Next Article in Journal
High Efficient and Cost Effective Titanium Doped Tin Dioxide Based Photocatalysts Synthesized via Co-precipitation Approach
Next Article in Special Issue
Stabilization of Lecitase Ultra® by Immobilization and Fixation of Bimolecular Aggregates. Release of Omega-3 Fatty Acids by Enzymatic Hydrolysis of Krill Oil
Previous Article in Journal
A Review of the Desulfurization Processes Used for Waste Tire Pyrolysis Oil
Previous Article in Special Issue
Sugarcane Bagasse Saccharification by Enzymatic Hydrolysis Using Endocellulase and β-glucosidase Immobilized on Different Supports
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Green and High Effective Scale Inhibitor Based on Ring-Opening Graft Modification of Polyaspartic Acid

1
School of Mechanical Engineering, Hefei University of Technology, Tunxi Road, Hefei 230009, China
2
State Key Laboratory of Materials-Oriented Chemical Engineering, College of Biotechnology and Pharmaceutical Engineering, Nanjing Tech University, Nanjing 211816, China
*
Author to whom correspondence should be addressed.
Catalysts 2021, 11(7), 802; https://doi.org/10.3390/catal11070802
Submission received: 23 February 2021 / Revised: 28 June 2021 / Accepted: 28 June 2021 / Published: 30 June 2021
(This article belongs to the Special Issue Application of Immobilized Enzyme as Catalysts in Chemical Synthesis)

Abstract

:
Polyaspartic acid (PASP)-based green scale inhibitor has great potential application in water treatment. Here, we first synthesized PASP in ionic liquid. Then, an effective PASP-based green scale inhibitor was synthesized by ring-opening graft modification of PASP with both aspartic acid (ASP) and monoethanolamine (MEA). Its chemical composition was characterized by gel chromatography (GPC), Fourier transform infrared spectroscopy (FTIR), and 1H nuclear magnetic resonance (1H NMR). Scale inhibition efficiency was measured by static scale inhibition tests. The results showed that the new PASP-based scale inhibitor has high scale inhibition to both CaCO3 and Ca3(PO4)2. When the concentration was increased to 2 mg/L, the inhibition efficiency of the new PASP-based scale inhibitor was 99% for CaCO3, while when the concentration was raised to only 4 mg/L, its inhibition efficiency increased to 100% for Ca3(PO4)2. Scanning electronic microscopy (SEM) and X-ray diffraction (XRD) were used to analyze the changes of crystal structure for CaCO3 and Ca3(PO4)2 after adding the new PASP-based scale inhibitor. The crystal size of CaCO3 and Ca3(PO4)2 became smaller and the crystal form became amorphous after adding the modified PASPs compared with adding pure PASP. Moreover, the modified PASP showed good biodegradation performance.

Graphical Abstract

1. Introduction

In recent years, with the increase of worldwide population and expansion of modern society, the consumption and pollution of industrial water increases shortage of water resources significantly [1,2,3]. Water treatment is an important technology to deal with the shortage of water resources [4,5,6]. Among various methods explored for water treatment, the circulating cooling water system is widely utilized in industrial processes [7,8]. However, scale deposition is one of the major problems in this system [9]. Scales degrade the performance of heat exchangers by increasing the resistance to heat transfer and eventually result in tremendous economic loss due to energy waste by increasing the power requirements of pumps. In order to control scale formation, a variety of scale inhibitors have been explored and used widely in cooling water systems. Scale inhibitors are usually non-polymeric phosphonates (ATMP, HEDP, PBTC) and polymers with functional groups. The most common polymers are phosphonate, carboxylate, and sulfonate polymers. However, phosphonate and sulfonate polymers are being restricted due to environmental legislation [10,11,12]. With the increase of environmental awareness, environmentally friendly scale inhibitors have been drawing broad attention all over the world [13,14,15,16,17].
Polyaspartic acid (PASP) is a promising environmentally friendly scale inhibitor with nontoxicity and good biodegradability. It has excellent scale inhibiting performance for CaCO3. However, its scale inhibiting performance for Ca3(PO4)2 is poor. In order to improve the comprehensive scale inhibition performance of PASP, it is essential to modify it by introducing functional groups into its side chain, such as hydroxyl groups, carboxylic groups, sulfonic groups, and/or phosphonyl groups [18,19].
Many studies have already reported on PASP modification. For example, Zhang et al. synthesized PASP/Urea by introducing an acylamino group into the side chains of PASP. It was shown that the modified PASP has better scaling inhibition to CaSO4 [20]. Chen et al. synthesized Ser-PASP via introducing a hydroxylic group into the side chain of PASP. When Ser-PASP was added, the scale crystals became irregular, and complete inhibition to CaSO4 was achieved [21]. Xu et al. obtained PASP–melamine grafted copolymer via polysuccinimide (PSI) ring-opening by melamine [22]. Xu et al. synthesized PASP/aminobenzenesulfonic acid (ABSA) copolymer using sulphanilic acid and PSI. It was found that the PASP/ABSA copolymer is able to efficiently inhibit CaCO3 [23]. Although single-functional-group-modified PASP has shown improved scale inhibition to CaCO3 and CaSO4, it is still a challenge to achieve high scale inhibition to Ca3(PO4)2.
Here, we prepared a new green and highly effective PASP-based scale inhibitor for water treatment (Figure 1). PASP was first synthesized in ionic liquid. Ionic liquids are a subset of molten salts with melting points at or below 100 °C [24]. Spurred by the green chemistry movement, ionic liquids are considered as promising alternative solvents to replace traditional volatile organic compounds (VOCs) due to their low volatility [25]. Then, a ring-opening graft modification strategy was adopted to modify the synthesized PASP with both aspartic acid (ASP) and monoethanolamine (MEA). The ASP and MEA provide a carboxyl and hydroxyl group to PASP, respectively. The introduced carboxyl can enhance the calcium chelation ability and solubilization for the crystals of CaCO3 and Ca3(PO4)2, while the hydroxyl group may result in distorting crystal lattices. In general, the synergy effect of the double-functional-group-modified PASP showed high scale inhibition performance for CaCO3 and Ca3(PO4)2, and it also showed good biodegradation performance.

