DOI:
10.1039/C6RA03635J
(Paper)
RSC Adv., 2016,
6, 31019-31030
Alternate cyclopolymer of diallylglutamic acid and sulfur dioxide
Received
8th February 2016
, Accepted 18th March 2016
First published on 21st March 2016
Abstract
The monomer N,N-diallylglutamic acid hydrochloride [(CH2
CH–CH2)2NH+CH(CO2H)CH2CH2CO2H·Cl−] (I)/SO2, underwent cyclopolymerization to afford the alternate cationic polyelectrolyte (CPE) poly(I-alt-SO2) (II) in very good yields. Upon depletion of HCl during soaking in water, water-soluble triprotic (+) II was transformed to water-insoluble diprotic zwitterionic (±) III containing residues of glutamic acid in each repeating unit. Upon treatment with 1 and 2 equivalents of NaOH, polyzwitterions (±) III, was converted into water-soluble monoprotic poly(zwitterion-anion) (±−) IV and the fully deprotonated polydianion (=) V, respectively. Basicity strengths of the chelation centers of CO2− and nitrogen in (=) V have been determined. The polymer demonstrated its effectiveness as an antiscalant in the inhibition of CaSO4 scaling. Keeping in view the pH-dependent solubility of the polymers, a recyclable aqueous two-phase system (ATPS) comprising of (=) V and polyethylene glycol has been constructed for the potential purification of biomolecules. After its use, the ATPS can be recycled by precipitating the ionic polymer in the form of (±) III at a lower pH.
1. Introduction
Glutamic acid is a non-essential amino acid that is used in the biosynthesis of proteins (Scheme 1). Its side chain CO2H with a pKa of 4.1 exists almost entirely in the CO2− form in the physiological pH range 7.35–7.45. The carboxylate anion in glutamate is an important excitatory neurotransmitter in neural activation.1,2 All meat, fish, dairy products, wheat, etc., are excellent sources of glutamic acid. Glutamate, an abundantly available molecule in the brain, is involved in learning and memory,3 it is also a key compound in cellular metabolism.4 A series of disodium sulphonamides of glutamic acid complexes with copper(II), nickel(II) and ruthenium(III) ions, have demonstrated good anticancer activities.5
 |
| Scheme 1 Glutamic acid, α- and γ-poly(glutamic acid) and polymer containing glutamic acid residues. | |
The biodegradable, non-toxic and non-immunogenic properties of naturally occurring poly-γ-glutamic acid (γ-PGA) have rendered its use in the food, medical and wastewater industries (Scheme 1).6–8 α-PGA has been utilized in many medical applications including drug delivery6 and cancer treatment.9 γ-Glutamic acid enhances calcium absorption in the small intestine and increases the solubility of Ca2+, suggesting its chelating ability.10,11 Glutamic acid-derived polymers have the potential to act as polychelatogenes to scavenge toxic metal ions.12 The cross-linking of microbial γ-PGA with glucose has led to a hydrogel which has been shown to be a superadsorbent of water (3000 g g−1).13 A biodegradable PGA/gadolinium chelate has been evaluated as a contrast agent for magnetic resonance imaging (MRI); this is an important development since most of the currently evaluated macromolecular contrast agents are not biodegradable.14 The recently developed Fe3O4–PGA nanoparticle has been shown to hold great promise to be used as a contrast agent for MRI of tumors.15 γ-PGA-functionalized alumina nanoparticles (γ-PAN) has been evaluated for cytotoxicity towards human prostate cancer cell PC-3; the study provides a basis for further screening of the promising material for future biomedical applications.16
The abundant availability of inexpensive glutamic acid has led us to synthesize and cyclopolymerize17–19 diallylglutamic acid 4 to obtain poly(diallylglutamic acid) 5 which retained all three original functionalities of the amino acid (Scheme 1).20 Note that the nitrogen in the peptide bond in PGA loses its basic character while retaining only one of the original carboxylate motifs of the amino acid. Glutamic acid and materials derived from it are of tremendous importance for biomedical and material research. We anticipated to use 5 as a polymer component in developing a recyclable aqueous two-phase system (ATPS) for its utilization in bioseparation.21,22 However, failure to synthesize 5 having high molar mass and its solubility behavior (vide infra) impeded the development of a recyclable ATPS in which its pH-controlled solubility would permit precipitating out and hence reuse the polymer.
