on Phosphorus Adsorption

It can be seen from Figure 6 that when the pH was between 6 and 7, the adsorption effect of HTO-D201 on phosphorus was the best. As the pH continued to rise, the adsorption capacity began to drop continuously. With SO42− shielding the role of carrier D201, the lower the pH value, the better the adsorption effect.
Under acidic conditions, phosphorus mainly exists in the form of H3PO4 and H2PO4 (Figure 1), and as the pH value increases, the proportion of negatively charged H2PO4 starts to rise [45]. The Donnan film effect makes it easier to concentrate on the positively charged D201 surface, increasing the possibility of contact between phosphorus and HTO [46]. When the pH continues to rise, the negative charge generated by the deprotonation of HTO repels the phosphate ion, and the increasing number of OH ions in the solution compete with the phosphate ion for the ion exchange sites on the D201 carrier. As a result, the adsorption capacity continues to decrease [47]. After using the SO42− shielding carrier D201, the adsorption capacity decreased as the pH increased. This is mainly because under acidic conditions, HTO is positively charged due to protonation, making it easier to adsorb phosphate. Under alkaline conditions, due to deprotonation, electrostatic repulsion is formed between HTO and phosphate ions, so when the pH value was greater than 8, there was almost no adsorption effect [48].

3.4. The Effect of Coexisting Ions on Phosphorus Adsorption

Figure 7 shows the equilibrium adsorption capacity of D201 and HTO-D201 under different molecular ratios of competitive ions and P, also considering variance [49]. As the concentration of competing ions increased, the equilibrium adsorption capacity of D201 gradually decreased and approached zero, while the equilibrium adsorption capacity of HTO-D201 also decreased, but could still maintain a certain adsorption effect, indicating that the specificity of HTO-D201 for phosphorus selection was stronger than that of the carrier D201 [50].
The decrease in equilibrium adsorption capacity of HTO-D201 occurred because the adsorption capacity of D201 was inhibited by competing ions. The main adsorption effect is due to Ti, which accounts for only 12.80% of the mass of HTO-D201. In addition, the existence of competing ions makes the Donnan membrane effect of D201 unable to preferentially enrich phosphate, and the collision probability of titanium and phosphate in the resin pores is reduced, which also affects the equilibrium adsorption capacity of HTO [51,52]. Comparing the influence of SO42−, NO3 and Cl on the ability of HTO-D201 to adsorb phosphorus, it was found that SO42− had the greatest influence, followed by NO3 and finally Cl. This is consistent with D201’s ability to adsorb anions. The negatively divalent SO42− has more charges than the negatively monovalent NO3 and Cl, and the coulombic force between D201 and D201 is stronger, making it is easier to adsorb to D201 [53].

3.5. Column Adsorption Experiment

As shown in Figure 8, when the first batch was adsorbed, HTO-D201 was in the first 300 BV, and could maintain a good adsorption effect. After 300 BV, the phosphorus concentration of the effluent began to rise significantly, and at about 320 BV, the phosphorus concentration of the effluent reached 0.20 mg·L−1. At 360 BV, the effluent phosphorus concentration exceeded 0.50 mg·L−1, and when D201 was 40 BV, the effluent phosphorus concentration exceeded 0.50 mg·L−1. When adsorption was saturated, the processing capacity of HTO-D201 and D201 were 420 BV and 100 BV, respectively. It can be seen that the saturated adsorption capacity of HTO-D201 was much larger than that of D201. After two adsorptions and two desorptions, the adsorption capacity of the third batch did not decrease significantly, indicating that HTO-D201 has a good regeneration capacity [54].