2. Results and Discussion

2.1. Ring-Opening Graft Modification of PASP

We first synthesized PASP in ionic liquid. Ionic liquid has unique advantages, including good thermal stability, strong solubility, nonvolatility, free design of anion and anion, etc. It can replace traditional organic solvents and catalysts to achieve green synthesis of PASP [26]. Then, a ring-opening graft modification strategy was adopted to modify PASP with both carboxyl and hydroxyl groups. Here, we put emphasis on discussing the graft modification process on the grafting ratio of ASP and MEA. The grafting ratio of both molecules are largely determined by the molar ratio of reagent (PSI–ASP–MEA), reaction time, and reaction temperature (Figure 2). First, the grafting ratio showed a significant increase with the increase of molar ratio of ASP and MEA (Figure 2a). When the molar ratio of the reagent (PSI–ASP–MEA) was fixed at 1:1:0, the grafting ratio of ASP was 74.0%, while when the molar ratio of the reagent (PSI–ASP–MEA) was fixed at 1:0:1, the grafting ratio of MEA was 89.8%. The difference of the grafting ratio for the two molecules may be ascribed to the difference of the steric hindrance effect between them. When successively increasing the molar ratio of ASP or MEA, the grafting ratio of ASP and MEA obviously increased. This can be ascribed to the increased reaction probability during the whole reaction process. Typically, when the molar ratio of the reagent (PSI–ASP–MEA) was fixed at 1:0.5:0.5, the grafting ratio of ASP and MEA was probably 36.22% and 43.3%, respectively. Then, when increasing the reaction time from 12 h to 24 h, the increase first brought a significant enhancement of the grafting ratio for ASP from 32.8% to 36.2% and for MEA from 39.8% to 43.4%. This may be because there were more reactive sites generated during the initial reaction from 12 h to 24 h (Figure 2b). However, little change to the grafting ratios of either ASP or MEA occurred when further increasing the reaction time from 24 h to 60 h (Figure 2b). This indicates that the reaction reached equilibrium at 24 h. As shown in Figure 2c, the grafting ratio of both ASP and MEA decreased significantly when increasing the reaction temperature from 0 °C to 40 °C. Specifically, the grafting ratio of ASP decreased from 38.8% to 31.9%, and the grafting ratio of MEA decreased from 43.2% to 24.2%. As is well known, higher temperature usually leads to an increased number of active molecules, increased effective collision, and higher reaction rate. It should theoretically lead to an increased grafting ratio of both ASP and MEA during the fixed time. However, because the ring-opening graft modification is an exothermic reaction, the increased temperature inhibited the reaction process [27] and led to hydrolysis of the formed graft bond. Thus, the grafting ratio of both ASP and MEA inevitably decreased.

2.2. Chemical Characterization of the Modified PASP

From Table 1, we can obtain the conclusion that the molecular mass of PASP synthesized in the ionic liquid was Mn 4.87 kDa (PDI~1.63). Compared with the pure PASP, GPC results show that the molecular mass of the modified PASP increased with the increase of MEA or ASP proportion in the monomer solution. Particularly, when the molar ratio of PSI–ASP–MEA was 1:0.5:0.5 (at reaction temperature of 0 °C and reaction time of 24 h), the molecular mass of the modified PASP was increased to Mn 4.90 kDa. The corresponding grafting ratios of ASP and MEA were calculated to be increased to 36.2% and 43.3%, respectively. This is reasonable, because increasing the molar ratio of ASP or MEA increases the grating ratio of the two molecules, inevitably enhancing the ring-opening graft modification. Thus, graft modification obviously increased the molecular weight of PASP.
We then verified that the ASP and MEA were conjugated onto the PASP side chain by analyzing the molecular structure using FTIR (Figure 3). The absorption peaks at ~3422 cm−1, ~1598 cm−1 and ~1401 cm−1 are attributed to the stretching vibration of N-H and C=O in –CONH and the absorption peak at C-N [28,29]. This indicates that the PASP was successfully synthesized in the ionic liquid. In addition, for the modified PASP, a peak corresponding to the stretching vibration of the C=O group was observed at ~1703 cm−1, which indicates that the ASP was successfully grafted onto the side chain of PASP. Moreover, the stretching vibrations peak of the -C-O- group at ~1164 cm−1 and ~1064 cm−1 indicate that the MEA was successfully introduced to the side chain of PASP. In addition, 1H NMR was also used to verify the graft modification process (Figure S1). Proton NMR results show that for the PASP main chain, the peaks corresponding to -CH- and -CH2- were observed at 4.45 ppm and 2.74 ppm, respectively. The peaks at 3.60~3.66 ppm and 3.16 ppm are attributed to -CH- and -CH2- of the side chain for carboxyl group-modified PASP (PASP-ASP). Similarly, the peaks at 3.8 ppm and 3.12 ppm are attributed to -CH- and -CH2- of the side chain for hydroxyl group-modified PASP (PASP-MEA). Furthermore, we obtained all of the above characteristic peaks for the product after both the carboxyl and hydroxyl group modifications to PASP, which indicates that the PASP-ASP-MEA was successfully synthesized.

