Zinc oxide (ZnO) was synthesized and used to investigate the mechanism of phosphate removal from aqueous solution. ZnO particles were characterized by X-ray diffraction, scanning electron microscope and Fourier transform infrared spectroscopy before and after adsorption. Batch experiments were carried out to investigate the kinetics, isotherms, effects of initial pH and co-existing anions. The adsorption process was rapid and equilibrium was almost reached within 150 min. The adsorption kinetics were described well by a pseudo-second-order equation, and the maximum phosphate adsorption capacity was 163.4 mg/g at 298 K and pH ∼6.2 ± 0.1. Thermodynamic analysis indicated the phosphate adsorption onto ZnO was endothermic and spontaneous. The point of zero charge of ZnO was around 8.4 according to the pH-drift method. Phosphate adsorption capacity reduced with the increasing initial solution pH values. The ligand exchange and Lewis acid-base interaction dominated the adsorption process in the lower and the higher pH range, respectively. Nitrate, sulfate and chloride ions had a negligible effect on phosphate removal, while carbonate displayed significant inhibition behavior.

As a principal nutrient, phosphate (P) plays a vital role in ecosystems. However, if excessive amounts of phosphate, released from agricultural fertilizer runoff, municipal wastewater, or effluent from chemical and food processing industries, are discharged into water bodies, there might be overgrowth of aquatic plants, depletion of dissolved oxygen and subsequent deterioration of water quality. Accordingly, it is necessary to reduce the phosphate concentration and control the discharge of phosphate.

To achieve this goal, various techniques, including chemical precipitation, biological removal, adsorption, crystallization and anion exchange, have been studied (Morse et al. 1998). However, the biological process is not steady and relies heavily on water quality, and chemical precipitation can be subject to problems with sludge treatment and disposal (Yeon et al. 2008). Anion exchange also has drawbacks, e.g., low selectivity in the presence of competing anions (Tsuji 2002). Compared with the traditional methods mentioned above, adsorption provides faster phosphate removal (Chouyyok et al. 2010), and can be applied to different phosphate concentrations under various environmental conditions (Wang et al. 2007; Haghseresht et al. 2009; Rentz et al. 2009; Pitakteeratham et al. 2013). It is considered a promising technique for removing phosphate. For this reason, a great many low-cost and easily available materials, including industrial by-products, natural materials, metal-modified minerals, man-made products, inorganic–organic hybrid materials and metal oxide/multi-metal composite oxides have been tested (Liao et al. 2006; Mohan & Pittman 2007; Vohla et al. 2011; Jiang & Ashekuzzaman 2012).

Zinc oxide (ZnO), as an environment-friendly material (Wan 2006), has been used for diverse applications (Hariharan 2006; Wang et al. 2011; Moezzi et al. 2012) and is considered a mature engineering material with annual production of one and a half million tons (Moezzi et al. 2012). Kikuchi et al. (2006) have reported its sorption performance toward metal ions and Liufu et al. (2004) investigated the effect of polyethylene glycol on the adsorption onto ZnO/polymer solution interface, and found that the blockage of the adsorption sites played a major role in the change of zeta potential. Regarding phosphorus, Namasivayam & Sangeetha (2004) found that ZnCl2-activated carbon developed from coir pith is a good sorbent for phosphate adsorption over a wide pH range (pH ∼ 3–10). Moreover, Cheng et al. (2009) and Zhou et al. (2011) reported that zinc-aluminum hydroxides (oxides) can effectively remove phosphate with a high content of ZNO at a Zn/Al molar ratio of 2, sintered at 573 K. He et al. (2010) reported that ZnAl layered double hydroxides show selective adsorption of phosphate and the calcined samples possibly proceed through ligand complexation or electrostatic attraction between phosphate ions and hydrated ZnO formed after calcination. However, up to now, phosphate adsorption by more stable ZnO particles has not been reported.

In the present study, ZnO was synthesized and used to investigate the mechanism of phosphate removal. The effect of pH and co-existing anions on phosphate capture, as well as the adsorption isotherm and kinetics, were evaluated. X-ray diffraction (XRD), scanning electron microscope (SEM) and Fourier transform infrared spectroscopy (FT-IR) were also performed to preliminarily elucidate the underlying sorption mechanism of ZnO. With this information, a better understanding of adsorption behaviors and properties could be obtained, and the feasibility of the use of ZnO in water/waste water will become clearer.