The current work envisages the application of copolymerization protocol23 to obtain 4/SO2 alternate copolymer poly(4-alt-SO2) 8 (Scheme 2) with the anticipation that its high molar mass and pH-dependent solubility behavior would permit us to develop a recyclable ATPS. The study would also permit us to evaluate the effects of SO2 spacer on the basicity constants of various ligand centers and antiscalant properties. The current study would pave the way to synthesize cross-linked adsorbents containing the metal chelating centers of glutamic acid for the removal of toxic metal ions.23–25
 |
| Scheme 2 Synthesis of cyclopolymers containing glutamic acid residues. | |
2. Experimental
2.1. Materials
L-Glutamic acid from Fluka AG (Buchs, Switzerland) was used as received. Sodium 3-trimethylsilylpropionate-2,2,3,3-d4 (TSP-deuterated), an NMR internal standard, was purchased from Merck Sharp & Dohme Canada Ltd, Montreal, Canada. 2,2′-Azoisobutyronitrile (AIBN) from Fluka AG (Buchs, Switzerland) was crystallized from C2H5OH/CHCl3 mixture. Calcium hydride-dried dimethylsulfoxide (DMSO) was distilled at a boiling point of 64–65 °C (4 mmHg). Poly(ethylene glycol) (PEG) of number average molecular weight (
n) of 35
000 was purchased from MERCK-Schuchardt.
Dimethyl N,N-diallylglutamate (7), synthesized from dimethyl glutamate (6), was hydrolyzed with NaOH and acidified with HCl to obtain N,N-diallylglutamic acid hydrochloride 4 as described.20
2.2. Physical methods for structural characterization
A Perkin Elmer (Series II Model 2400) elemental analyzer and a Fourier transform infrared (FTIR) spectrometer (Perkin Elmer 16F PC) were utilized for elemental analyses and IR spectroscopy, respectively. The nuclear magnetic resonance (NMR) spectra were recorded using a 500 MHz JEOL LA spectrometer. Tetramethylsilane (TMS) in CDCl3 and TSP-deuterated in D2O were used as internal standards with 1H signal at δ 0 ppm. The 13C chemical shifts in D2O were referenced against 13C peak of external standard dioxane at δ 67.4 ppm. A pH meter (Sartorius PB 11) and an Orion Versa Star benchtop meter (Thermoscientific) were used to measure pH and conductivity, respectively. Viscosity was measured under N2 in an Ubbelohde viscometer (constant = 0.005317 mm2 s−2) using polymer solution prepared in CO2-free water. An SDT analyzer (Q600: TA Instruments) was utilized to carry out thermogravimetric analysis (TGA) under N2.
2.3. Procedure for 4/SO2 copolymerization
The conditions of polymerizations are summarized in Table 1. For instance, initiator AIBN was added under N2 to a homogeneous mixture of monomer 4 (3.96 g, 15 mmol) and SO2 (960 mg, 15 mmol) in DMSO (3.75 g) in a 25 mL-RB flask. After stirring the contents in the closed flask at 60 °C for 48 h, the viscous polymer mixture was precipitated in water, filtered and dried at 60 °C under vacuum to a constant weight of copolymer polyzwitterion acid (PZA) 9. Thermal decomposition: stable up to 240 °C, slight phase change and expansion between 240 and 245 °C, turned tan at 280 °C with evolution of gas which continue up to 400 °C leaving behind dark particles (found: C, 45.0; H, 6.1; N, 4.7; S, 10.8. C11H17NO6S requires C, 45.35; H, 5.88; N, 4.81; S, 11.01%); νmax (KBr): 3400 (v br), 2978, 1724, 1625, 1405, 1306, 1126, 878, 812, 767, 640 and 512 cm−1.