2.3. Scale Inhibition Performance of the Modified PASP

To investigate the scale inhibition properties of the modified PASP, we chose CaCO3 and Ca3(PO4)2 as the two typical crystals. Figure 4a,b presents the scale inhibition performance of PASP-ASP-MEA against CaCO3 and Ca3(PO4)2 scales. When its concentration was 1 mg/L and 2 mg/mL, the inhibition efficiency of PASP-ASP-MEA against CaCO3 scale reached 95% and 99%, respectively, while its inhibition efficiency against Ca3(PO4)2 scale reached 75% and 89%, respectively. The maximum inhibition efficiency (100%) to both CaCO3 and Ca3(PO4)2 was achieved at a scale inhibitor dosage of 4 mg/L. Compared to PASP, PASP-ASP, and PASP-MEA, PASP-ASP-MEA had better inhibition performance against CaCO3 and Ca3(PO4)2 scales. The reasons may be listed as follows. Firstly, the sparingly soluble salts, such as CaCO3, Ca3(PO4)2, and gypsum, are crystallized on the nano/microdust impurities in the aqueous solution [30]. Those nano/microdusts are not uniform and consist of different ingredients. The antiscalants PASP and their derivatives can block the nano/microdust crystallization centers of the salts. Therefore, the rate of crystallization of sparingly soluble salts decreases. Secondly, the grafted polymer has a better sorption on nano/microdust surface than a non-grafted one [31]. Because the nano/microdusts are not uniform, it is reasonable that some fractions of the dusts are better blocked by PASP-ASP and others by PASP-MEA. Thus, synergy becomes inevitable due to PASP-ASP-MEA combining the properties of PASP-ASP and PASP-MEA.
The CaCO3 and Ca3(PO4)2 scale deposits were observed by scanning electron microscope (SEM). As shown in Figure 5 and Figure 6, in the absence of PASP and the modified PASPs, CaCO3 scale deposits show a calcite structure with regular shape and glossy surface, and Ca3(PO4)2 scale deposits have an irregular polygon-shaped structure, not parallel to the surface to bulk. In comparison, when PASP and the modified PASPs were introduced into the solution, the shapes of CaCO3 and Ca3(PO4)2 scale deposits were irregular, and their crystalline grain tended to be finer. Both PASP and the modified PASPs (PASP-ASP, PASP-MEA and PASP-ASP-MEA) showed extremely good inhibition effects on CaCO3 and Ca3(PO4)2 scales. Particularly, PASP-ASP-MEA showed the smallest crystal scale deposits, and it had the best inhibition effects on CaCO3 and Ca3(PO4)2 scales. The synergy of the introduced hydroxyl and carboxylic groups of PASP-ASP-MEA can enhance the blocking of the nano/microdust crystallization centers of CaCO3 and Ca3(PO4)2. Therefore, the rate of crystallization of CaCO3 and Ca3(PO4)2 decreased, which finally resulted in the smaller crystals [30,31,32].
XRD spectra of CaCO3 crystals are shown in Figure 7. There are strong diffraction peaks of 26.2°, 27.2°, 33.1°, 36.2°, 37.3°, 45.9°, 50.2°, and 51.9° in Figure 7a, which are characteristic peaks of aragonite, and there are also diffraction peaks of 29.4° and 43.2° for calcite. These results indicate that in the absence of PASP and the modified PASPs, the calcium carbonate precipitate is the mixture of aragonite, which is the main crystal form, and some calcite [33]. In the other spectra (Figure 7b–e), diffraction peaks (24.9°, 27.1°, 32.8°, 43.9°, and 50.1°) corresponding to vaterite are very strong, which demonstrates that vaterite is the main crystal form in the presence of PASP and the modified PASPs [34]. Obviously, the vaterite peaks are the weakest for PASP-ASP-MEA. The change of crystal forms indicates that PASP-ASP-MEA contributed to the distortion of CaCO3 crystals and showed the best inhibition effect on CaCO3 scales.
The XRD spectra for Ca3(PO4)2 crystals in the absence and in the presence of PASP and the modified PASPs (PASP-ASP, PASP-MEA and PASP-ASP-MEA) are shown in Figure 8. In spectrum a, there are strong diffraction peaks at the 22.9°, 29.3°, 36.0°, 39.3°, 43.0°, 47.4°, 48.3°, 57.5°, 60.6°, 64.6° and 65.5°, which are characteristic peaks of Ca3(PO4)2 crystals [35], and the peak at 29.3° is obviously reduced in spectra b, c, d, and e. The addition of PASP and the modified PASPs (PASP-ASP, PASP- MEA, and PASP-ASP-MEA) showed a large influence on the crystal structure, and the diffraction peaks became quite weak after the addition of PASP and the modified PASPs, which implies that the surface morphology and particle size changed in the presence of the inhibitor. The diffraction peaks are the weakest for PASP-ASP-MEA, which shows that PASP-ASP-MEA had the best inhibition effect on Ca3(PO4)2 scales.
According to the above SEM images and XRD analysis, the introduction of the modified PASP disturbed the crystal growth habits and distorted the lattice, which resulted changes in the crystal morphology of the precipitates. The synergy of the introduced functional groups of PASP-ASP-MEA made the scale crystal become much smaller compared with the effects of other inhibitors. Consequently, the scales become floppy and can be removed easily.

2.4. Biodegradation Performance of the Modified PASP

The biodegradation property of the modified PASP was investigated, and the results are shown in Figure 9. Compared with the pure PASP, the modified PASP showed a decreased biodegradation rate. This may be attributed to the formed cross-linked chemical bonds in the modified PASP [36,37]. However, it still experienced an easy degradation process; more than 60 and 70 wt% was lost after 21 and 28 days of biodegradation, respectively. When the polymers were catalyzed by the bacteria, the cross-linking chain and their backbone were both degraded and ruptured. They were first degraded to oligomers and smaller molecules and then biocatalyzed to carbon dioxide and water [38,39]. The modified PASP showed good biodegradable performance; it is indubitably an environmentally friendly scale inhibitor.

