In this paper, soybean protein isolate (SPI) was used as template, hydroxyapatite was crystallized on protein chains of SPI by in-situ synthesis, then the obtained inorganic HA/biopolymer SPI composite (HA@SPI) was calcined at suitable temperature, which afforded a novel hydroxyapatite-based porous materials (HApM). The results indicated that the product showed a porous morphology structure and excellent absorption performance for Pb2+. HApM maximum removal of lead was attained (96.25%) at an initial pH value of 7.4, temperature of 25 °C and contact time of 30 min with an initial metal concentration of 60 mg/L. In order to identify composition, structure and functional groups involved in the uptake of Pb2+, Fourier transform infrared spectrometer (FTIR), thermogravimetric analysis (TG), X-ray diffraction (XRD) scanning electron microscopy (SEM), energy dispersive spectroscopy (EDS), and Brunauer–Emmett–Teller (BET) analysis were carried out. Therefore, the hydroxyapatite-based porous materials (HApM) is a promising candidate for the treatment of liquid wastes containing toxic Pb2+ metal ion, heavy metal ion antidotes and other related fields.

Hydroxyapatite (HA) is the main inorganic constituent of vertebrate hard tissues such as bones and teeth, which has been chosen to study the system of biomolecule/biocompatible-surface interactions (Rimola et al. 2008). As the major constituent of bone (Akram et al. 2014; Liu et al. 2016), HA was coupled with inorganics (Fu et al. 2015; Anitha et al. 2017) and organics (Lv et al. 2013; Lin et al. 2016) for bone tissue engineering application. Since excellent biocompatibility, biodegradability, non-toxicity and high stability, it was extensively used in biomedical fields including dental restoration (Lee et al. 2017), controlling transport of proteins (Zhu et al. 2017), genes (Escudero et al. 2013; Deshmukh et al. 2016) and other applications (Fujita et al. 2016; Zhang et al. 2017). Furthermore, HA materials are of great significance in many other fields such as removing harmful volatile organic compounds (VOCs) (Fang et al. 2016), fuel cells (Wei & Matthew 2012) and corrosion inhibitors (Snihirova et al. 2010).

Due to Ca2+ exchange performance, HA has adsorption and immobilization effect on metal ions, such as Cd2+, Pd2+. Because of this feature, HA has been successfully used in heavy metal-contaminated soil (Wei et al. 2016) and water treatment (Meski et al. 2011; Wijesinghe et al. 2018). To improve absorption efficiency, biopolymer-hydroxyapatite composite was employed to remove heavy metals from aqueous solutions. Vila et al. (2011) have developed 3D-macroporous biopolymer-coated hydroxyapatite (HA) foams as potential devices for the treatment of Pb2+, cadmium and copper contamination of consumable water (Vila et al. 2011). Narwade et al. (2018) used in-situ synthesis method to prepare hydroxyapatite in the presence of cellulose nanofibrils (CNF) or TEMPO-oxidized CNF (TCNF) by wet chemical precipitation process, which combined with a low-cost biosorbent to remove heavy metal ion. Moreover, scientists found that HA particles with different morphologies, such as fibrous, plate-like, agglomerate-like, rodlike, needle-like and spherical, highly affected the efficiency of remove heavy metal ions (Qi et al. 2016; Wang et al. 2016).

The absorbent's surface structure and morphology deeply influence remove rate. We found that polymer modified inorganic clay could efficiently remove heavy metal ions (He et al. 2012), and plant-protein-modified semiconductor photocatalyst increased activity in photodegradation of dyes (Wang et al. 2018). Recently, increasing emphases were turned to natural polymers in the development of environmentally-friendly products (Wang et al. 2017; Al-Othman et al. 2012; Naushad & ALOthman 2015). As the soy protein isolate (SPI) with reproducible resource was used as template, this work could provide an efficient new treatment without high cost. In addition, HA and SPI are multifunctional biomaterials with biocompatibility and environmental friendliness. Herein, we prepared a novel HA-based hybrid with special structure and morphology by using SPI as bio-template. The obtained hybrid material was used as adsorbent to remove heavy metal ions from aqueous solutions to investigate its adsorption property. The influence of time, initial concentration of heavy metal ions, and the amount of adsorbent was tested. These hybrids have been studied in terms of isotherm models and sorption kinetics analysis at neutral medium in order to consider their application as heavy metal removers from aqueous solutions.

Materials

SPI (Shanghai Yuanye Bio-Technology Co. Ltd), urea (Tianjin Zhiyuan Chemical Reagent Co., Ltd), (NH4)2HPO4 (Shanghai Zhongqin Chemical Co., Ltd), Ca(NO3)2 (Tianjin Kexin Chemical Industry Co., Ltd), and Pb(NO3)2 (Tianjin Daming Chemical Reagent Factory) were obtained commercially and used without further purification. Distilled water was used in all experiments.