3.6. The Adsorption Mechanism of Phosphorus on HTO-D201

D 201 and HTO-D 201 after phosphorus adsorption were investigated by XPS to discover the mechanism of action of HTO-D 201. Figure 9a shows the XPS total spectrum of HTO-D201 with phosphorus adsorbed. Characteristic absorption peaks of titanium and phosphorus appear in the figure, indicating that titanium was successfully loaded on D201 and phosphorus was adsorbed. As observed from Figure 9b, D 201-P exhibited single peaks at around 132.9 eV, which was ascribed to exchange of Cl on the quaternary ammonium group and H2PO4 in the solution [55]. Ti-D201-P exhibited two binding peaks at around 132.90 eV and 132.30 eV. The peaks at 132.90 eV were ascribed to phosphorus absorbed by the D201 matrix, and the peak at 132.30 eV corresponded to complex coordination between phosphate and HTO. All of the results were highly consistent with a previous report, confirming the existence of HTO [56] XPS research results showed that after titanium was loaded, the number of active sites for phosphorus adsorption by the material increased, both the carrier and HTO had adsorption effects on phosphorus, and the adsorption process was more complicated [57].
HTO-D201 combined the advantages of D201 and HTO to achieve complementary advantages and enhanced functions. HTO-D201 has two types of functional groups: the quaternary amine group on the resin, and the hydroxyl group on the HTO [58,59]. Phosphate ions in the solution accumulated on the surface of HTO-D201 due to the pre-concentration effect of the Donnan membrane. This process was rapid and continuous throughout the adsorption process. After entering HTO-D201, the phosphate ion displaced Cl on the quaternary amine group and the OH on the HTO group, so as to be adsorbed on HTO-D201 (Figure 10) [60]. The ion exchange interaction with Cl was not selective, and its effect was weakened in the presence of competing ions, while the complex pairing interaction with HTO was highly selective, which ensured the selective adsorption performance of HTO-D201 adsorbents [61,62].

4. Conclusions

In this work, nano hydrated titanium oxide (HTO) was loaded onto D201 resin by an ion exchange-in-situ precipitation method to prepare a novel resin-based nano-hydrated titanium oxide adsorbent (HTO-D201) that was used for efficient dephosphorization in water. SEM results showed that HTO-D201 retained a porous structure with HTO particles uniformly within the carrier material. TEM images implied that HTO particles were uniformly loaded onto the D201 resin, which provided a larger specific surface area and active sites for dephosphorization. Adsorption isotherm experiments and thermodynamic analysis showed that the phosphorus adsorption on HTO-D201 was more spontaneous than that on D201, and the adsorption capacity of HTO-D201 was also much higher than that of D201. HTO-D201 exhibited better adsorption effect in low pH condition. A competitive adsorption test confirmed that HTO-D201 still maintained adsorption ability in high concentrations of competing ions, compared with loss of adsorption ability of D201. After regeneration, HTO-D201 exhibited significantly greater adsorption compared with that of D201, suggesting excellent reusability of HTO-D 201. XPS analysis showed that phosphorus adsorbed on HTO-D201 resulted in two different binding energies, suggesting two kinds of adsorption sites for phosphorus: the quaternary ammonium group in the D201 matrix and the hydroxyl group on the HTO particles. The mechanism of action of THO-D201 was discussed. This work provides a novel strategy for phosphate adsorbent with practical application prospects.