Materials

All chemicals of analytical grade were purchased from the Tianjin Fuchen Chemical Reagents Factory (Tianjin, China), and used without any further purifications. ZnO particles were fabricated by the chemical precipitation-thermal decomposition method with minor modifications. In brief, 0.1 M NaOH solution was dropwise added into 80 mL distilled water containing 10 g of ZnSO4•7H2O with continuous stirring, until the solution pH reached around 8 ± 0.1. The suspended solution was stirred for 3 h at room temperature. The obtained pastes were harvested by repeating the following procedure three times: centrifuging the solution at 4,000 rpm for 5 min, and then washing with distilled water. The pastes were then put into the Muffle furnace and kept at 573 K for 6 h (Zhou et al. 2011) to produce the ZnO particles. The obtained ZnO particles were ground into powder with an agate mortar and kept in a dryer for further use.

Analytical methods

All the samples were filtered through 0.45 μm membranes before determination. Phosphate concentrations were determined by a Metrohm 881 ion chromatograph coupled with a Metrosep A Supp 4 column. A solution comprising 1.8 mM of Na2CO3 and 1.7 mM of NaHCO3 was used as a mobile phase at a flow rate of 1.2 mL/min.

Phosphate adsorption capacities (qe, mg P/g) were determined by the following equation:
1
2
where V (mL) was the volume of the working solution, Co and Ce (mg/L) were the concentration in the working solution and filtrate, respectively, and m (g) was the adsorbent mass.

The surface morphology and element composition of the ZnO particles were obtained by a SUTW-Sapphire system with energy dispersive X-ray analysis (EDAX) (PHILIPS XL30 ESEM, Eindhoven, The Netherlands). The XRD pattern was used to characterize the relative intensity of ZnO particles, which were fabricated as described in the ‘Materials' section above, and recorded by a Bruker D8 Advance (Bruker, Bremen, Germany) using Cu Kα radiation within the 2θ range of 4–90°. The FT-IR spectroscopy was obtained using a Vertex 70 with a DTGS (deuterated triglycine sulfate) detector (Bruker, Bremen, Germany) at 4,000–400 cm−1, with a pellet of powdered potassium bromide. For comparison, the XRD, SEM and FT-IR of the ZnO particles after adsorption were also conducted.

Experimental procedures

A series of batch experiments was conducted to investigate the characteristics of phosphate adsorption onto ZnO. Kinetic experiments were performed by sampling a 2 mL solution from the 1,000 mL working solution containing 400 ± 0.03 mg of ZnO particles and the desired concentration of working solution (10, 20 and 30 mg/L) at various shaking times (0, 5, 10, 20, 30, 45, 60, 90, 150, 270, 390 510, 690 min) in an incubator shaker (WR-1, Shanghai, China) at 298 K.

Sorption isotherm experiments were conducted in 50 mL sealed Erlenmeyer flasks containing 25 mL working solution and 10 ± 0.03 mg of ZnO particles with various phosphate concentrations (0–100 mg/L) and a shaking time of 690 min at desired temperatures.

To determine the effect of pH, 10 ± 0.03 mg of ZnO particles was added into 50 mL Erlenmeyer flasks containing 25 mL working solution (20 mg/L) at desired pH values (3–11) adjusted by 0.1 M HCl and 0.1 M NaOH solution at 298 K. These flasks were sealed with parafilms (Pechiney Plastic Packaging Co., Washington, NJ, USA), and then transferred into the incubator shaker for 1,440 min. The solutions were then filtered through 0.45 μm membranes to determine the equilibrium pH. The concentrations of the releasing Zn2+ were determined by inductively coupled plasma (ICP) (PLASMA-SPEC I, Shelton, CT, USA).

The effects of co-existing anions (sulfate, carbonate, nitrate and chloride ion) on phosphate adsorption were investigated in 25 mL working solution (20 mg/L and 10 ± 0.03 mg of ZnO particles), which contained 10 or 20 mg/L of , , , or Cl ions at 298 K.

All working solutions were directly diluted from phosphate stock solution (counted as P, 1,000 mg/L) and each of the experiments was conducted twice.