Table 1 Cyclocopolymerization of 4/SO2 at 63 °C for 36 h
Entry |
Monomer (mmol) |
SO2 (g) (mmol) |
DMSO (g) |
AIBNa (mg) |
Yield (%) |
[η]b dL g−1 |
Azobisisobutyronitrile. Intrinsic viscosity of 1–0.0625% polymer 9 treated with 2 equivalents NaOH in 0.1 M NaCl at 30 °C was measured with an Ubbelohde Viscometer (K = 0.005317 mm2 s−2). |
1 |
15 |
15 |
3.75 |
45 |
48 |
0.873 |
2 |
2 × 15 |
2 × 15 |
2 × 3.75 |
2 × 75 |
82 |
1.03 |
3 |
15 |
15 |
3.75 |
105 |
87 |
1.22 |
2.4. Conversion of 9 into 11 by NaOH
A mixture of PZA 9 (1.46 g, 5.0 mmol) in water (10 cm3) at 0 °C was treated with NaOH (0.4 g, 10 mmol). Another portion of NaOH (0.2 g, 5 mmol) was added and the solution was quickly dropped onto acetone (75 cm3). The resultant white polydianion electrolyte (PDE) 11 was filtered and washed with MeOH/acetone 1
:
1 mixture. The polymer was dried under vacuum till constant weight (1.6 g, 91%). Thermal decomposition: stable up to 290 °C, slight phase change and expansion between 290 and 325 °C, turned tan at 330 °C with evolution of gas which continued up to 400 °C leaving behind dark particles. νmax (KBr): 3400 (v br), 2962, 2829, 1579 (v br), 1448, 1409, 1300, 1123, 891, 827, 786, and 514 cm−1 (found: C, 37.1; H, 4.9; N, 3.9; S, 8.8. C11H15NNa2O6S·1H2O requires C, 37.40; H, 4.85; N, 3.96; S, 9.07%).
2.5. Solubility measurements
An aqueous solution of PZA 9 (1% w/w) containing NaBr or NaI or HCl were titrated salt-free water at 23 °C until turbidity appeared. Triplicate measurements gave critical (minimum) salt concentration (CSC) of 2.16 M NaBr, 0.581 M NaI, and 8.59 M HCl. PZA 9 was found to be practically insoluble in any possible concentration of NaCl.
2.6. Potentiometric titrations
The procedure for the determination of basicity constants (K) is described elsewhere.26,27 A certain millimole (in terms of repeating unit (RU)) of water-insoluble PZA 9 (ZH2±) was dissolved by treating with 3 equivalents of NaOH using a 0.0978 M NaOH, then diluted with CO2-free water to 200 mL, which thus contained a solution of PDE 11 in the presence of 1 equivalent excess NaOH. For water-soluble PDE 11, its solution was directly prepared in salt-free water (200 mL). The solutions were then titrated by gradual addition of 0.1222 M HCl (0.05–0.15 mL) as described in Table 2. The pH values, recorded after each addition of the titrant, were used to calculate log
Ki by the Henderson–Hasselbalch eqn (2) (see eqn (1)–(3) embedded in Scheme 2). The insolubility of PZA 9 did not permit the determination of log
K3 associated with the equilibration: 9 (ZH2±) + H+
(ZH3+) 8.