Author Contributions

Methodology, Y.Z.; Formal analysis, M.Z.; Investigation, M.Z., Y.T., D.W. and J.Z.; Resources, M.Z., D.W. and Y.Z.; Writing—original draft, M.Z. and Y.T.; Writing—review & editing, Z.L.; Project administration, Z.L.; Funding acquisition, Z.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Key Research and Development Project of China during the 13th Five-Year Plan Period (2017YFB0602505); the University Science Research Project of Jiangsu Province (16KJA610002).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wilfert, P.; Kumar, P.S.; Korving, L.; Witkamp, G.; van Loosdrecht, M.C.M. The Relevance of Phosphorus and Iron Chemistry to the Recovery of Phosphorus from Wastewater: A Review. Environ. Sci. Technol. 2015, 49, 9400–9414. [Google Scholar] [CrossRef] [PubMed]
  2. Feng, C.; Zhang, S.; Wang, Y.; Wang, G.; Pan, X.; Zhong, Q.; Xu, X.; Luo, L.; Long, L.; Yao, P. Synchronous removal of ammonium and phosphate from swine wastewater by two agricultural waste based adsorbents: Performance and mechanisms. Bioresour. Technol. 2020, 307, 123231. [Google Scholar] [CrossRef] [PubMed]
  3. Wang, Y.; Zhao, W.; Tian, X.; Song, H.; Gao, R.; Tang, X.; Zhang, X.; Hao, Y.; Tang, Y. High-efficiency recognition and detection of sulindac in sewage using hydrophilic imprinted resorcinol-formaldehyde resin magnetic nano-spheres as SPE adsorbents combined with HPLC. Chem. Eng. J. 2020, 392, 123716. [Google Scholar] [CrossRef]
  4. Yang, C.; Chang, J.; Lee, D. Covalent organic framework EB-COF:Br as adsorbent for phosphorus (V) or arsenic (V) removal from nearly neutral waters. Chemosphere 2020, 253, 126736. [Google Scholar] [CrossRef]
  5. Egle, L.; Rechberger, H.; Zessner, M. Overview and description of technologies for recovering phosphorus from municipal wastewater. Resour. Conserv. Recycl. 2015, 105, 325–346. [Google Scholar] [CrossRef]
  6. Sukacova, K.; Trtilek, M.; Rataj, T. Phosphorus removal using a microalgal biofilm in a new biofilm photobioreactor for tertiary wastewater treatment. Water Res. 2015, 71, 55–63. [Google Scholar] [CrossRef]
  7. Zou, H.; Wang, Y. Phosphorus removal and recovery from domestic wastewater in a novel process of enhanced biological phosphorus removal coupled with crystallization. Bioresour. Technol. 2016, 211, 87–92. [Google Scholar] [CrossRef]
  8. Zhou, Z.; Hu, L.; Ren, W.; Zhao, Y.; Jiang, L.; Wang, L. Effect of humic substances on phosphorus removal by struvite precipitation. Chemosphere 2015, 141, 94–99. [Google Scholar] [CrossRef]
  9. Liu, S.; Zhu, Y.; Meng, W.; He, Z.; Feng, W.; Zhang, C.; Giesy, J.P. Characteristics and degradation of carbon and phosphorus from aquatic macrophytes in lakes: Insights from solid-state C-13 NMR and solution P-31 NMR spectroscopy. Sci. Total Environ. 2016, 543, 746–756. [Google Scholar] [CrossRef]
  10. Lalley, J.; Han, C.; Li, X.; Dionysiou, D.D.; Nadagouda, M.N. Phosphate adsorption using modified iron oxide-based sorbents in lake water: Kinetics, equilibrium, and column tests. Chem. Eng. J. 2016, 284, 1386–1396. [Google Scholar] [CrossRef]
  11. Qiu, G.; Law, Y.; Das, S.; Ting, Y. Direct and Complete Phosphorus Recovery from Municipal Wastewater Using a Hybrid Microfiltration-Forward Osmosis Membrane Bioreactor Process with Seawater Brine as Draw Solution. Environ. Sci. Technol. 2015, 49, 6156–6163. [Google Scholar] [CrossRef] [PubMed]