SEM and XRD

According to the SEM photographs shown in Figure 1, ZnO particles exhibited individual, sheet morphologies and the particles formed aggregates of different sizes. The EDAX spectrum of ZnO particles after phosphate adsorption was also analyzed and the P element occurred in the EDAX spectrum (Figure S1, in the supporting information, available online at http://www.iwaponline.com/jwh/013/210.pdf), which confirmed the presence of phosphate on the surface of the ZnO particles after adsorption.

Figure 1

SEM photograph of ZnO particles after phosphate adsorption.

Figure 1

SEM photograph of ZnO particles after phosphate adsorption.

Close modal

Figure 2 shows the XRD results of fresh ZnO samples and ZnO samples obtained from the kinetic experiments. The strong and narrow peaks indicate that the adsorbent was well crystallized during the calcination process. The diffraction peaks of ZnO samples before and after adsorption correspond well with the standard data for ZnO (Zincite, syn. JCPDS card #36-1451).

Figure 2

XRD patterns of ZnO (a) ZnO after adsorption (b) ZnO before phosphate adsorption.

Figure 2

XRD patterns of ZnO (a) ZnO after adsorption (b) ZnO before phosphate adsorption.

Close modal

FT-IR analysis

The FT-IR spectra of fresh and phosphate-loaded ZnO particles are presented in Figure S2 (available online at http://www.iwaponline.com/jwh/013/210.pdf). The broad peaks apparent around the 3,400–1,630 cm−1 region might be attributed to the existence of the O-H stretching and bending vibrations, which were mainly caused by the adsorbed H2O molecules (Lv et al. 2013; Zhang et al. 2013). The peaks appearing at around 1,114, 1,069 and 965 cm−1 might be assigned to the vibration of (Wang et al. 2006) from the raw materials (ZnSO4·7H2O). The peaks presenting at around 415–425 cm−1 might correspond to the stretching frequency of Zn-O (Musić et al. 2002; Prasad et al. 2006). After adsorption, marked peaks appearing at 1,026, 1,006 and 945 cm−1 might be attributable to the symmetrical stretching vibration of phosphate (Arai & Sparks 2001; Ning et al. 2013). The peak at 965 cm−1 disappearing after phosphate adsorption might be covered by that of phosphate. It is also noteworthy that the frequency of Zn-O vibration moved from 415 to 425 cm−1 after uptake of phosphate, which might be due to the fact that the active sites reacted with phosphate.

Kinetics

The experimental data obtained as a function of reaction time were analyzed using the pseudo-first-order (Equation (3)) and pseudo-second-order (Equation (4)) kinetic models:
3
4
where qe and qt (mg P/g) were the amounts of phosphate adsorbed on ZnO at equilibrium and time t (min), respectively, k1 (min−1) was the rate constant of the pseudo-first-order kinetic model and k2 (g/(mg·min)) was the rate constant of the pseudo-second-order kinetic model.

As can be seen from Figure 3(a), the phosphate adsorption capacity of ZnO increased rapidly in the first 150 min, and then gradually increased until the equilibrium was reached at 510 min. The further extension of contact time did not affect the adsorption amounts of ZnO noticeably. The phosphate adsorption amount increased with the increasing initial phosphate concentration (from 23.6 mg P/g in 10 mg/L to 60.5 mg P/g in 30 mg/L).

Figure 3

Kinetic behaviors of phosphate adsorption onto ZnO. (a) Effect of contact time and initial phosphate concentration. (b) Pseudo-first-order kinetic model for phosphate adsorption. (c) Pseudo-second-order kinetic model for phosphate adsorption.

Figure 3

Kinetic behaviors of phosphate adsorption onto ZnO. (a) Effect of contact time and initial phosphate concentration. (b) Pseudo-first-order kinetic model for phosphate adsorption. (c) Pseudo-second-order kinetic model for phosphate adsorption.