Table 2 Protonation of polymer 11 (Z=) and 9 (ZH±) at 23 °C in salt-free water
Run |
ZH2± (mmol) |
CTa (mol dm−3) |
α-Range |
pH-Range |
Pointsb |
log Koic |
n1c |
R2d |
(+)ve values describe titrations with HCl. Number of data points. Standard deviations in the last digit are given under the parentheses. R = correlation coefficient. log Ki = log Koi + (ni − 1)log[(1 − α)/α]. Titration was carried out in the presence of 6.43 and 7.50 mL of added 0.0978 M NaOH, respectively, to solubilize and attain the required values of the α. Titration was carried out in the presence of 6.43 and 7.50 mL of added 0.0978 M NaOH, respectively, to solubilize and attain the required values of the α. |
Polymer 9 (ZH±) or 11 (Z=) in salt-free water |
1 |
0.2007 (Z=) |
+0.1222 |
0.23–0.91 |
10.37–7.26 |
22 |
9.35 |
2.09 |
0.9942 |
2 |
0.2066f (ZH2±) |
+0.1222 |
0.18–0.88 |
10.83–7.60 |
22 |
9.30 |
2.07 |
0.9945 |
3 |
0.2413g (ZH2±) |
+0.1222 |
0.15–0.88 |
10.80–7.50 |
24 |
9.33 |
2.04 |
0.9950 |
Average |
|
|
|
|
|
9.33(3) |
2.07(3) |
|
log K1e = 9.33 + 1.07 log[(1 − α)/α] |
![[thin space (1/6-em)]](https://www.rsc.org/images/entities/char_2009.gif) |
1 |
0.2007 (Z=) |
+0.1222 |
0.16–0.66 |
6.36–4.93 |
17 |
5.40 |
1.45 |
0.9934 |
2 |
0.2066f (ZH2±) |
+0.1222 |
0.12–0.75 |
6.60–4.74 |
20 |
5.45 |
1.40 |
0.9943 |
3 |
0.2413g (ZH2±) |
+0.1222 |
0.09–0.87 |
6.85–4.20 |
23 |
5.49 |
1.46 |
0.9953 |
Average |
|
|
|
|
|
5.45(5) |
1.44(3) |
|
log K1e = 5.45 + 0.44 log[(1 − α)/α] |
The log
K1 and log
K2 represent the basicity constants of the basic centers in PDE 11 (Z=) and polyzwitterion-anion (PZAN) 10 (ZH±−), respectively. The degree of protonation (α) of 11 and 10 is calculated by [ZH±−]eq/[Z]o, and [ZH2±]eq/[Z]o, respectively, where [ZH±−]eq and [ZH2±]eq represent the corresponding equilibrium concentrations of the protonated species 10 and 9. [Z]o represents the initial polymer concentration in terms of RUs.
Because of addition of 8–11 mL of 0.0978 M NaOH, which is more than two equivalents of the RUs in 9 (ZH2±), the polymer is converted to its dianionic form 11 (Z=) by neutralization with two equivalents of NaOH. By considering the excess NaOH as added OH−, the concentration of the protonated species 10 [ZH±−] during the titration of 11 (Z=) with HCl to determine log
K1 was given by [ZH±−]eq = CH+ − COH− − [H+] + [OH−], where COH− represents the concentration of the added ‘excess NaOH’. The equilibrium concentrations of [H+] and [OH−] were determined using the pH values, whereas CH+ represents the concentration of added HCl during titrations. Continuing the titration, log
K2 was calculated using titrant volume after subtracting one-equivalent, from the total volume.
2.7. Evaluation of antiscalant behavior
The inhibition of CaSO4 scale formation were evaluated using supersaturated solution of CaSO4 containing Ca2+ (2600 ppm) and SO42− (6300 ppm) in the presence of newly synthesized antiscalant 9 (x ppm) (Table 1, entry 2) (treated with minimum quantity of NaHCO3 to dissolve it) at 40 ± 1 °C using the procedure described elsewhere.28 The concentrations of the ions are three times the concentrations found in the reject brine of a reverse osmosis plant.29 A rapid decrease in conductivity happened at the induction time indicating the beginning of precipitation of CaSO4 (Table 3). Visual inspections for any turbidity were performed.
Table 3 Percent inhibition against precipitation at various times in the presence of various concentrations of the synthesized polymer 9 in a supersaturated CaSO4 solution at 40 °C
Entry |
Sample (ppm) |
Percent inhibition at times (min) of |
Induction time (min) |
30 |
60 |
90 |
120 |
150 |
500 |
1000 |
No induction observed on the studied time range. |
1 |
5 |
94 |
67 |
15 |
8.1 |
7.1 |
7.3 |
6.4 |
60 |
2 |
10 |
98 |
91 |
82 |
71 |
38 |
8.5 |
8.5 |
90 |
3 |
20 |
99 |
99 |
99 |
98 |
98 |
77 |
12 |
600 |
4 |
30 |
100 |
100 |
100 |
100 |
100 |
100 |
99 |
—a |
2.8. Phase compositions of PZA 9 (+2 equivs NaOH)–PEG–H2O (NaCl) systems
2.8.1. The tie-lines by 1H NMR spectroscopy. Several ATPSs of volume ≈ 7 cm3 having varying compositions of component polymers PZA 9 (+2.0 equivalents of NaOH) (Table 1, entry 2) and PEG were prepared in calibrated cylinders using their stock solutions (20 wt%) in 0.6 M NaCl and 0.6 M NaCl as the diluents. After centrifuging and equilibrating at 23 °C for 24 h, volume and density of the top and bottom layers were measured. The composition of the polymers in the phases were determined by 1H NMR spectra of the D2O exchanged phases in the presence of K2CO3 (vide infra).