  12. Nancharaiah, Y.V.; Mohan, S.V.; Lens, P.N.L. Recent advances in nutrient removal and recovery in biological and bioelectrochemical systems. Bioresour. Technol. 2016, 215, 173–185. [Google Scholar] [CrossRef] [PubMed]
  13. Cieslik, B.; Konieczka, P. A review of phosphorus recovery methods at various steps of wastewater treatment and sewage sludge management. The concept of "no solid waste generation" and analytical methods. J. Clean Prod. 2017, 142, 1728–1740. [Google Scholar] [CrossRef]
  14. Peng, L.; Dai, H.; Wu, Y.; Peng, Y.; Lu, X. A comprehensive review of phosphorus recovery from wastewater by crystallization processes. Chemosphere 2018, 197, 768–781. [Google Scholar] [CrossRef]
  15. Shen, Z.; Dong, X.; Shi, J.; Ma, Y.; Liu, D.; Fan, J. Simultaneous removal of nitrate/phosphate with bimetallic nanoparticles of Fe coupled with copper or nickel supported on chelating resin. Environ. Sci. Pollut. Res. 2019, 26, 16568–16576. [Google Scholar] [CrossRef]
  16. Naghipour, D.; Taghavi, K.; Jaafari, J.; Mahdavi, Y.; Ghozikali, M.G.; Ameri, R.; Jamshidi, A.; Mahvi, A.H. Statistical modeling and optimization of the phosphorus biosorption by modified Lemna minor from aqueous solution using response surface methodology (RSM). Desalin. Water Treat. 2016, 57, 19431–19442. [Google Scholar] [CrossRef]
  17. Wei, L.; Wang, G.; Jiang, J.; Li, G.; Zhang, X.; Zhao, Q.; Cui, F. Co-removal of phosphorus and nitrogen with commercial 201 × 7 anion exchange resin during tertiary treatment of WWTP effluent and phosphate recovery. Desalin. Water Treat. 2015, 56, 1633–1647. [Google Scholar] [CrossRef]
  18. Paudyal, H.; Inoue, K.; Ohto, K.; Kawakita, H.; Harada, H. Recovery of Phosphorus from Incineration Ash of Chicken Droppings by Citric Acid Leaching Followed by Adsorption Using a Porous Resin Containing Hydrated Zirconium Oxide Powder. J. Chem. Eng. JPN 2019, 52, 465–470. [Google Scholar] [CrossRef]
  19. Li, Q.; Li, X.; Sun, J.; Song, H.; Wu, J.; Wang, G.; Li, A. Removal of organic and inorganic matters from secondary effluent using resin adsorption and reuse of desorption eluate using ozone oxidation. Chemosphere 2020, 251, 126442. [Google Scholar] [CrossRef]
  20. Jin, J.; Feng, T.; Ma, Y.; Wang, W.; Wang, Y.; Zhou, Q.; Li, A. Novel magnetic carboxyl modified hypercrosslinked resins for effective removal of typical PPCPs. Chemosphere 2017, 185, 563–573. [Google Scholar] [CrossRef]
  21. Zheng, S.; Li, X.; Zhang, X.; Wang, W.; Yuan, S. Effect of inorganic regenerant properties on pharmaceutical adsorption and desorption performance on polymer anion exchange resin. Chemosphere 2017, 182, 325–331. [Google Scholar] [CrossRef] [PubMed]
  22. Li, Q.; Wang, Z.; Li, Q.; Shuang, C.; Zhou, Q.; Li, A.; Gao, C. Competition and enhancement effect in coremoval of atenolol and copper by an easily regenerative magnetic cation exchange resin. Chemosphere 2017, 179, 1–9. [Google Scholar] [CrossRef] [PubMed]
  23. Zhu, X.; Li, W.; Tang, S.; Zeng, M.; Bai, P.; Chen, L. Selective recovery of vanadium and scandium by ion exchange with D201 and solvent extraction using P507 from hydrochloric acid leaching solution of red mud. Chemosphere 2017, 175, 365–372. [Google Scholar] [CrossRef] [PubMed]
  24. Wu, B.; Wan, J.; Zhang, Y.; Pan, B.; Lo, I.M.C. Selective Phosphate Removal from Water and Wastewater using Sorption: Process Fundamentals and Removal Mechanisms. Environ. Sci. Technol. 2020, 54, 50–66. [Google Scholar] [CrossRef] [PubMed]