Close modal

To further study the phosphate adsorption onto ZnO, the adsorption kinetic data were fitted to an integrated first (or second)-order kinetic expression, and are presented in Figure 3(b) and 3(c); the fitting parameters are listed in Table 1. According to the first-order kinetic model, the adsorption process could be divided into two parts by the adsorption rate constant (kp1 > kp2, kp1 and kp2 were the rate constant of fast and slow adsorption phase, respectively, and the values are listed in Table 1), i.e., fast adsorption and slow adsorption, which were also verified by the effect of contact time on the adsorption process shown in Figure 3(a). The fast adsorption process might be attributed to electrostatic attraction, or the driving force of a concentration gradient at higher initial phosphate concentration, which resulted in the rapid transportation of phosphate to the surface of the adsorbent, with respect to that of the adsorbent surface. The following slow adsorption process suggests that the driving force of the concentration gradient became weaker or the intraparticle diffusion dominated the adsorption process at that stage.

Table 1

Comparison of the pseudo-first- and second-order adsorption rate constants and experimental values for different initial phosphate concentrations

Co (mg P/L)qe(exp) (mg P/g)Pseudo-second-order kinetics
Pseudo-first-order kinetics
k2 (g/(mg·min))qe(cal) (mg P/g)R2kp1 (min−1)qe1(cal) (mg P/g)R2kp2 (min−1)qe2(cal) (mg P/g)R2
10 23.6 7.3 × 10−4 26.0 0.994 2.0 × 10−2 27.6 0.997 6.4 × 10−3 1.7 0.978 
20 41.9 3.4 × 10−4 46.5 0.991 1.9 × 10−2 46.3 0.952 4.5 × 10−3 5.5 0.920 
30 60.5 4.0 × 10−4 64.9 0.998 2.1 × 10−2 62.5 0.969 3.1 × 10−3 8.8 0.977 
Co (mg P/L)qe(exp) (mg P/g)Pseudo-second-order kinetics
Pseudo-first-order kinetics
k2 (g/(mg·min))qe(cal) (mg P/g)R2kp1 (min−1)qe1(cal) (mg P/g)R2kp2 (min−1)qe2(cal) (mg P/g)R2
10 23.6 7.3 × 10−4 26.0 0.994 2.0 × 10−2 27.6 0.997 6.4 × 10−3 1.7 0.978 
20 41.9 3.4 × 10−4 46.5 0.991 1.9 × 10−2 46.3 0.952 4.5 × 10−3 5.5 0.920 
30 60.5 4.0 × 10−4 64.9 0.998 2.1 × 10−2 62.5 0.969 3.1 × 10−3 8.8 0.977 

In addition, the pseudo-second-order model was also introduced to fit the kinetic data, and the results suggest that the adsorption of phosphate onto ZnO was chemisorption, or chemical bonding between adsorbent-active sites (Wang et al. 2006). That conclusion could also be supported by the close values of qe(exp) and qe(cal) of the pseudo-second-order model.

Isotherms

The experimental data obtained as a function of initial phosphate concentrations were analyzed using Langmuir (Equations (5) and (6)) and Freundlich (Equation (7)) models:
5
6
7
where, qm (mg P/g) was the theoretical maximum phosphate adsorption capacity, Co and Ce (mg/L) were phosphate concentrations in working solution and equilibrium solution, respectively, b (L/mg) was the affinity constant, RL was a dimensionless affinity constant, known as a separation factor adsorption, which indicated the adsorption nature to be either unfavorable (RL > 1), linear (RL = 1), favorable (0 < RL < 1) or irreversible (RL = 0), and Kf (mg/g) and n were Freundlich constants.

The adsorption of phosphate onto ZnO at different temperatures was fitted by Langmuir and Freundlich isotherm models and is depicted in Figure 4. According to Table 2, at 298 K, the isotherm data could be better fitted by the Langmuir model rather than the Freundlich model, indicating that the production of zinc phosphate might be neglected because the Langmuir model could have the optimum performance to characterize the adsorption processes without precipitation (Del Bubba et al. 2003). Comparing the values of parameter RL (Figure S3, available online at http://www.iwaponline.com/jwh/013/210.pdf), it is suggested that the adsorption of phosphate was more favorable at higher initial concentrations than at lower ones (Foo & Hameed 2010).