Preparation of HApM hybrids

In 20 mL of urea solution (8 mol/L), 1.000 g SPI was dissolved by stirring at room temperature, and then stirred for 12 hrs. Subsequently, the SPI solution was treated for 4 h and denatured in a heating bath at 90 °C. After adding 5 mL of (NH4)2HPO4 aqueous solution (1.195 mol/L), the mixture was stirred for 30 min. Next, 5 mL of Ca(NO3)2 aqueous solution (1.952 mol/L) was added dropwise into the mixture. Subsequently, adjusting pH value to 7.4 by the ammonia, the mixture was reacted for 3 h with stirring at 90 °C. Then, the system was continued to stir for 12 h at 25 °C. The product was centrifugally separated, and washed with distilled water for removing urea. After freeze-drying, HA@SPI composite was obtained.

The HA@SPI composites obtained by 3 g were calcined at the temperatures of 200 °C, 400 °C, 500 °C, 600 °C, 700 °C, 800 °C for 60 min by temperature-programmed calcination in muffle furnace to afford the hybrids and was named as SPI/HA-200, SPI/HA-400, PI/HA-500, SPI/HA-600 (HApM), SPI/HA-700, SPI/HA-800, respectively (Figure S1, Supplementary Information). The result showed that HApM (calcined at 600 °C) showed the best adsorption ability.

Characterizations

Fourier transform infrared (FTIR) spectra of the adsorbents were recorded between 4,000 and 400 cm−1 through the KBr method with a FTS3000 spectrophotometer (FTS3000, DigiLAB Merlin, USA). Thermogravimetric analysis (TGA) was carried out at a condition of N2 protection and temperature range being 25 °C–800 °C, heating rate being 10 °C/min (PerkinElmer, model Pyris Diamond, USA). The crystal structures and photocurrent responses were characterized by X-ray diffraction (XRD) (D/max-2400, Rigaku Corporation, Japan). A morphology analysis of certain features was visualized by using scanning electron microscopy (SEM) (ULTRA PLUS, ZEISS, Germany).

Adsorption performance of HApM hybrid

The Pb2+ were desorbed from HA, SPI, HA@SPI and HApM by shaking them in a 50 mL Pb2+ solution at a temperature of 25 °C and a speed of 125 rpm, respectively. After the adsorption process, the adsorbent was separated from the samples by filtering. Filtrate was analyzed by UV-Vis (Agilent 8,453, Agilent Technologies, USA). Parameters affecting the adsorption process, such as adsorbent dosage, initial Pb2+ concentration, contact time were studied. The removal rate and adsorption capacity were computed as following equations.
formula
(1)
formula
(2)
where C0 and Ce are the initial and equilibrium concentration of metal ions (mg/L) in the solution, respectively. V is the volume of metal ions solution (L), m is the weight of the adsorbent (g).

As we know, SPI is a kind of edible, low-cost and plant-based protein whose chains show a loop domain containing several helices structure (Maruyama et al. 2004). Here, using SPI as biopolymer template, the microcrystal of hydroxyapatite was crystallized along with loop protein chains and HA@SPI composite was synthesized. Subsequently, the HA@SPI was calcined at 600 °C, in the process of calcination, hydroxyapatite was mineralized and a part of SPI was carbonized. Finally, HA based hybrids (HApM) were successfully prepared (Figure 1).

Figure 1

Preparation of HA@SPI and HApM.

Figure 1

Preparation of HA@SPI and HApM.

Close modal

IR analysis

The obtained hydroxyapatite-based porous materials (HApM) were characterized by FTIR spectra, and compared with HA@SPI, templates materials (SPI), pure HA,and SPI/HA-Tx (Figure S2, Supplementary Information). The results are shown in Figure 2. It indicates that HA@SPI displayed characteristic absorption peaks both of SPI (1,649 cm−1, 1,535 cm−1 and 1,244 cm−1 are corresponding to -CONH-) and HA (PO43− vibration bands at 1,033, 603, and 566 cm−1 and -OH band at 3,441 cm−1). Characteristic absorption peaks of HApM are similar to HA@SPI, but the peaks of HA in HApM are more obvious. The amide two new characteristic peaks appeared at 2,050 cm−1 and 2,450 cm−1 which could be caused by carbonization. These results indicated that SPI of HA@SPI was turned to activated carbon at heating condition, and hydroxyapatite-based porous material was obtained (HApM). At the same time, a sharp peak at 1,628 cm−1 corresponded to H–O–H bending band, being also representative of the strongly bonded –OH groups in the matrix. The O–H stretching bands merged together and shifted to lower frequency in the spectra of HApM. This was due to the possibility of H-bonding. It is also apparent from Figure 2 that after adsorption of Pb2+ (HApM-after), the intensity and position of 1,628 cm−1 peak was changed, and the intensity of the 3,441 cm−1 peak weakens slightly, which demonstrated the involvement of –OH groups in the adsorption of Pb2+.