  25. Choi, J.; Choi, Y.; Kim, D.; Lee, S. Characteristics of phosphorus adsorption for a titanium mesostructure synthesized with various surfactants. Environ. Technol. 2011, 32, 1419–1425. [Google Scholar] [CrossRef] [PubMed]
  26. Liu, X.; Chen, G.; Erwin, J.G.; Adam, N.K.; Su, C. Release of phosphorous impurity from TiO2 anatase and rutile nanoparticles in aquatic environments and its implications. Water Res. 2013, 47, 6149–6156. [Google Scholar] [CrossRef]
  27. Luo, H.; Xie, Y.; Niu, J.; Xiao, Y.; Li, Y.; Wang, Y.; Zhang, Y.; Xie, T. Cobalt-doped biogenic manganese oxides for enhanced tetracycline degradation by activation of peroxymonosulfate. J. Chem. Technol. Biot. 2019, 94, 752–760. [Google Scholar] [CrossRef]
  28. Wu, J.; Li, Q.; Li, W.; Li, Y.; Wang, G.; Li, A.; Li, H. Efficient removal of acid dyes using permanent magnetic resin and its preliminary investigation for advanced treatment of dyeing effluents. J. Clean Prod. 2020, 251, 119694. [Google Scholar] [CrossRef]
  29. Liu, Q.; Hu, P.; Wang, J.; Zhang, L.; Huang, R. Phosphate adsorption from aqueous solutions by Zirconium (IV) loaded cross-linked chitosan particles. J. Taiwan Inst. Chem. Eng. 2016, 59, 311–319. [Google Scholar] [CrossRef]
  30. Chen, L.; Zhao, X.; Pan, B.; Zhang, W.; Hua, M.; Lv, L.; Zhang, W. Preferable removal of phosphate from water using hydrous zirconium oxide-based nanocomposite of high stability. J. Hazard. Mater. 2015, 284, 35–42. [Google Scholar] [CrossRef]
  31. Nie, G.; Wu, L.; Du, Y.; Wang, H.; Xu, Y.; Ding, Z.; Liu, Z. Efficient removal of phosphate by a millimeter-sized nanocomposite of titanium oxides encapsulated in positively charged polymer. Chem. Eng. J. 2019, 360, 1128–1136. [Google Scholar] [CrossRef]
  32. Wan, B.; Yan, Y.; Liu, F.; Tan, W.; He, J.; Feng, X. Surface speciation of myo-inositol hexakisphosphate adsorbed on TiO2 nanoparticles and its impact on their colloidal stability in aqueous suspension: A comparative study with orthophosphate. Sci. Total Environ. 2016, 544, 134–142. [Google Scholar] [CrossRef]
  33. Wang, Y.; Xie, X.; Chen, X.; Huang, C.; Yang, S. Biochar-loaded Ce3+-enriched ultra-fine ceria nanoparticles for phosphate adsorption. J. Hazard. Mater. 2020, 396, 122626. [Google Scholar] [CrossRef]
  34. Xiao, G.; Wen, R.; Liu, A.; He, G.; Wu, D. Adsorption performance of salicylic acid on a novel resin with distinctive double pore structure. J. Hazard. Mater. 2017, 329, 77–83. [Google Scholar] [CrossRef] [PubMed]
  35. Rajendran, S.; Inwati, G.K.; Yadav, V.K.; Choudhary, N.; Solanki, M.B.; Abdellattif, M.H.; Yadav, K.K.; Gupta, N.; Islam, S.; Jeon, B.H. Enriched catalytic activity of TiO2 nanoparticles supported by activated carbon for noxious pollutant elimination. Nanomaterials 2021, 11, 2808. [Google Scholar] [CrossRef] [PubMed]
  36. Kumar, P.; Inwati, G.K.; Mathpal, M.C.; Ghosh, S.; Roos, W.D.; Swart, H.C. Defects induced enhancement of antifungal activities of Zn doped CuO nanostructures. Appl. Surf. Sci. 2021, 560, 150026. [Google Scholar] [CrossRef]
  37. Zhang, Y.; He, F.; Xia, S.; Kong, L.; Xu, D.; Wu, Z. Adsorption of sediment phosphorus by porous ceramic filter media coated with nano-titanium dioxide film. Ecol. Eng. 2014, 64, 186–192. [Google Scholar] [CrossRef]
  38. Wang, S.; Xiao, K.; Mo, Y.; Yang, B.; Vincent, T.; Faur, C.; Guibal, E. Selenium(VI) and copper(II) adsorption using polyethyleneimine-based resins: Effect of glutaraldehyde crosslinking and storage condition. J. Hazard. Mater. 2020, 386, 121637. [Google Scholar] [CrossRef]