Table 2

Langmuir and Freundlich isotherm parameters for phosphate adsorption onto ZnO

Temperature (K)Langmuir
Separation factor
Freundlich
b (L/mg)qm (mg P/g)R2RLNKf (mg P/g)R2
288 0.0863 131.78 0.963 0.104 < RL < 0.698 3.37 18.49 0.968 
298 0.0803 168.37 0.967 0.111 < RL < 0.714 2.15 23.84 0.981 
308 0.2591 145.33 0.861 0.037 < RL < 0.436 2.15 47.37 0.970 
Temperature (K)Langmuir
Separation factor
Freundlich
b (L/mg)qm (mg P/g)R2RLNKf (mg P/g)R2
288 0.0863 131.78 0.963 0.104 < RL < 0.698 3.37 18.49 0.968 
298 0.0803 168.37 0.967 0.111 < RL < 0.714 2.15 23.84 0.981 
308 0.2591 145.33 0.861 0.037 < RL < 0.436 2.15 47.37 0.970 
Figure 4

Langmuir and Freundlich isotherms of ZnO at 288, 298, and 308 K.

Figure 4

Langmuir and Freundlich isotherms of ZnO at 288, 298, and 308 K.

Close modal

Simultaneously, the n values obtained from the Freundlich model ranged from 1 to 10, which also implies favorable adsorption of phosphate onto ZnO. In addition, at 308 K, the correlation coefficient R2 (0.970) suggests that high temperature strengthened the adsorption process, which led to disordered, overlapped distributions of phosphate ions on the surface of the ZnO particles, or chemical reaction between ZnO and phosphate to form crystals of zinc phosphate.

The qm values might not provide an accurate estimation of the long-term sorption capacity, but could still be useful for comparing alternative materials. As shown in Table 3, the adsorption capacity was 168.4 mg P/g for ZnO, which is much higher than most of the other adsorbents reported before, and suggests that ZnO has a relatively high potential for use as a phosphate adsorbent.

Table 3

Maximum phosphate adsorption capacity of different metal oxide adsorbents

AdsorbentAdsorption capacity (mg P/g)T (K)pHReference
ZnO 168.4 298 6.2 Present study 
Fe-Al-Mn trimetal oxides 48.3 298 6.8 Lv et al. (2013)  
Zn-Al oxides (573 K) 40.8 – 6.8 Cheng et al. (2009)  
Zn-Al LDH (Curea = 0.4 M) 76.1 303 Zhou et al. (2011)  
Zn-Al LDO (573 K, Curea = 0.4 M) 232.0 303 Zhou et al. (2011)  
ZnCl2-activated coir pith carbon 5.1 308 – Namasivayam & Sangeetha (2004)  
Manganese ore tailings 26.3 298 Liu et al. (2012)  
Pseudo-boehmite (γ-Al2O313.6 293 4.0 Yang et al. (2007)  
Al-Fe oxide (343 K) 71.6 – 4.8 Harvey & Rhue (2008)  
Fe(III)–Cu(II) binary oxides 35.2 298 7.0 Li et al. (2014)  
AdsorbentAdsorption capacity (mg P/g)T (K)pHReference
ZnO 168.4 298 6.2 Present study 
Fe-Al-Mn trimetal oxides 48.3 298 6.8 Lv et al. (2013)  
Zn-Al oxides (573 K) 40.8 – 6.8 Cheng et al. (2009)  
Zn-Al LDH (Curea = 0.4 M) 76.1 303 Zhou et al. (2011)  
Zn-Al LDO (573 K, Curea = 0.4 M) 232.0 303 Zhou et al. (2011)  
ZnCl2-activated coir pith carbon 5.1 308 – Namasivayam & Sangeetha (2004)  
Manganese ore tailings 26.3 298 Liu et al. (2012)  
Pseudo-boehmite (γ-Al2O313.6 293 4.0 Yang et al. (2007)  
Al-Fe oxide (343 K) 71.6 – 4.8 Harvey & Rhue (2008)  
Fe(III)–Cu(II) binary oxides 35.2 298 7.0 Li et al. (2014)  

– Not explicitly mentioned in the reference.

Thermodynamics

The thermodynamic parameters (ΔG°, ΔH° and ΔS°) were calculated according to the following equations (Equations (8) and (9)) (Deng & Yu 2012) to evaluate the thermodynamic feasibility and the nature of the adsorption process.
8
9
where ΔG°, ΔH° and ΔS° are the free energy of sorption (kJ/mol), the standard enthalpy change (kJ/mol) and entropy change (J/(mol·K), respectively, T (K) is the absolute temperature in Kelvin, R (8.314 J/(mol·K)) is the universal gas constant, Kd is the thermodynamic equilibrium constant, and ΔH° and ΔS° were determined as the slope and intercept of the linear plot of lnKd vs. 1/T, respectively.