3. Experiment

3.1. Materials

L-aspartic acid was obtained from Changmao Biochemical Engineering Co., Ltd. (Changzhou, China). Ethanol, sodium hydrate, phosphoric acid, monoethanolamine (MEA), bromoethane, ethyl acetate, N,N-dimethylformamide (DMF), ethylene diamine tetraacetic acid (EDTA), calcium chloride, potassium dihydrogen phosphate, sodium bicarbonate, and sodium borate were purchased from Sinopharm Chemical Reagent Co., Ltd. (Shanghai, China) and were used as received without further purification. N-methylimidazole was supplied by Shanghai Aladdin Bio-Chem Technology Co., Ltd. (Shanghai, China). EmimH2PO4 was synthesized in our laboratory (Scheme 1) [40]. Deionized water (ultrafiltered to 18 MW×cm using a Milipore Milli-Q gradient system) was used in all experiments.

3.2. Characterization

Fourier transform infrared spectrometry (FTIR) was carried out on a Perkin-Elmer SpectrumGX Fourier transform infrared spectrometer (Waltham, MA, USA). Proton NMR spectra were recorded on a Bruker Ascend 400 MHz nuclear magnetic resonance (NMR) spectrometer (Bruker, Billerica, MA, USA) using D2O as solvent. Gel chromatography (GPC) (PL-GPC50, UK) was used to determine the molecular mass of polymers. Scanning electron microscopy (SEM, HITACHI TM3000, Tokyo, Japan) was used to capture the crystal structure of CaCO3 and Ca3(PO4)2. X-ray diffraction (XRD) was used to analyze the crystal form of CaCO3 and Ca3(PO4)2.

3.3. Synthesis of Polyaspartic Acid (PASP) in Ionic Liquid

L-aspartic acid (5 g) was dissolved in a three-necked flask. Ionic liquid 1-ethyl-3-methylimidazol dihydrogen phosphate (EmimH2PO4) (15 mL) was slowly added into the flask, and the mixture was reacted at 180 °C for 3 h. Afterwards, the mixture was poured into anhydrous ethanol to form the precipitate. Then, light yellow solid product was obtained. The product polysuccinimide (PSI) was filtrated and dry.
PSI (5 g) and 10 wt% hydroxide solution were added into a round bottom flask together. The mixture was stirred at 40 °C for 1 h. Then, the solution was filtered, washed with absolute alcohol, and dried. Finally, the reddish-brown product was obtained. The relevant synthetic reaction is expressed in Scheme 2.

3.4. Ring-Opening Graft Modification of PASP

PSI (2 g) and NH3-NH4Cl buffer solution (pH = 10, 20 mL) were mixed at 0 °C. Then, ASP (0.055 g, 0.044 g, 0.027 g, 0.011 g) and MEA (0.025 g, 0.020 g, 0.013 g, 0.005 g) were added to the mixture together. The reaction was carried out at 0 °C for 24 h under stirring. The reddish-brown viscous solid was obtained after the precipitate was washed with ethanol. The relevant synthetic reaction is expressed in Scheme 3. High performance liquid chromatography (HPLC) was used to detect the surplus amount of ASP, and gas chromatography (GC) was used to detect the surplus amount of MEA.
The grafting ratio (GD, %) was calculated by the following Equation (1):
G D = m 1 m 2 m 1
where m1 is the feed amount of ASP or MEA and m2 is the surplus amount of ASP or MEA.

3.5. Measurement of the Efficiency of Static Scale Inhibition

Static scale inhibition tests were performed according to Chinese National Standards GB/T 16632-2008 and GB/T 22626-2008. The experimental condition of scale inhibition to CaCO3 was CaCl2 (240 mg·L−1) mixed with NaHCO3 (732 mg·L−1). The experimental condition of scale inhibition of Ca3(PO4)2 was CaCl2 (100 mg·L−1) mixed with KH2PO4 (5 mg·L−1). Borax buffer solution was used as the initial solution (pH = 9.0, 0.01 mol·L−1). Both brines passed filtration (220 nm filter membrane) before use. The reaction was carried out at 80 °C for 10 h with various amounts of scale inhibitors. When the reaction was finished, it was cooled to room temperature and filtered by filter paper. The filtrate was titrated with ethylene diamine tetraacetic acid (EDTA) standard solution to determine the concentration of Ca2+ (CaCO3).
The inhibition efficiency of CaCO3 was calculated by the following Equation (2):
η 1 = ρ 2 ρ 1 ρ 0 ρ 1 × 100 %
where ρ0 is the concentration of Ca2+ before experiment, ρ1 is the concentration of Ca2+ in the absence of scale inhibitor in the solution, and ρ2 is the concentration of Ca2+ in the presence of scale inhibitor in the solution.
The PO43− concentration in the filtrate was detected by the ammonium molybdate spectrophotometric method. The inhibition efficiency of Ca3(PO4)2 was calculated by the following Equation (3):
η 2 = ρ 2 ρ 1 ρ 0 ρ 1 × 100 %
where ρ0 is the concentration of PO43− before experiment, ρ1 is the concentration of PO43− in the absence of scale inhibitor, and ρ2 is the concentration of PO43− in the presence of inhibitor.

3.6. Biodegradation of the Modified PASP

The biodegradability of the modified PASP was estimated based on Chinese National Standard GB/T 21803-2008. The sample was treated with the standard activated sludge, which was obtained from Nanjing High Tech University Biological Technology Research Institute Co. Ltd., at 30 ± 1 °C for 28 days. The concentration of the activated sludge was 3~5 g/L. The concentration of the modified PASP was 100 mg/L. Sodium acetate was used as the standard molecule to estimate the activity of the standard activated sludge. The chemical oxygen demand (COD) detection was used to evaluate the biodegradability of the modified PASP. The COD was measured using a thermostat, DRB200 (Hach Co. Ltd., Loveland, CO, USA) and a spectrophotometer, DR1010 (Hach Co. Ltd., Loveland, CO, USA).