Figure 2

FTIR spectra of HA, SPI, HA@SPI, HApM and HApM-after.

Figure 2

FTIR spectra of HA, SPI, HA@SPI, HApM and HApM-after.

Close modal

TG analysis

TG curves are shown in Figure 3. Compare HA@SPI with SPI and pure HA, main weight loss of HA@SPI is due to SPI degradation, but for SPI itself, there was 22% remained at 800 °C which was according the degradation of SPI, its carbonization product. And it also can be deduced that the proportion of HA and SPI is 1:1. Although the calculated temperature was as high as 800 °C, the weight of HA@SPI is only 1.5% which owed to further carbonization of HA@SPI. These results suggested that the main interaction in preparing HApM progress was due to SPI carbonization of HA@SPI (Figure S3, Supplementary Information).

Figure 3

TG curves of HA, SPI, HA@SPI and HApM.

Figure 3

TG curves of HA, SPI, HA@SPI and HApM.

Close modal

XRD analysis

The XRD patterns of biopolymer-template SPI, pure HA, composite HA@SPI and HApM were given in Figure 4. It can be observed from the figure that when 2θ is 32°, the powder of SPI/HA-200, SPI/HA-400, SPI/HA-500, SPI/HA-600 and SPI/HA-800 all have (211) crystal plane diffraction peak, and when it is near 25°, The diffraction peak of (002) crystal plane appears, which is the most important characteristic diffraction peak of HA, indicating that SPI/HA, SPI/HA-200, SPI/HA-400, SPI/HA-500, HApM and SPI/HA-800 all formed HA crystals. At the same time, it can be seen from the diagram that the diffraction peaks of (102), (210), (300) and (202) planes tend to be sharp with the increase of calcination temperature, and the diffraction peaks of each crystal plane of SPI/HA-800 are sharp. The position and intensity are consistent with the characteristic diffraction peak of HA, indicating that the crystallization degree of HA in SPI/HA-800 complex is the best and the purity of the product is very high (Figure S4, Supplementary Information).

Figure 4

XRD patterns of HA, SPI, HA@SPI and HApM.

Figure 4

XRD patterns of HA, SPI, HA@SPI and HApM.

Close modal

SEM-EDS images

By observing the morphology of SPI/HA-200, SPI/HA-400, SPI/HA-500, HApM, SPI/HA-700 and SPI/HA-800 under scanning electron microscope (as shown in Figure S5, Supplementary Information), it can be seen that when the calcination temperature is 200 °C, the morphology of the composite SPI/HA-200 is close to that of HA@SPI, but the pore size increases slightly. When the calcination temperature is 400 °C, 500 °C and 600 °C, the pores of SPI/HA-400, SPI/HA-500 and HApM continue to increase, and there are different degrees of agglomeration at the same time. With the increase of temperature, when the calcination temperature rises to 700 °C, the pores of the composite SPI/HA-700 collapse and have obvious agglomeration, and the microscopic morphology is closer to that of HA. When the calcination temperature rises to 800 °C, the SPI/HA-800 agglomeration of the composite is serious. From the microscopic morphology of SPI/HA-Tx series, it can be seen that the SPI in the composite is gradually degraded with the increase of calcination temperature, which is consistent with the results of TG analysis. When the calcination temperature is 600 °C, the pore structure of SPI/HA-600(HApM) is the best. At the same time, the SEM pictures of HApM before and after adsorption of Pb2+ were compared. It was found that after the adsorption of Pb2+ ions, the porous morphology of HApM was filled in different degrees because of the adsorption of Pb2+on the surface and pores. The energy dispersive spectroscopy (EDS) elemental analysis of HApM after adsorption of Pb2+ was carried out, and the results were verified (Figure 5).

Figure 5

SEM images of (a) HA, (b) SPI, (c) HA@SPI, (d) HApM, (e) HApM after adsorption of Pb2+ and EDX spectra image of (f) HApM after adsorption of Pb2+.

Figure 5

SEM images of (a) HA, (b) SPI, (c) HA@SPI, (d) HApM, (e) HApM after adsorption of Pb2+ and EDX spectra image of (f) HApM after adsorption of Pb2+.

Close modal

BET analysis

The Brunauer–Emmett–Teller (BET) specific surface area (SBET) and pore-size distribution of the products were investigated. As shown in Figure 6(a) and 6(b), the specific surface area of HApM is 123.954 m2/g, which is higher than that of HApM-after (118.317 m2/g). The results show that due to the adsorption of Pb2+ on the surface and pores of HApM porous materials, the specific surface product of Pb2+ adsorbed decreases. The results were in good agreement with EDS analysis.