2.8.2. Binodals by turbidity method. The binodal of the PZA 9 (+2 equivs NaOH)–PEG–H2O (0.6 M NaCl) system was constructed by turbidimetric method at 23 °C as described elsewhere.30
3. Results and discussions
3.1. Monomer and polymer synthesis
Dimethyl glutamate (6) upon alkylation with allyl bromide afforded dimethyl N,N-diallylglutamate (7) which upon alkaline hydrolysis followed by acidification gave monomer N,N-diallylglutamic acid hydrochloride (4).25 Monomer 4 underwent AIBN-initiated cyclopolymerization with SO2 to give CPE (+) 8 which upon depletion of HCl during soaking in water was transformed to its water-insoluble zwitterionic form: PZA (±) 9 in 87% yield (Table 1, entry 3). Increasing concentration of the initiator increases both the yields and viscosities of the polymer. The intrinsic viscosities of water-insoluble (±) 9 was measured in 0.1 M NaCl in the presence of two equivalents of NaOH which converts it to water-soluble PDE (=) 11 and are reported in Table 1. An analytical sample of 11 was prepared by treating PZA 9 (Table 1, entry 2) with excess NaOH in 91% yield. Note that both the monomer and polymers retain the unquenched nitrogen valency as well as the two carboxyl groups of the glutamic acid.
The TGA curve of PZA 9 is shown in Fig. 1; a minor loss of 6% up to 200 °C is attributed to the loss of trapped moisture. An accelerated loss of 24% in the range 200–250 °C could be accounted by the loss of the SO2 units; note that the calculated mass of SO2 amounts to 22.0%. The gradual loss of 48% in the range 250–800 °C may be associated with the removal of the glutamate pendant containing CO2 units. At 800 °C, remaining mass of 22% belonged to some nitrogenated organic fraction. As can be seen from the TGA plot, the polymer remained stable up to 200 °C.
 |
| Fig. 1 TGA curve of PZA 9 (entry 2, Table 1). | |
Attempts to determine the molar masses of polymer 11 were unsuccessful owing to the strong adsorption aided by the chelating ligands amine and carboxy of the polymer to the materials of the GPC column.31
3.2. Infrared and NMR spectra
The presence of SO2 in the backbone of polymers 9 and 11 was confirmed by the appearance of strong IR bands at ≈1305 and 1125 cm−1 assigned to its asymmetric and symmetric stretching, respectively. For the dipolar motifs32 in (±) 9, the symmetric and asymmetric vibrations of COO− and the C
O stretching of COOH were revealed at 1405, 1625 and 1724 cm−1, respectively. The absorption bands at 1409 and 1579 cm−1 were attributed to the symmetric and asymmetric stretching of COO− in (=) 11, respectively.