As shown in Figure 5 and Table 4, the values of ΔG°, which indicate the spontaneity degree of the sorption process, are negative (−3.15 to −1.87 kJ/mol) and decreased gradually with the increasing temperature, suggesting that the adsorption process was spontaneous and was enhanced with the increasing temperature. The capacity for ZnO particles to adsorb phosphate also increased with increasing temperature (see Figure 4), which also supports this conclusion. The positive value of ΔH° (16.65 kJ/mol) demonstrates the endothermic nature of the reaction. The positive ΔS° value obtained (64.46 J/(mol·K)) indicates the degree of disorder, and the increase of randomness due to the increase in the number of species at the solid/liquid interface when the phosphate moves from the hydrous phase to the surface of the adsorbent (Wang et al. 2006).

Table 4

Thermodynamic parameters for phosphate adsorption on ZnO

Temperature (K)KdΔG° (kJ/mol)ΔH° (kJ/mol)ΔS° (J/mol·K)
288 2.18 −1.87 16.65 64.46 
298 2.92 −2.51 
308 3.42 −3.15 
Temperature (K)KdΔG° (kJ/mol)ΔH° (kJ/mol)ΔS° (J/mol·K)
288 2.18 −1.87 16.65 64.46 
298 2.92 −2.51 
308 3.42 −3.15 
Figure 5

Van't Hoff plot for phosphate adsorption onto ZnO.

Figure 5

Van't Hoff plot for phosphate adsorption onto ZnO.

Close modal

Effect of solution pH and sorption mechanism

The effect of pH on phosphate removal by ZnO was investigated at pH values ranging from 3 to 11. It is clear from Figure 6 that the adsorption is strongly dependent on pH value, and the adsorption capacity decreased significantly with the increasing pH value. The maximum adsorption capacity was about 50 mg P/g at pH ∼ 3.1, which was over 15 times that at pH ∼ 10.6. However increased pH value was found in the final solution after 690 min adsorption at an initial pH from 3 to 8.4 but this decreased at initial pH from 8.4 to 11.

Figure 6

Effect of solution pH on phosphate adsorption onto ZnO.

Figure 6

Effect of solution pH on phosphate adsorption onto ZnO.

Close modal

Phosphate exists in aqueous solution as different species, i.e., H3PO4///. The surface group properties of ZnO are associated with the point of zero charge (pHpzc), and variations of speciation distributions are attributed to the changes of solution pH. pHpzc values were determined by the pH-drift method described by Yang et al. (2004) and Liu et al. (2011). As can be seen from Figure 6, the pHpzc for ZnO is around 8.4. When solution pH is lower than 8.4, the surface of ZnO could be protonated as ≡Zn-OH + H+↔ ≡Zn-, where ≡Zn-OH represents the uncharged surface group, and ≡Zn-OH2+ represents the positively charged surface group. The positively charged surface groups are easier to replace at a lower pH from the metal binding sites than hydroxyl groups, which could facilitate the ligand exchange process (Chubar et al. 2005; He et al. 2010). Moreover, lower pH is favorable for phosphate adsorption through enlarging protonation to boost the positively charged sites, and thus leads to higher phosphate adsorption capacities. Simultaneously, anionic phosphate could interact with the positively charged surface group by electrostatic forces. However, with increasing pH value, becomes the dominant species instead of , progressively, and the capacity for continuous protonation is weakened, which leads to the decline of positively charged sites, and contributes to the decline in the ability of ZnO to adsorb for phosphate.

When solution pH is higher than 8.4, the dominant species are and , where the deprotonation of ≡Zn-OH, i.e., ≡Zn-OH – H+ ↔ ≡Zn-O + H+ exists, where ≡Zn-O represents the negatively charged surface group. The strong electrostatic repulsion between the two negative species Zn-O and become more dominant, which prevents from adsorbing onto the surface of ZnO. At the same time, the ligand exchange becomes weaker (He et al. 2010) because of deprotonation of the adsorbent, and the Lewis acid-base interaction is strengthened due to the increased numbers of oxygen anions in the different ionic forms of phosphate (Zhang et al. 2012). With increasing pH, the Lewis acid-base interaction dominates the adsorption process. There existed strong electrostatic repulsion between the surface of the adsorbent and the more negative-charged . At the same time, there was competition between the negatively charged phosphate species and the hydroxide groups adsorbed onto the more negatively charged adsorbent (Huang et al. 2013), and phosphate adsorption by electrical force might not occur. All these reasons led to the capacity for phosphate adsorption to drop to the minimum point, i.e., 2.9 mg P/g.