Figure 6

BET image of (a) HApM, (b) HApM after adsorption of Pb2+.

Figure 6

BET image of (a) HApM, (b) HApM after adsorption of Pb2+.

Close modal

Adsorption

Heavy metal ions in water can cause long-term and serious harm to people and the environment. In particular, Pb2+ can lead to endocrine and reproductive system disorders, teratogenicity and even carcinogenesis. Here, we used Pb2+ as model heavy metal ion, the respective adsorption ability of HA, SPI, HA@SPI and HApM were evaluated and shown in Table 1. The results showed that removal efficiency of Pb2+ by HApM was higher than HA, SPI and HA@SPI which implied that HAPM was an effective adsorbent.

Table 1

Adsorption ability of HA, SPI, HA@SPI and HApM in the removal of Pb2+

Adsorbent[Pb2+]Removal (%)Adsorption capacity (mg/g)
HApM 96.25 28.88 
HA@SPI 77.20 19.28 
HA 82.54 20.63 
SPI 8.54 2.13 
Adsorbent[Pb2+]Removal (%)Adsorption capacity (mg/g)
HApM 96.25 28.88 
HA@SPI 77.20 19.28 
HA 82.54 20.63 
SPI 8.54 2.13 

Optimization of the adsorption conditions of HApM

To investigate influence of adsorption conditions, the amount of adsorbent (HApM) (0.2, 0.1 and 0.05 g), adsorption time (30, 60 and 90 min), initial concentration of Pb2+ (30, 50, 60 mg/L) were selected. Removal rate (%) and adsorption capacity (mg/g) were calculated by orthogonal adsorption experiments. As shown in Table 2, in the condition of adsorbent dosage was 0.1 g, initial concentration of Pb2+ was 60 mg/L and adsorption time was 30 min, the adsorbent achieved its maximum adsorption efficiency of 96.25% removal rate and 28.88 mg/g adsorption capacity. If the adsorption capacity was taken into account, the removal rate (90.02%) and the adsorption capacity (45.01 mg/g) were relatively high in the condition of 0.05 g adsorbent, 50 mg/L initial concentration of Pb2+ and 60 min adsorption time.

Table 2

Orthogonal experiments of HApM about adsorbing Pb2+

No.Adsorbent (g)[Pb2+]Initial (mg/L)Time (min)Removal (%)Adsorption capacity (mg/g)
0.05 30 30 75.15 22.54 
0.1 30 60 77.51 11.63 
0.2 30 90 78.80 5.91 
0.05 50 60 90.02 45.01 
0.1 50 90 67.54 16.89 
0.2 50 30 88.38 11.05 
0.05 60 90 81.63 48.98 
0.1 60 30 96.25 28.88 
0.2 60 60 81.63 12.25 
No.Adsorbent (g)[Pb2+]Initial (mg/L)Time (min)Removal (%)Adsorption capacity (mg/g)
0.05 30 30 75.15 22.54 
0.1 30 60 77.51 11.63 
0.2 30 90 78.80 5.91 
0.05 50 60 90.02 45.01 
0.1 50 90 67.54 16.89 
0.2 50 30 88.38 11.05 
0.05 60 90 81.63 48.98 
0.1 60 30 96.25 28.88 
0.2 60 60 81.63 12.25 

Adsorption isotherm and kinetics

Adsorption isotherm Langmuir and Freundlich isotherm models were applied to establish the relationship between the amounts of Pb2+ adsorbed onto adsorbent and its equilibrium concentration in aqueous solution. The Langmuir adsorption isotherm was applied to equilibrium adsorption assuming monolayer adsorption onto a surface of the adsorbent, and adsorption occurs at finite number of identical sites within the adsorbent. The Langmuir isotherm was described as Equation (3). The Freundlich model stipulates that the ratio of solute adsorbed to the solute concentration was a function of the solution. The empirical model was shown to be consistent with exponential distribution of active center, characteristic of heterogeneous surfaces (Equation (4)).
formula
(3)
formula
(4)
where Ce is adsorption equilibrium concentration of adsorbate in solution (mg/L); qe is adsorption capacity (mg/g) per unit mass of adsorbent at equilibrium; qm is the maximum amount of monolayer adsorption per unit mass of adsorbent (mg/g); Ka is Langmuir isothermal adsorption energy constant (L/mg), KF is Freundlich isothermal adsorption empirical constant, which related to adsorption capacity and adsorption intensity.