4. Conclusions

A new PASP-based green scale inhibitor (PASP-ASP-MEA) was obtained by ring-opening graft modification of PASP with both aspartic acid (ASP) and monoethanolamine (MEA). The modified PASP had excellent inhibition efficiency against CaCO3 and Ca3(PO4)2 scales. When the concentration of PASP-ASP-MEA was increased to 2 mg/L, its inhibition efficiency increased to 99% against CaCO3 and 89% against Ca3(PO4)2. Inhibition efficiency of 100% against both CaCO3 and Ca3(PO4)2 was achieved at a scale inhibitor dosage of 4 mg/L. The functional groups of PASP-ASP-MEA have good sorption on nano/microdust surface. PASP-ASP-MEA can therefore block the nano/microdust crystallization centers of CaCO3 and Ca3(PO4)2 and significantly decrease the crystallization rate of these sparingly soluble salts. Compared with PASP-ASP and PASP-MEA, the synergy of the introduced groups (-OH and -CO3) of PASP-ASP-MEA inevitably led better blocking of nano/microdusts. Thus, the scale crystals of CaCO3 and Ca3(PO4)2 became much smaller, increasing the solubility of these calcium salts in water. Moreover, the modified PASP also shows good biodegradable performance. In short, it is a promising green scale inhibitor.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/catal11070802/s1, Figure S1: Proton NMR spectra of PASP and PASP derivatives: (a) PASP, (b) PASP-ASP, (c) PASP-MEA, and (d) PASP-ASP-MEA. (For PASP-ASP-MEA, the molar ratio of PSI–ASP–MEA was 1:0.5:0.5, reaction temperature was 0 °C, and reaction time is 24 h).

Author Contributions

Formal Analysis, Y.Z.; Investigation, Y.Z. and Y.F.; Writing—Original Draft Preparation, Y.Z.; Writing—Review & Editing, J.W.; Supervision, Y.F.; Project Administration, Y.F.; Funding Acquisition, Y.F. Validation, Y.F. and Y.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China, grant number 21805135; the Jiangsu Province Natural Science Foundation for Youths, grant number BK20180712 and the Postgraduate Research & Practice Innovation Program of Jiangsu Province (SJCX21_0546); The APC was funded by the National Natural Science Foundation of China and the Jiangsu Province Natural Science Foundation for Youths.