As an environment-friendly engineered material, it is difficult to ignore the solubility of ZnO in acid or base solutions, which is shown in Figure S4 (available online at http://www.iwaponline.com/jwh/013/210.pdf). For a better understanding of the following analysis, the equations below are suggested (Miao et al. 2010; Bian et al. 2011):
10
11
12
13
Phosphate with a low P/Zn molar ratio would reduce the release of Zn2+ into aqueous solution rapidly and substantially (Lv et al. 2012), and in this work, the fact that the concentration of the released Zn2+ was low might be due to a high P/Zn molar ratio in the working solution. The existence of Zn2+ or Zn(OH)+ in the acidic solution could enhance the removal efficiency through the precipitation of zinc phosphate or a combination of Zn(OH)+ and / in the acid solution. Otherwise, the appearance of or could restrain the removal efficiency by electrostatic repulsion. In addition, the attachment of or onto the surface of ZnO also leads to a decrease in removal efficiency.

Effect of co-existing anions

Competitive sorption is an important factor that influences the removal efficiency. Anions such as , , and Cl commonly exist in water and wastewater, and might interfere with the adsorption of phosphate by competing for sorptive sites on the surface of adsorbents. As shown in Figure 7, , and Cl do not show any negative effects on phosphate adsorption at concentrations from 0 to 20 mg/L. Anions such as are often weakly bounded with surface sites of metal hydroxides to form outer-sphere surface complexes, but phosphates are relatively strongly bounded with the surface sites forming inner-sphere surface complexes (Zhang et al. 2008), so the phosphate adsorption process is barely affected by , and Cl. However, Rahnemaie et al. (2007) suggested that carbonate was able to act as a competitor for phosphate at a relatively high concentration. The removal efficiency decreased from 96.5% to 53.5% when the concentration of was increased from 0 to 20 mg/L, which was probably due to the higher affinity of to the surface of ZnO to form or . According to the solubility product constants of zinc carbonate (Ksp = 1.46 × 10−10) and zinc phosphate (Ksp = 9.1 × 10−33) (Lide 1997-1998), zinc phosphate was easier to precipitate and more settlable in water solution. However, the addition of carbonate decreased the phosphate removal efficiency, suggesting that phosphate adsorption onto ZnO might not be caused by chemical precipitation but rather by chemical adsorption through exchange with OH on the hydroxylation surface of ZnO. However, carbonate addition during the adsorption process that consumed some H+ and improved solution pH values also decreased the phosphate removal efficiency.

Figure 7

Effect of co-existing anions on phosphate adsorption onto ZnO.

Figure 7

Effect of co-existing anions on phosphate adsorption onto ZnO.

Close modal

In the present study, ZNO was synthesized using the chemical precipitation-thermal decomposition method to investigate its capacity for phosphate adsorption. The obtained ZnO particles were well crystallized and exhibited individual, sheet morphologies. ZnO has a strong phosphorus removal ability of 163.4 mg/g and the dominant phosphate removal mechanism was probably due to adsorption rather than precipitation. Adsorption of phosphate was highly pH-dependent, and the adsorption capacity decreased with increasing pH. At the same time, Lewis acid-base interaction took the place of ligand exchange to be the dominant mechanism. significantly affected phosphate adsorption due to the higher affinity of for the surface of ZnO. Adsorption performance analysis can provide valuable information, representing practical value for the technological applications of phosphorus removal from aqueous solutions.

The authors wish to thank the Natural Science Foundation of China (No. 51238001) and the China Scholarship Council (CSC No. 201404480081) for their financial support.