As shown in Table 3 and Figure 7, the parameters of Langmuir and Freundlich for the adsorption of Pb2+ by the HApM hybrid are obtained, which is 0.9616 and 0.9853 for Langmuir and Freundlich. It can be concluded that the Freundlich isotherm is more suitable to describe the adsorption processes of Pb2+ onto HApM than the Langmuir isotherm. In addition, the adsorption index 1/n of this experiment is 0.4282, which is between 0.1 and 0.5, indicating that HApM is easy to adsorb Pb2+.

Table 3

Parameters of Langmuir and Freundlich isotherms

Adsorption isothermsAdsorption isotherms
LangmuirParameterFreundlichParameter
qm (mg/g) 28.1690 KF [(mg/g) (L/g)/n] 5.3134 
Ka (dm3/mg) 0.1211 1/n 0.4282 
R2 0.9616 R2 0.9853 
Adsorption isothermsAdsorption isotherms
LangmuirParameterFreundlichParameter
qm (mg/g) 28.1690 KF [(mg/g) (L/g)/n] 5.3134 
Ka (dm3/mg) 0.1211 1/n 0.4282 
R2 0.9616 R2 0.9853 
Figure 7

Langmuir and Freundlich isotherms for removal of Pb2+ by HApM. (a) Langmuir adsorption isotherm, (b) Freundlich adsorption isotherm.

Figure 7

Langmuir and Freundlich isotherms for removal of Pb2+ by HApM. (a) Langmuir adsorption isotherm, (b) Freundlich adsorption isotherm.

Close modal

Adsorption kinetics

Adsorption kinetics was used to explain the adsorption mechanism and adsorption characteristics. Pseudo-first-order model is rendered as the rate of occupation of the adsorption sites to be proportional to the number of unoccupied sites; pseudo-second-order kinetic model is assumed the chemical reaction mechanisms, and that the adsorption rate is controlled by chemical adsorption through sharing or exchange of electrons between the adsorbate and adsorbent (Naushad et al. 2017). Here, the pseudo-first-order and pseudo-second-order kinetic models have been employed to fit the experimental data. The pseudo-first-order model and pseudo-second-order kinetic model can be explained by Equations (5) and (6).
formula
(5)
formula
(6)
where qe and qt (mg/g) are the amount of Pb2+ adsorbed per unit mass of adsorbent at the equilibrium adsorption and at time period t (min), and k1 (min−1), k2 (g/m/min) are the first-order kinetic and pseudo-second-order rate constant.

As shown in Table 4 and Figure 8, the R2 of the pseudo-first-order and pseudo-second-order kinetic parameters is 0.9269 and 0.9984, respectively. Moreover, the values of qe (25.84 mg/g) calculated from pseudo-second-order kinetics are in agree with the experimental values of qe(exp) (25.18 mg/g). It indicates that the pseudo-second-order model is appropriate for the adsorption phenomena, and chemical adsorption plays a major role in the adsorption process.

Table 4

Parameters of pseudo-first and second-order kinetic models

C0 (mg/L)qe, exp (mg/g)Pseudo-first-order kinetics model
Pseudo-second-order kinetics model
qe (mg/g)k1 (min−1)R2qe (mg/g)k2[g/(mg min)]R2
60 25.18 7.4887 0.02381 0.9269 25.8398 0.00946 0.9984 
C0 (mg/L)qe, exp (mg/g)Pseudo-first-order kinetics model
Pseudo-second-order kinetics model
qe (mg/g)k1 (min−1)R2qe (mg/g)k2[g/(mg min)]R2
60 25.18 7.4887 0.02381 0.9269 25.8398 0.00946 0.9984 
Figure 8

Kinetic models for Pb2+ ions (a) pseudo-first-order, (b) pseudo-second-order.

Figure 8

Kinetic models for Pb2+ ions (a) pseudo-first-order, (b) pseudo-second-order.

Close modal

Mechanism of HApM and comparison with reported adsorbents

Based on above results, we considered it belongs to chemisorption. After absorption of Pb2+, HApM-after was characterized by SEM, FTIR and BET, and results are shown in Figures 2, 5 and 6, respectively. Comparing with HApM, in HApM-after (Figure 2), the intensity and position of 1,628 cm−1 peak was changed, and the intensity of the 3,441 cm−1 peak weakens slightly that demonstrated the involvement of –OH groups in the adsorption of Pb2+. From SEM images, it was found that, the surface morphology of HApM changed after adsorption of Pb2+ ions (Figure 5(e)). BET analysis also confirmed that the surface area of HApM decreased due to the adsorption of lead ion on the surface of HApM (Figure 6). Therefore, we speculated that lead ions react with hydroxyl groups so that lead ions were firmly adsorbed on the surface of hydroxyapatite.