Acknowledgments

We are grateful for the technical discussions with Yuzhen Li for the measurement of the morphology of the Efficiency of Static Scale Inhibition, and Xiao Chen for the analysis of the biodegradation.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Antony, A.; Low, J.H.; Gray, S.; Childress, A.E.; Le-Clech, P.; Leslie, G. Scale formation and control in high pressure membrane water treatment systems: A review. J. Membr. Sci. 2011, 383, 1–16. [Google Scholar] [CrossRef]
  2. Wolowiec, M.; Komorowska-Kaufman, M.; Pruss, A.; Rzepa, G.; Bajda, T. Removal of heavy metals and metalloids from water using drinking water treatment residuals as adsorbents: A review. Minerals 2019, 9, 17. [Google Scholar] [CrossRef] [Green Version]
  3. Guo, Y.R.; Hu, Y.; Shi, K.; Bilan, Y. Valuation of water resource green efficiency based on SBM-TOBIT panel model: Case study from henan province, China. Sustainability 2020, 12, 6944. [Google Scholar] [CrossRef]
  4. Zhang, S.P.; Qu, H.J.; Yang, Z.; Fu, C.E.; Tian, Z.Q.; Yang, W.B. Scale inhibition performance and mechanism of sulfamic/amino acids modified polyaspartic acid against calcium sulfate. Desalination 2017, 419, 152–159. [Google Scholar] [CrossRef]
  5. Yu, W.; Song, D.; Chen, W.; Yang, H. Antiscalants in RO membrane scaling control. Water Res. 2020, 183, 115985. [Google Scholar] [CrossRef]
  6. Yin, Y.M.; Jeong, N.; Minjarez, R.; Robbins, C.A.; Carlson, K.H.; Tong, T.Z. Contrasting behaviors between gypsum and silica scaling in the presence of antiscalants during membrane distillation. Environ. Sci. Technol. 2021, 55, 5335–5346. [Google Scholar] [CrossRef] [PubMed]
  7. Feng, J.Y.; Gao, L.J.; Wen, R.Z.; Deng, Y.Y.; Wu, X.J.; Deng, S.L. Fluorescent polyaspartic acid with an enhanced inhibition performance against calcium phosphate. Desalination 2014, 345, 72–76. [Google Scholar] [CrossRef]
  8. Chen, J.X.; Chen, F.J.; Han, J.; Su, M.; Li, Y.H. Evaluation of scale and corrosion inhibition of modified polyaspartic acid. Chem. Eng. Technol. 2020, 43, 1048–1058. [Google Scholar] [CrossRef]
  9. Chen, Y.Y.; Zhou, Y.M.; Yao, Q.Z.; Nan, Q.L.; Zhang, M.J.; Sun, W. Synthesis of modified polyepoxysuccinic acid and evaluation of its scale inhibition on CaCO3, CaSO4, and Ca3(PO4)2 precipitation for industrial recycling water. Desalin. Water Treat. 2019, 152, 16–25. [Google Scholar] [CrossRef]
  10. Hasson, D.; Shemer, H.; Sher, A. State of the art of friendly “green” scale control inhibitors: A review article. Ind. Eng. Chem. Res. 2011, 50, 7601–7607. [Google Scholar] [CrossRef]
  11. Li, H.S.; Liu, W.; Qi, X.J. Evaluation of a novel CaSO4 scale inhibitor for a reverse osmosis system. Desalination 2007, 214, 193–199. [Google Scholar] [CrossRef]
  12. Zhu, T.Z.; Wang, L.D.; Sun, W.; Yang, Z.Q.; Wang, S.L.; Zhao, L.Q.; Xiao, G.J.; Wang, G.Z.; Liu, Z.M.; Shu, X.Q.; et al. Local scaling of CaCO3 on carbon steel surface with different corrosion types. Powder Technol. 2019, 356, 990–1000. [Google Scholar] [CrossRef]
  13. Mady, M.F.; Bayat, P.; Kelland, M.A. Environmentally friendly phosphonated polyetheramine scale inhibitors-excellent calcium compatibility for oilfield applications. Ind. Eng. Chem. Res. 2020, 59, 9808–9818. [Google Scholar] [CrossRef]
  14. Can, H.K.; Uner, G. Water-soluble anhydride containing alternating copolymers as scale inhibitors. Desalination 2015, 355, 225–232. [Google Scholar] [CrossRef]
  15. Martinod, A.; Neville, A.; Euvrad, M.; Sorbie, K. Electrodeposition of a calcareous layer: Effects of green inhibitors. Chem. Eng. Sci. 2009, 64, 2413–2421. [Google Scholar] [CrossRef]
  16. Mazumder, M.A.J. A review of green scale inhibitors: Process, types, mechanism and properties. Coatings 2020, 10, 928. [Google Scholar] [CrossRef]
  17. Zhang, Z.J.; Lu, M.L.; Liu, J.; Chen, H.L.; Chen, Q.L.; Wang, B. Fluorescent-tagged hyper-branched polyester for inhibition of CaSO4 scale and the scale inhibition mechanism. Mater. Today Commun. 2020, 25, 101359. [Google Scholar] [CrossRef]
  18. Migahed, M.A.; Rashwan, S.M.; Kamel, M.M.; Habib, R.E. Synthesis, characterization of polyaspartic acid-glycine adduct and evaluation of their performance as scale and corrosion inhibitor in desalination water plants. J. Mol. Liq. 2016, 224, 849–858. [Google Scholar] [CrossRef]
  19. Ketrane, R.; Saidani, B.; Gil, O.; Leleyter, L.; Baraud, F. Efficiency of five scale inhibitors on calcium carbonate precipitation from hard water: Effect of temperature and concentration. Desalination 2009, 249, 1397–1404. [Google Scholar] [CrossRef]
  20. Ling, L.; Zhou, Y.M.; Huang, J.Y.; Yao, Q.Z.; Liu, G.Q.; Zhang, P.X.; Sun, W.; Wu, W.D. Carboxylate-terminated double-hydrophilic block copolymer as an effective and environmental inhibitor in cooling water systems. Desalination 2012, 304, 33–40. [Google Scholar] [CrossRef]
  21. Nudelman, F.; Pieterse, K.; George, A.; Bomans, P.H.H.; Friedrich, H.; Brylka, L.J.; Hilbers, P.A.J.; de With, G.; Sommerdijk, N. The role of collagen in bone apatite formation in the presence of hydroxyapatite nucleation inhibitors. Nat. Mater. 2010, 9, 1004–1009. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Nemethy, A.; Solti, K.; Kiss, L.; Gyarmati, B.; Deli, M.A.; Csanyi, E.; Szilagyi, A. pH- and temperature-responsive poly(aspartic acid)-l-poly (N-isopropylacrylamide) conetwork hydrogel. Eur. Polym. J. 2013, 49, 2392–2403. [Google Scholar] [CrossRef]