  39. Kausar, A.; Iqbal, M.; Javed, A.; Aftab, K.; Nazli, Z.; Bhatti, H.N.; Nouren, S. Dyes adsorption using clay and modified clay: A review. J. Mol. Liq. 2018, 256, 395–407. [Google Scholar] [CrossRef]
  40. Lee, K.; Jutidamrongphan, W.; Lee, S.; Park, K.Y. Adsorption kinetics and isotherms of phosphate and its removal from wastewater using mesoporous titanium oxide. Membr. Water Treat. 2017, 8, 161–169. [Google Scholar] [CrossRef]
  41. Mitrogiannis, D.; Psychoyou, M.; Baziotis, I.; Inglezakis, V.J.; Koukouzas, N.; Tsoukalas, N.; Palles, D.; Kamitsos, E.; Oikonomou, G.; Markou, G. Removal of phosphate from aqueous solutions by adsorption onto Ca(OH)(2) treated natural clinoptilolite. Chem. Eng. J. 2017, 320, 510–522. [Google Scholar] [CrossRef]
  42. Wang, N.; Feng, J.; Chen, J.; Wang, J.; Yan, W. Adsorption mechanism of phosphate by polyaniline/TiO2 composite from wastewater. Chem. Eng. J. 2017, 316, 33–40. [Google Scholar] [CrossRef]
  43. Robshaw, T.; Tukra, S.; Hammond, D.B.; Leggett, G.J.; Ogden, M.D. Highly efficient fluoride extraction from simulant leachate of spent potlining via La-loaded chelating resin. An equilibrium study. J. Hazard. Mater. 2019, 361, 200–209. [Google Scholar] [CrossRef] [PubMed]
  44. Li, H.; Chen, Y.; Long, J.; Jiang, D.; Liu, J.; Li, S.; Qi, J.; Zhang, P.; Wang, J.; Gong, J.; et al. Simultaneous removal of thallium and chloride from a highly saline industrial wastewater using modified anion exchange resins. J. Hazard. Mater. 2017, 333, 179–185. [Google Scholar] [CrossRef]
  45. Tan, W.; Li, G.; Wang, T.; Yang, M.; Peng, J.; Barrow, C.J.; Yang, W.; Wang, H. Preparation and adsorption of phosphorus by new heteropolyacid salt-lanthanum oxide composites. Desalin. Water Treat. 2016, 57, 7874–7880. [Google Scholar]
  46. Deng, Z.; Oraby, E.A.; Eksteen, J.J. Cu adsorption behaviours onto chelating resins from glycine-cyanide solutions: Isotherms, kinetics and regeneration studies. Sep. Purif. Technol. 2020, 236, 116280. [Google Scholar] [CrossRef]
  47. Nur, T.; Loganathan, P.; Kandasamy, J.; Vigneswaran, S. Phosphate Adsorption from Membrane Bioreactor Effluent Using Dowex 21K XLT and Recovery as Struvite and Hydroxyapatite. Int. J. Environ. Res. Pub. Health 2016, 13, 277. [Google Scholar] [CrossRef] [Green Version]
  48. Lai, L.; Xie, Q.; Chi, L.; Gu, W.; Wu, D. Adsorption of phosphate from water by easily separable Fe3O4@SiO2 core/shell magnetic nanoparticles functionalized with hydrous lanthanum oxide. J. Colloid Interface Sci. 2016, 465, 76–82. [Google Scholar] [CrossRef]
  49. Ahmadi, S.; Khormali, A.; Khoutoriansky, F.M. Optimization of the demulsification of water-in-heavy crude oil emulsions using response surface methodology. Fuel 2022, 323, 124270. [Google Scholar] [CrossRef]
  50. Eljamal, O.; Khalil, A.M.E.; Sugihara, Y.; Matsunaga, N. Phosphorus removal from aqueous solution by nanoscale zero valent iron in the presence of copper chloride. Chem. Eng. J. 2016, 293, 225–231. [Google Scholar] [CrossRef]
  51. Awual, M.R.; Jyo, A. Assessing of phosphorus removal by polymeric anion exchangers. Desalination 2011, 281, 111–117. [Google Scholar] [CrossRef]
  52. Li, H.; Ye, Z.; Lin, Y.; Wang, F. Phosphorus recovery as struvite from eutropic waters by XDA-7 resin. Water Sci. Technol. 2012, 65, 2091–2097. [Google Scholar] [CrossRef] [PubMed]