The adsorption HApM was also compared with reported results (Table 5) (Mittal et al. 2016; Naushad 2014; Gnanasekaran et al. 2018; Pugazhendhi et al. 2018). Ghasemi et al. (2014a; 2014b) reported that the Ash, nFe-A and FSAC adsorbents have excellent adsorption capacity of Pb2+. Separately, and the maximum adsorption capacities are 588.24 mg/g, 833.33 mg/g, 80.645 mg/g. Naushad et al. (2015) found that the removal of Pb2+ got to 95% using TIV as adsorbent, and the maximum adsorption capacity was 18.8 mg/g. Le et al. (2019) used porous hydroxyapatite granules as an effective adsorbent for the Removal of Aqueous Pb(II) Ions which the removal of Pb2+ got to 95.9%, and the maximum adsorption capacity was 7.99 mg/g. Here, using HApM as adsorbent, the removal of lead got to 96.25%, while the adsorption capacity was 28.88 mg/g.

Table 5

Comparison of maximum monolayer adsorption capacities for Pb2+ removal

AdsorbentsOptimum condition
Removal of Pb2+ (%)qmax (mg/g)
pHTemp (°C)Adsorption time
KTSMBNL 13 7.0 35 26 h 90.0 – 
SDS-AZS 6.0 25 40 min 90.5 21.01 
Ash 6.0 25 40 min 89.0 588.24 
nFe-A 6.0 25 40 min 94.0 833.33 
FSAC 4.0 25 60 min 95.8 80.645 
TIV 6.0 20–55 60 min 95.0 18.8 
MWCNTs/ThO2 5.5 45 50 min 94.6 – 
PVA/HAp 5.5 30 40 min 95.9 7.99 
HApM 7.4 25 30 min 96.3 28.88 
AdsorbentsOptimum condition
Removal of Pb2+ (%)qmax (mg/g)
pHTemp (°C)Adsorption time
KTSMBNL 13 7.0 35 26 h 90.0 – 
SDS-AZS 6.0 25 40 min 90.5 21.01 
Ash 6.0 25 40 min 89.0 588.24 
nFe-A 6.0 25 40 min 94.0 833.33 
FSAC 4.0 25 60 min 95.8 80.645 
TIV 6.0 20–55 60 min 95.0 18.8 
MWCNTs/ThO2 5.5 45 50 min 94.6 – 
PVA/HAp 5.5 30 40 min 95.9 7.99 
HApM 7.4 25 30 min 96.3 28.88 

In summary, a kind of HA based hybrid has been successfully fabricated and tested as Pb2+ removers from aqueous solutions. Morphology and structure of the HApM suggest that protein chains were shank and carbonized in the calcined process and formed a highly porous HA based hybrid which have proved an excellent adsorption capacity of 96.25% for Pb2+. The adsorption data fitted well with Freundlich isotherm model, and being followed by the chemisorption according to the pseudosecond-order. All of those suggest that the HA based hybrid will be an ideal Pb2+ adsorbent for the treatment of liquid wastes containing toxic Pb2+ metal ion, which could provide high effectiveness with low-cost and environmental friendliness.

The project was supported by the National Natural Science Foundation of China (Grant No. 21865030; 21364012).

The Supplementary Material for this paper is available online at http://dx.doi.org/10.2166/wst.2019.370.