  23. Xu, Y.; Wang, L.N.; Zhao, L.L.; Cui, Y.C. Synthesis of polyaspartic acid-aminobenzenesulfonic acid grafted copolymer and its scale inhibition performance and dispersion capacity. Water Sci. Technol. 2011, 64, 423–430. [Google Scholar] [CrossRef]
  24. Zhang, Q.; Vigier, K.D.O.; Royer, S.; Jérôme, F. Deep eutectic solvents: Syntheses, properties and applications. Chem. Soc. Rev. 2012, 41, 7108–7146. [Google Scholar] [CrossRef]
  25. Wang, B.S.; Qin, L.; Mu, T.C.; Xue, Z.M.; Gao, G.H. Are ionic liquids chemically stable? Chem. Rev. 2017, 117, 7113–7131. [Google Scholar] [CrossRef]
  26. Reddy, P.N.; Padmaja, P.; Reddy, B.V.S.; Rambabu, G. Ionic liquid/water mixture promoted organic transformations. RSC Adv. 2015, 5, 51035–51054. [Google Scholar] [CrossRef]
  27. Coxon, J.M.; Townsend, M.A.E. Computational study on the ring-opening reaction of protonated oxirane and methylpropene. Tetrahedron 2007, 63, 5665–5668. [Google Scholar] [CrossRef]
  28. Quan, Z.H.; Chen, Y.C.; Wang, X.R.; Shi, C.; Liu, Y.J.; Ma, C.F. Experimental study on scale inhibition performance of a green scale inhibitor polyaspartic acid. Sci. China Ser. B Chem. 2008, 51, 695–699. [Google Scholar] [CrossRef]
  29. Tang, Y.M.; Yang, W.Z.; Yin, X.S.; Liu, Y.; Yin, P.W.; Wang, J.T. Investigation of CaCO3 scale inhibition by PAA, ATMP and PAPEMP. Desalination 2008, 228, 55–60. [Google Scholar] [CrossRef]
  30. Oshchepkov, M.; Golovesov, V.; Ryabova, A.; Frolova, S.; Tkachenko, S.; Kamagurov, S.; Rudakova, G.; Popov, K. Synthesis and visualization of a novel fluorescent-taggedpolymeric antiscalant during gypsum crystallization in combination with bisphosphonate fluorophore. Crystals 2020, 10, 992. [Google Scholar] [CrossRef]
  31. Oshchepkov, M.; Golovesov, V.; Ryabova, A.; Tkachenko, S.; Redchuk, A.; Ronkkomaki, H.; Rudakova, G.; Pervov, A.; Popov, K. Visualization of a novel fluorescent-tagged bisphosphonate behavior during reverse osmosis desalination of water with high sulfate content. Sep. Purif. Technol. 2021, 255, 117382. [Google Scholar] [CrossRef]
  32. Chai, C.X.; Xu, Y.H.; Li, D.Y.; Zhao, X.W.; Xu, Y.; Zhang, L.; Wu, Y.F. Cysteamine modified polyaspartic acid as a new class of green corrosion inhibitor for mild steel in sulfuric acid medium: Synthesis, electrochemical, surface study and theoretical calculation. Prog. Org. Coat. 2019, 129, 159–170. [Google Scholar] [CrossRef]
  33. Liu, D.; Dong, W.B.; Li, F.T.; Hui, F.; Ledion, J. Comparative performance of polyepoxysuccinic acid and polyaspartic acid on scaling inhibition by static and rapid controlled precipitation methods. Desalination 2012, 304, 1–10. [Google Scholar] [CrossRef]
  34. Wang, H.C.; Zhou, Y.M.; Yao, Q.Z.; Ma, S.S.; Wu, W.D.; Sun, W. Synthesis of fluorescent-tagged scale inhibitor and evaluation of its calcium carbonate precipitation performance. Desalination 2014, 340, 1–10. [Google Scholar] [CrossRef]
  35. Naciri, Y.; Hsini, A.; Ajmal, Z.; Bouddouch, A.; Bakiz, B.; Navio, J.A.; Albourine, A.; Valmalette, J.C.; Ezahri, M.; Benlhachemi, A. Influence of Sr-doping on structural, optical and photocatalytic properties of synthesized Ca3(PO4)2. J. Colloid Interface Sci. 2020, 572, 269–280. [Google Scholar] [CrossRef] [PubMed]
  36. Nakato, T.; Tomida, M.; Suwa, M.; Morishima, Y.; Kusuno, A.; Kakuchi, T. Preparation and characterization of dodecylamine-modified poly(aspartic acid) as a biodegradable water-soluble polymeric material. Polym. Bull. 2000, 44, 385–391. [Google Scholar] [CrossRef]
  37. Juriga, D.; Nagy, K.; Jedlovszky-Hajdu, A.; Perczel-Kovach, K.; Chen, Y.M.; Varga, G.; Zrinyi, M. Biodegradation and osteosarcoma cell cultivation on poly(aspartic acid) based hydrogels. ACS Appl. Mater. Interfaces 2016, 8, 23463–23476. [Google Scholar] [CrossRef] [PubMed]
  38. Huang, K.; Kong, L. Preparation and characterization of poly(aspartic acid) derivatives as biodegradable water treatment agents. Asian J. Chem. 2013, 25, 10233–10237. [Google Scholar] [CrossRef]
  39. Yang, J.H.; Liu, T.; Liu, H.B.; Zhang, D.; Zhai, L.M.; Liu, J.; Wang, M.; Chen, Y.X.; Chen, B.Y.; Wang, H.Y. Biodegradable PASP can effectively inhibit nitrification, moderate NH3 emission, and promote crop yield. Arch. Agron. Soil Sci. 2019, 65, 1273–1286. [Google Scholar] [CrossRef]
  40. Dong, F.; Jun, L.; Zhou, X.L.; Liu, Z.L. Mannich reaction in water using acidic ionic liquid as recoverable and reusable catalyst. Catal. Lett. 2007, 116, 76–80. [Google Scholar] [CrossRef]
Figure 1. Schematic illustration of the new scale inhibitor based on ring-opening graft modification of polyaspartic acid for water treatment.
Figure 1. Schematic illustration of the new scale inhibitor based on ring-opening graft modification of polyaspartic acid for water treatment.
Catalysts 11 00802 g001
Figure 2. There are three typical influencing factors which affect graft ratio of ASP and MEA. (a) Influence of molar ratio of PSI–ASP–MEA on the graft ratio of ASP and MEA (reaction time is 24 h, reaction temperature is 0 °C); (b) influence of reaction time on the graft ratio of ASP and MEA (molar ratio of PSI–ASP–MEA is 1:0.5:0.5, reaction temperature is 0 °C); (c) influence of reaction temperature on the graft ratio of ASP and MEA (molar ratio of PSI–ASP–MEA was 1:0.5:0.5 and reaction time was 24 h).