  53. Nur, T.; Johir, M.A.H.; Loganathan, P.; Nguyen, T.; Vigneswaran, S.; Kandasamy, J. Phosphate removal from water using an iron oxide impregnated strong base anion exchange resin. J. Ind. Eng. Chem. 2014, 20, 1301–1307. [Google Scholar] [CrossRef] [Green Version]
  54. Simsek, E.B.; Tuna, O.; Balta, Z. Construction of stable perovskite-type LaFeO3 particles on polymeric resin with boosted photocatalytic Fenton-like decaffeination under solar irradiation. Sep. Purif. Technol. 2020, 237, 116384. [Google Scholar] [CrossRef]
  55. Trung, H.B.; Hong, S.P.; Yoon, J. Development of nanoscale zirconium molybdate, embedded anion exchange resin for selective removal of phosphate. Water Res. 2018, 134, 22–31. [Google Scholar]
  56. Yu, Y.; Chen, N.; Hu, W.; Feng, C. Application of Taguchi experimental design methodology in optimization for adsorption of phosphorus onto Al/Ca-impregnated granular clay material. Desalin. Water Treat. 2015, 56, 2994–3004. [Google Scholar] [CrossRef]
  57. O’Neal, J.A.; Boyer, T.H. Phosphorus recovery from urine and anaerobic digester filtrate: Comparison of adsorption-precipitation with direct precipitation. Environ. Sci. Water Res. 2015, 1, 481–492. [Google Scholar] [CrossRef]
  58. Ren, J.; Li, N.; Zhao, L.; Ren, N. Enhanced adsorption of phosphate by loading nanosized ferric oxyhydroxide on anion resin. Front. Environ. Sci. Eng. 2014, 8, 531–538. [Google Scholar] [CrossRef]
  59. Xiong, C.H.; Zhu, S.; Yao, C.P.; Chen, Q. Adsorption behavior of Pd(II) from aqueous solutions by D201 resin. Rare Metals 2011, 30, 470–476. [Google Scholar] [CrossRef]
  60. Lin, L.; Li, Z.; Song, X.; Jiao, Y.; Zhou, C. Preparation of chitosan/lanthanum hydroxide composite aerogel beads for higher phosphorus adsorption. Mater. Lett. 2018, 218, 201–204. [Google Scholar] [CrossRef]
  61. Zhang, Q.; Du, Q.; Jiao, T.; Pan, B.; Zhang, Z.; Sun, Q.; Wang, S.; Wang, T.; Gao, F. Selective removal of phosphate in waters using a novel of cation adsorbent: Zirconium phosphate (ZrP) behavior and mechanism. Chem. Eng. J. 2013, 221, 315–321. [Google Scholar] [CrossRef]
  62. Winter, J.; Wray, H.E.; Schulz, M.; Vortisch, R.; Barbeau, B.; Berube, P.R. The impact of loading approach and biological activity on NOM removal by ion exchange resins. Water Res. 2018, 134, 301–310. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Fabrication process of HTO-D201.
Figure 1. Fabrication process of HTO-D201.
Minerals 12 01596 g001
Figure 2. Schematic diagram of the column adsorption experiment.
Figure 2. Schematic diagram of the column adsorption experiment.
Minerals 12 01596 g002
Figure 3. (a,b) SEM image of HTO-D201; (c) the radial distribution of Ti; (d) the radial distribution of P; (e) TEM image of HTO-D201; (f) particle size distribution histograms of HTO-D201.
Figure 3. (a,b) SEM image of HTO-D201; (c) the radial distribution of Ti; (d) the radial distribution of P; (e) TEM image of HTO-D201; (f) particle size distribution histograms of HTO-D201.
Minerals 12 01596 g003aMinerals 12 01596 g003bMinerals 12 01596 g003c
Figure 4. XRD patterns of D201 and HTO-D201.
Figure 4. XRD patterns of D201 and HTO-D201.
Minerals 12 01596 g004
Figure 5. (a) Langmuir adsorption isotherm of HTO-D201; (b) Freundlich adsorption isotherm of HTO-D201.
Figure 5. (a) Langmuir adsorption isotherm of HTO-D201; (b) Freundlich adsorption isotherm of HTO-D201.
Minerals 12 01596 g005
Figure 6. Effect of pH on phosphate removal by HTO-D201.