Akram
M.
,
Ahmed
R.
,
Shakir
I.
,
Ibrahim
W. A. W.
,
Hussain
R.
2014
Extracting hydroxyapatite and its precursors from natural resources
.
Journal of Materials Science
49
,
1461
1475
.
Anitha
A.
,
Menon
D.
,
Sivanarayanan
T. B.
,
Koyakutty
M.
,
Mohan
C.
,
Shantikumar
V. N.
,
Manitha
B. N.
2017
Bioinspired composite matrix containing hydroxyapatite-silica core-shell nanorods for bone tissue engineering
.
ACS Applied Materials & Interfaces
9
,
26707
26718
.
Deshmukh
K.
,
Shaik
M. M.
,
Ramanan
S. R.
,
Kowshik
M.
2016
Self-activated fluorescent hydroxyapatite nanoparticles: a promising agent for bioimaging and biolabeling
.
ACS Biomaterials Science & Engineering
2
,
1257
1264
.
Fang
B. B.
,
Yan
Y. B.
,
Yang
Y.
,
Wang
F. L.
,
Chu
Z.
,
Sun
X. Y.
,
Li
J. S.
,
Wang
L. J.
2016
Adsorption of Pb2+ from aqueous solution using spinel ferrite prepared from steel pickling sludge
.
Water Science & Technology
73
(
5
),
1112
.
Fu
C.
,
Savino
K.
,
Gabrys
P.
,
Zeng
A. B.
,
Guan
B. H.
,
Olvera
D.
,
Wang
C. G.
,
Song
B. A.
,
Awad
H.
,
Gao
Y. L.
,
Yates
M. Z.
2015
Hydroxyapatite thin films with giant electrical polarization
.
Chemistry of Materials
27
,
1164
1171
.
Fujita
H.
,
Kudo
T. A.
,
Kanetaka
H.
,
Miyazaki
T.
,
Hashimoto
M.
,
Kawashita
M.
2016
Adsorption of laminin on hydroxyapatite and alumina and the MC3T3-E1 cell response
.
ACS Biomaterials Science & Engineering
2
,
1162
1168
.
Ghasemi
M.
,
Naushad
M.
,
Ghasemi
N.
,
Khosravi-Fard
Y.
2014a
Adsorption of Pb (II) from aqueous solution using new adsorbents prepared from agricultural waste: adsorption isotherm and kinetic studies
.
Journal of Industrial and Engineering Chemistry
20
(
4
),
2193
2199
.
Ghasemi
M.
,
Naushad
M.
,
Ghasemi
N.
,
Khosravi-Fard
Y.
2014b
A novel agricultural waste based adsorbent for the removal of Pb (II) from aqueous solution: kinetics, equilibrium and thermodynamic studies
.
Journal of Industrial and Engineering Chemistry
20
(
2
),
454
461
.
Gnanasekaran
L.
,
Hemamalini
R.
,
Naushad
M.
2018
Efficient photocatalytic degradation of toxic dyes using nanostructured TiO2/polyaniline nanocomposite
.
Desalination and Water Treatment
108
,
322
328
.
He
Y.-H.
,
Zhang
L.
,
Wang
R.-M.
,
Li
H.-R.
,
Wang
Y.
2012
Loess clay based copolymer for removing Pb(II) ions
.
Journal of Hazardous Materials
227
,
334
340
.
Le
D. T.
,
Le
T. P. T.
,
Do
H. T.
,
Vo
H. T.
,
Pham
N. T.
,
Nguyen
T. T.
,
Tran
D. L.
2019
Fabrication of porous hydroxyapatite granules as an effective adsorbent for the removal of aqueous Pb (II) ions
.
Journal of Chemistry
2019
, 1–10.
Lee
H. S.
,
Myers
C.
,
Zaidel
L.
,
Nalam
P. C.
,
Caporizzo
M. A.
,
Daep
C. A.
,
Eckmann
D. M.
,
Masters
J. G.
,
Composto
R. J.
2017
Competitive adsorption of polyelectrolytes onto and into pellicle-coated hydroxyapatite investigated by QCM–D and force spectroscopy
.
ACS Applied Materials & Interfaces
9
,
13079
13091
.
Lin
K. F.
,
He
S.
,
Song
Y.
,
Wang
C. M.
,
Gao
Y.
,
Li
J. Q.
,
Tang
P.
,
Wang
Z.
,
Bi
L.
,
Pei
G. X.
2016
Low-temperature additive manufacturing of biomimic three dimensional hydroxyapatite/collagen scaffolds for bone regeneration
.
ACS Applied Materials & Interfaces
8
,
6905
6916
.
Liu
C.
,
Zhai
H. L.
,
Zhang
Z. S.
,
Li
Y. l.
,
Xu
X. R.
,
Tang
R. K.
2016
Cells recognize and prefer bone-like hydroxyapatite: biochemical understanding of ultrathin mineral platelets in bone
.
ACS Applied Materials & Interfaces
8
,
29997
30004
.
Lv
L. X.
,
Zhang
X. F.
,
Wang
Y. Y.
,
Ortiz
L.
,
Mao
X.
,
Jiang
Z. L.
,
Xiao
Z. D.
,
Huang
N. P.
2013
Effects of hydroxyapatite-containing composite nanofibers on osteogenesis of mesenchymal stem cells in vitro and bone regeneration In vivo
.
ACS Applied Materials & Interfaces
5
,
319
330
.
Maruyama
Y.
,
Maruyama
N.
,
Mikami
B.
,
Utsumi
S.
2004
Structure of the core region of the soybean β-conglycinin αʹsubunit