Chouyyok
W.
Wiacek
R. J.
Pattamakomsan
K.
Sangvanich
T.
Grudzien
R. M.
Fryxell
G. E.
Yantasee
W.
2010
Phosphate removal by anion binding on functionalized nanoporous sorbents
.
Environ. Sci. Technol.
44
,
3073
3078
.
Chubar
N. I.
Kanibolotskyy
V. A.
Strelko
V. V.
Gallios
G. G.
Samanidou
V. F.
Shaposhnikova
T. O.
Milgrandt
V. G.
Zhuravlev
I. Z.
2005
Adsorption of phosphate ions on novel inorganic ion exchangers
.
Colloid Surface A
255
,
55
63
.
Jiang
J. Q.
Ashekuzzaman
S.
2012
Development of novel inorganic adsorbent for water treatment
.
Curr. Opin. Chem. Eng.
1
,
191
199
.
Liao
X. P.
Ding
Y.
Wang
B.
Shi
B.
2006
Adsorption behavior of phosphate on metal-ions-loaded collagen fiber
.
Ind. Eng. Chem. Res.
45
,
3896
3901
.
Lide
D. R.
1997–1998
Handbook of Chemistry and Physics
, 78th edn.
CRC press
,
Florida
,
USA
.
Lv
J. T.
Zhang
S. Z.
Luo
L.
Han
W.
Zhang
J.
Yang
K.
Christie
P.
2012
Dissolution and microstructural transformation of ZnO nanoparticles under the influence of phosphate
.
Environ. Sci. Technol.
46
,
7215
7221
.
Miao
A. J.
Zhang
X. Y.
Luo
Z. P.
Chen
C. S.
Chin
W. C.
Santschi
P. H.
Quigg
A.
2010
Zinc oxide-engineered nanoparticles: dissolution and toxicity to marine phytoplankton
.
Environ. Toxicol. Chem.
29
,
2814
2822
.
Moezzi
A.
McDonagh
A. M.
Cortie
M. B.
2012
Zinc oxide particles: synthesis, properties and applications
.
Chem. Eng. J.
185
,
1
22
.
Morse
G.
Brett
S.
Guy
J.
Lester
J.
1998
Review: phosphorus removal and recovery technologies
.
Sci. Total. Environ.
212
,
69
81
.
Musić
S.
Popović
S.
Maljković
M.
Dragčević
Đ.
2002
Influence of synthesis procedure on the formation and properties of zinc oxide
.
Mater. Chem. Phy.
77
,
521
530
.
Ning
Z. L.
Li
W. J.
Sun
C. Y.
Che
P.
Chang
Z. D.
2013
Synthesis and optical properties of zinc phosphate microspheres
.
Trans. Nonferrous Met. Soc. China
23
,
718
724
.
Prasad
V.
Souza
C. D.
Yadav
D.
Shaikh
A.
Vigneshwaran
N.
2006
Spectroscopic characterization of zinc oxide nanorods synthesized by solid-state reaction
.
Spectrochim. Acta. A
65
,
173
178
.
Rahnemaie
R.
Hiemstra
T.
Riemsdijk
W. H.
2007
Carbonate adsorption on goethite in competition with phosphate
.
J. Colloid Interf. Sci.
315
,
415
425
.
Rentz
J. A.
Turner
I. P.
Ullman
J. L.
2009
Removal of phosphorus from solution using biogenic iron oxides
.
Water Res.
43
,
2029
2035
.
Wang
S. L.
Cheng
C. Y.
Tzou
Y. M.
Liaw
R. B.
Chang
T. W.
Chen
J. H.
2007
Phosphate removal from water using lithium intercalated gibbsite
.
J. Hazard Mater.
147
,
205
212
.
Wang
L.
Zheng
Y.
Li
X.
Dong
W.
Tang
W.
Chen
B.
Li
C.
Li
X.
Zhang
T.
Xu
W.
2011
Nanostructured porous ZnO film with enhanced photocatalytic activity
.
Thin Solid Films
519
(
16
),
5673
5678
.
Yeon
K. H.
Park
H.
Lee
S. H.
Park
Y. M.
Lee
S. H.
Iwamoto
M.
2008
Zirconium mesostructures immobilized in calcium alginate for phosphate removal
.
Korean J. Chem. Eng.
25
,
1040
1046
.
Zhang
L.
Zhou
Q.
Liu
J.
Chang
N.
Wan
L.
Chen
J.
2012
Phosphate adsorption on lanthanum hydroxide-doped activated carbon fiber
.
Chem. Eng. J.
185–186
,
160
167
.

Supplementary data