Figure 6. Effect of pH on phosphate removal by HTO-D201.
Minerals 12 01596 g006
Figure 7. Factors of competitive ions on phosphate uptake by HTO-D201 and D201: (a) SO42−, (b) NO3, (c) Cl.
Figure 7. Factors of competitive ions on phosphate uptake by HTO-D201 and D201: (a) SO42−, (b) NO3, (c) Cl.
Minerals 12 01596 g007aMinerals 12 01596 g007b
Figure 8. Column adsorption curve.
Figure 8. Column adsorption curve.
Minerals 12 01596 g008
Figure 9. (a) Wide scan XPS spectra of phosphate loaded HTO-D201. (b) P2p XPS spectra of phosphate loaded D201 and HTO-D201.
Figure 9. (a) Wide scan XPS spectra of phosphate loaded HTO-D201. (b) P2p XPS spectra of phosphate loaded D201 and HTO-D201.
Minerals 12 01596 g009
Figure 10. Simulation of HTO-D201 structure and adsorption principle.
Figure 10. Simulation of HTO-D201 structure and adsorption principle.
Minerals 12 01596 g010
Table 1. Comparison of basic properties of D201 and HTO-D201.
Table 1. Comparison of basic properties of D201 and HTO-D201.
MaterialsD201HTO-D201
Cage constructionpolystyrenepolystyrene
Activity levelR-N+(CH3)3ClR-N+(CH3)3Cl and titanium oxides
BET (m2·g−1)23.6127.52
Average pore size (nm)24.8720.03
Mass fraction of Ti (%)012.80
Table 2. Fitting parameters of Freundlich and Langmuir model for phosphate adsorption by HTO-D201.
Table 2. Fitting parameters of Freundlich and Langmuir model for phosphate adsorption by HTO-D201.
AdsorbentT/KFreundlichLangmuir
K/mg·g−1·(g·L−1)nnR2Qmax/mg·g−1b/L·mg−1R2
28814.543.740.9631.571.210.97
HTO-D20129816.663.920.9634.081.490.96
30817.213.670.9836.891.230.96
D2012986.702.290.9925.020.150.97
Table 3. Thermodynamic parameters of phosphate adsorption on HTO-D201.
Table 3. Thermodynamic parameters of phosphate adsorption on HTO-D201.
AdsorbentT/KK0ΔG/kJ·mol−1ΔH/kJ·mol−1ΔS/J·(K·mol)−1
2881.48−0.953.8116.69
HTO-D2012981.64−1.27
3081.66−1.27
D2012980.880.33
Table 4. Comparison of phosphorus adsorption between HTO-D201 and other composites adsorbents.
Table 4. Comparison of phosphorus adsorption between HTO-D201 and other composites adsorbents.
Adsorbentqmax/mg·g−1T/KpH
ACF-ZrFe27.032987.00
Zirconium modified kaolin5.703037.00
Zirconia loaded ceramic 10.792987.00
FM-CD13.302987.00
Iron loaded ceramics12.50-6.60
Lanthanum—activated carbon fiber29.40-4.00
HMO-PN30.902887.00
Zirconium modified diatomite10.562986.30
Iron oxide coated sand1.502985.00
Carboxymethyl cellulose/Fe(II) treated
aspen wood fiber
4.302984.80
Al-loaded skin split waste21.65298-
HTO-D20134.082986.80
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhu, M.; Teng, Y.; Wu, D.; Zhu, J.; Zhang, Y.; Liu, Z. Development of Nanoscale Hydrated Titanium Oxides Support Anion Exchange Resin for Efficient Phosphate Removal from Water. Minerals 2022, 12, 1596. https://0-doi-org.brum.beds.ac.uk/10.3390/min12121596

AMA Style

Zhu M, Teng Y, Wu D, Zhu J, Zhang Y, Liu Z. Development of Nanoscale Hydrated Titanium Oxides Support Anion Exchange Resin for Efficient Phosphate Removal from Water. Minerals. 2022; 12(12):1596. https://0-doi-org.brum.beds.ac.uk/10.3390/min12121596

Chicago/Turabian Style

Zhu, Mingxin, Yue Teng, Dong Wu, Jiawei Zhu, Yi Zhang, and Zhiying Liu. 2022. "Development of Nanoscale Hydrated Titanium Oxides Support Anion Exchange Resin for Efficient Phosphate Removal from Water" Minerals 12, no. 12: 1596. https://0-doi-org.brum.beds.ac.uk/10.3390/min12121596

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