Figure 2. There are three typical influencing factors which affect graft ratio of ASP and MEA. (a) Influence of molar ratio of PSI–ASP–MEA on the graft ratio of ASP and MEA (reaction time is 24 h, reaction temperature is 0 °C); (b) influence of reaction time on the graft ratio of ASP and MEA (molar ratio of PSI–ASP–MEA is 1:0.5:0.5, reaction temperature is 0 °C); (c) influence of reaction temperature on the graft ratio of ASP and MEA (molar ratio of PSI–ASP–MEA was 1:0.5:0.5 and reaction time was 24 h).
Catalysts 11 00802 g002
Figure 3. FTIR spectrum of PASP and PASP derivatives: (a) PASP, (b) PASP-ASP, (c) PASP-MEA, and (d) PASP-ASP-MEA. (For PASP-ASP-MEA, the molar ratio of PSI–ASP–MEA was 1:0.5:0.5, reaction temperature was 0 °C, and reaction time was 24 h).
Figure 3. FTIR spectrum of PASP and PASP derivatives: (a) PASP, (b) PASP-ASP, (c) PASP-MEA, and (d) PASP-ASP-MEA. (For PASP-ASP-MEA, the molar ratio of PSI–ASP–MEA was 1:0.5:0.5, reaction temperature was 0 °C, and reaction time was 24 h).
Catalysts 11 00802 g003
Figure 4. The inhibition efficiency of PASP and the modified PASPs (PASP-ASP, PASP-MEA, and PASP-ASP-MEA) against CaCO3 scales (a) and Ca3(PO4)2 scales (b). (For PASP-ASP-MEA, the molar ratio of PSI–ASP–MEA was 1:0.5:0.5, reaction temperature was 0 °C, and reaction time was 24 h).
Figure 4. The inhibition efficiency of PASP and the modified PASPs (PASP-ASP, PASP-MEA, and PASP-ASP-MEA) against CaCO3 scales (a) and Ca3(PO4)2 scales (b). (For PASP-ASP-MEA, the molar ratio of PSI–ASP–MEA was 1:0.5:0.5, reaction temperature was 0 °C, and reaction time was 24 h).
Catalysts 11 00802 g004
Figure 5. SEM images of CaCO3 with PASP and its derivatives: (a) blank, (b) PASP, (c) PASP-ASP, (d) PASP-MEA, (e) PASP-ASP-MEA. (For PASP-ASP-MEA, the molar ratio of PSI–ASP–MEA was 1:0.5:0.5, reaction temperature was 0 °C, and reaction time was 24 h).
Figure 5. SEM images of CaCO3 with PASP and its derivatives: (a) blank, (b) PASP, (c) PASP-ASP, (d) PASP-MEA, (e) PASP-ASP-MEA. (For PASP-ASP-MEA, the molar ratio of PSI–ASP–MEA was 1:0.5:0.5, reaction temperature was 0 °C, and reaction time was 24 h).
Catalysts 11 00802 g005
Figure 6. SEM images of Ca3(PO4)2 with PASP and its derivatives: (a) blank, (b) PASP, (c) PASP-ASP, (d) PASP-MEA, (e) PASP-ASP-MEA. (For PASP-ASP-MEA, the molar ratio of PSI–ASP–MEA was 1:0.5:0.5, reaction temperature was 0 °C, and reaction time was 24 h).
Figure 6. SEM images of Ca3(PO4)2 with PASP and its derivatives: (a) blank, (b) PASP, (c) PASP-ASP, (d) PASP-MEA, (e) PASP-ASP-MEA. (For PASP-ASP-MEA, the molar ratio of PSI–ASP–MEA was 1:0.5:0.5, reaction temperature was 0 °C, and reaction time was 24 h).
Catalysts 11 00802 g006
Figure 7. The XRD patterns of the CaCO3 crystals: (a) blank, (b) PASP, (c) PASP-ASP, (d) PASP-MEA, (e) PASP-ASP-MEA. (For PASP-ASP-MEA, the molar ratio of PSI–ASP–MEA was 1:0.5:0.5, reaction temperature was 0 °C, and reaction time was 24 h).
Figure 7. The XRD patterns of the CaCO3 crystals: (a) blank, (b) PASP, (c) PASP-ASP, (d) PASP-MEA, (e) PASP-ASP-MEA. (For PASP-ASP-MEA, the molar ratio of PSI–ASP–MEA was 1:0.5:0.5, reaction temperature was 0 °C, and reaction time was 24 h).
Catalysts 11 00802 g007
Figure 8. The XRD patterns of the Ca3(PO4)2 crystals: (a) blank, (b) PASP, (c) PASP-ASP, (d) PASP-MEA, (e) PASP-ASP-MEA. (For PASP-ASP-MEA, the molar ratio of PSI–ASP–MEA was 1:0.5:0.5, reaction temperature was 0 °C, and reaction time was 24 h).
Figure 8. The XRD patterns of the Ca3(PO4)2 crystals: (a) blank, (b) PASP, (c) PASP-ASP, (d) PASP-MEA, (e) PASP-ASP-MEA. (For PASP-ASP-MEA, the molar ratio of PSI–ASP–MEA was 1:0.5:0.5, reaction temperature was 0 °C, and reaction time was 24 h).
Catalysts 11 00802 g008
Figure 9. Degradation of PASP and the modified PASPs. (For PASP-ASP-MEA, the molar ratio of PSI–ASP–MEA was 1:0.5:0.5, reaction temperature was 0 °C, and reaction time was 24 h).
Figure 9. Degradation of PASP and the modified PASPs. (For PASP-ASP-MEA, the molar ratio of PSI–ASP–MEA was 1:0.5:0.5, reaction temperature was 0 °C, and reaction time was 24 h).
Catalysts 11 00802 g009
Scheme 1. Synthesis route of EmimH2PO4.
Scheme 1. Synthesis route of EmimH2PO4.
Catalysts 11 00802 sch001
Scheme 2. Synthesis of PASP in ionic liquid.
Scheme 2. Synthesis of PASP in ionic liquid.
Catalysts 11 00802 sch002
Scheme 3. Ring-opening graft modification of PASP.
Scheme 3. Ring-opening graft modification of PASP.
Catalysts 11 00802 sch003
Table 1. Molecular mass of PASP and the modified PASP calculated by GPC.
Table 1. Molecular mass of PASP and the modified PASP calculated by GPC.
PSI-ASP-MEAMn (kDa)Mw (kDa)Mw/Mn
1:0:04.877.931.63
1:1:04.968.091.63
1:0.8:0.24.948.191.66
1:0.5:0.54.907.891.61
1:0.2:0.84.917.801.59
1:0:14.927.921.61
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhou, Y.; Wang, J.; Fang, Y. Green and High Effective Scale Inhibitor Based on Ring-Opening Graft Modification of Polyaspartic Acid. Catalysts 2021, 11, 802. https://doi.org/10.3390/catal11070802

AMA Style

Zhou Y, Wang J, Fang Y. Green and High Effective Scale Inhibitor Based on Ring-Opening Graft Modification of Polyaspartic Acid. Catalysts. 2021; 11(7):802. https://doi.org/10.3390/catal11070802

Chicago/Turabian Style

Zhou, Yongsheng, Jie Wang, and Yan Fang. 2021. "Green and High Effective Scale Inhibitor Based on Ring-Opening Graft Modification of Polyaspartic Acid" Catalysts 11, no. 7: 802. https://doi.org/10.3390/catal11070802

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