.
Acta Crystallographica Section D-Structural Biology
60
(
2
),
289
297
.
Meski
S.
,
Ziani
S.
,
Khireddine
H.
,
Yataghane
F.
,
Ferguene
N.
2011
Elaboration of the hydroxyapatite with different precursors and application for the retention of the Pb2+
.
Water Science & Technology
63
(
10
),
2087
2096
.
Mittal
A.
,
Naushad
M.
,
Sharma
G.
,
ALothman
Z. A.
,
Wabaidur
S. M.
,
Alam
M.
2016
Fabrication of MWCNTs/ThO2 nanocomposite and its adsorption behavior for the removal of Pb (II) metal from aqueous medium
.
Desalination and Water Treatment
57
(
46
),
21863
21869
.
Narwade
V. N.
,
Khairnar
R. S.
,
Kokol
V.
2018
In situ synthesized hydroxyapatite – cellulose nanofibrils as biosorbents for heavy metal ions removal
.
Journal of Polymers and the Environment
26
(
5
),
2130
2141
.
Naushad
M.
,
Ahamad
T.
,
Al-Maswari
B. M.
,
Alqadami
A. A.
,
Alshehri
S. M.
2017
Nickel ferrite bearing nitrogen-doped mesoporous carbon as efficient adsorbent for the removal of highly toxic metal ion from aqueous medium
.
Chemical Engineering Journal
330
,
1351
1360
.
Pugazhendhi
A.
,
Boovaragamoorthy
G. M.
,
Ranganathan
K.
,
Naushad
M.
,
Kaliannan
T.
2018
New insight into effective biosorption of Pb2+ from aqueous solution using Ralstonia solanacearum: characterization and mechanism studies
.
Journal of Cleaner Production
174
,
1234
1239
.
Qi
Y. C.
,
Shen
J.
,
Jiang
Q. Y.
,
Jin
B.
,
Chen
J. W.
,
Zhang
X.
,
Su
J. L.
2016
Hierarchical porous hydroxyapatite microspheres: synthesis and application in water treatment
.
Journal of Materials Science
51
,
2598
2607
.
Rimola
A.
,
Corno
M.
,
Zicovich-Wilson
C. M.
,
Ugliengo
P.
2008
Ab initio modeling of protein/biomaterial interactions: glycine adsorption at hydroxyapatite surfaces
.
Journal of the American Chemical Society
130
,
16181
16183
.
Snihirova
D.
,
Lamaka
S. V.
,
Taryba
M.
,
Salak
A. N.
,
Kallip
S.
,
Zheludkevich
M. L.
,
Ferreira
M. G. S.
,
Montemor
M. F.
2010
Hydroxyapatite microparticles as feedback-active reservoirs of corrosion inhibitors
.
ACS Applied Materials & Interfaces
2
(
11
),
3011
3022
.
Vila
M.
,
Sánchezsalcedo
S.
,
Cicuéndez
M.
,
Izquierdo-Barba
I.
,
Vallet-Regí
M.
2011
Novel biopolymer-coated hydroxyapatite foams for removing heavy-metals from polluted water
.
Journal of Hazardous Materials
192
(
1
),
71
77
.
Wang
D. D.
,
Guan
X. M.
,
Huang
F. Z.
,
Li
S. K.
,
Shen
Y. H.
,
Chen
J.
,
Long
H. B.
2016
Removal of heavy metal ions by biogenic hydroxyapatite: morphology influence and mechanism study
.
Russian Journal of Physical Chemistry A
90
(
8
),
1557
1562
.
Wang
J. F.
,
Qian
W. Z.
,
He
Y. F.
,
Xiong
Y. B.
,
Song
P. F.
,
Wang
R.-M.
2017
Reutilization of discarded biomass for preparing functional polymer materials
.
Waste Management
65
,
11
21
.
Wang
Y.
,
She
H. D.
,
Ma
Q.
,
Lian
J. H.
,
Zhong
J. B.
,
Li
J. Z.
,
Tong
J. H.
,
He
Y. F.
,
Wang
R.-M.
,
Wang
Q. Z.
2018
Plant-protein-modified TiO2(SPI@TiO2) for photodegradation of dyes
.
Chemistry Select
3
(
11
),
3127
3132
.
Wei
X.
,
Matthew
Z. Y.
2012
Yttrium-doped hydroxyapatite membranes with high proton conductivity
.
Chemistry of Materials
24
,
1738
1743
.
Wei
L.
,
Wang
S. T.
,
Zuo
Q. Q.
,
Liang
S. X.
,
Shen
S. G.
,
Zhao
C. X.
2016
Nano-hydroxyapatite alleviates the detrimental effects of heavy metals on plant growth and soil microbes in e-waste-contaminated soil
.
Environmental Science-Processes & Impacts
18
(
6
),
760
767
.
Wijesinghe
W. P. S. L.
,
Mantilaka
M. M. M. G. P. G.
,
Nirmal Peiris
T. A.
,
Rajapakse
R. M. G.
,
Upul Wijayantha
K. G.
,
Pitawala
H. M. T. G. A.
,
Premachandra
T. N.
,
Herathg
H. M. T. U.
,
Rajapaksef
R. P. V. J.
2018
Preparation and characterization of mesoporous hydroxyapatite with non-cytotoxicity and heavy metal adsorption capacity
.
New Journal of Chemistry
42
(
12
),
10271
10278
.

Supplementary data