The NMR spectra (Fig. 2 and 3) of monomer 4 and polymers 9 and 11 revealed the absence of residual double bonds in the macromolecules. The finding suggests that the chain termination happened via degradative chain transfer to the monomer33 as well as by coupling process.34 Integration of the relevant carbon signals revealed a 67
:
33 cis/trans ratio of the ring substituents at Cb,b (Scheme 1; Fig. 3c).35,36
 |
| Fig. 2 1H NMR spectra of (a) 4, (b) 9 (+NaI) and (c) 11 in D2O (referenced using signal of trimethylsilylpropionate-2,2,3,3-d4 (TSP) at δ 0 ppm of as internal standard). | |
 |
| Fig. 3 13NMR spectra of (a) 4, (b) 9 (+NaI) and (c) 11 in D2O (referenced using δ 67.4 ppm of dioxane as external standard). | |
3.3. Solubility behavior
Like the majority of known polyzwitterions,37–39 water-insoluble (±) 9 becomes soluble with the addition of various salts of small molar masses. For NaBr and NaI, the CSCs at 23 °C were determined to be 2.16 and 0.581 M, respectively. Iodide ions, being the most polarizable (soft), effectively neutralize the positive nitrogens so as to disrupt the intragroup, intra- and interchain ionic crosslinks and thereby, enhance solubility in water.18 Zwitterionic (±) 9 was found to be soluble in 8.59 M HCl as a result of shifting the mobile equilibrium: (±) 9 + H+
(+) 8 toward right where the zwitterionic interactions required for water-insolubility vanishes. The presence of any concentration of NaCl was not able to disrupt the zwitterionic interactions and promote solubility of polymer (±) 9. This is an interesting solubility behavior which could be exploited in the development of recyclable ATPS (vide infra). Polyelectrolytes PZAN 10 and PDE 11 with charge imbalances are found to be water-soluble; anionic portion in (±−) 10 overcomes the zwitterionic interactions so as to impart water-solubility. It has been observed during the potentiometric titrations that the polymer solutions become cloudy at pH below ≈4.6, whereby the polymer backbone is calculated to have a (±) 9/(±−) 10 ratio of ≈80
:
20. Thus, increasing the zwitterionic portion to more than 80% leads to the polymer's insolubility in salt-free water.
3.4. Viscosity measurements
Polyzwitterion (±) 9 is expected to have the lowest viscosity; however it cannot be measured in the presence of usual salt NaCl since the polymer is insoluble (vide supra). Fig. 4a and b displays the viscosity plots for (±−) 10 and (=) 11 (derived from entry 2, Table 1). The viscosity plots for the polymers in salt-free water were concave upwards like any polyelectrolytes (Fig. 4a-i and ii). However, at lower concentration regime, the viscosity values fall off as a result of progressive shifting of the mobile equilibria: [(=) 11 or (±−) 10] + H2O
[(±−) 10 or (±) 9] + OH− towards right. As per general rule of hydrolysis, the degree of transformation of 11 to 10 (or 10 to 9) increases with decreasing concentration; overall decrease in the charge imbalance on the polymer chains decreases, thereby leading to lesser electrostatic repulsions hence lesser viscosity values.
 |
| Fig. 4 (a) The viscosity behavior of sample derived entry 2, Table 1 using an Ubbelohde viscometer at 30 °C: (i) ■ (±−) 10 and (ii) □ (=) 11 in salt-free water; (b) (i) □ (=) 11 and (ii) ■ (±−) 10 in 0.1 M NaCl; (c) plot for the apparent (i) log K1 versus degree of protonation (α) (entry 3, Table 2) and (ii) log K2 versus α for PDA 11 (derived from entry 2, Table 1) in salt-free water (entry 3, Table 2); (d) reduced viscosity (ηsp/C) at 30 °C of a 0.00858 M (i.e. 0.25 g dL−1) solution of polymer PZA 9 (ZH2±) in salt-free water (•) versus equivalent of added NaOH. Distribution curves (dashed lines) of the various ionized species [■ 9 (ZH2±), □ 10 (ZH±−), Δ 11 (Z=)] calculated using eqn (2) (Scheme 2) and pH of the solutions in salt-free water at 23 °C. | |
The higher viscosity values for zwitterionic/anionic (±−) 10 (Fig. 4a-i) than that of its dianionic counterpart (=) 11 (Fig. 4a-ii), as confirmed by careful triplicate measurements in salt-free water under N2, is indeed unexpected. The greater repulsion among the (−)ve charges in PDE (=) 11 having the highest degree of charge asymmetry is supposed to impart higher viscosity values. A possible rationale for the higher viscosity of (±−) 10 could be attributed to the interchain H-bonding leading to the higher hydrodynamic volume as shown in Scheme 3. Note that the more distant negative oxygens of dianions (=) 11 may impart less effective repulsions. However, in 0.1 M NaCl, the interchain hydrogen bonding is discouraged because of the neutralization of the (+)ve charges in (±−) 10 by Cl− ions. The viscosity plots in 0.1 M NaCl were found to be in line with the expectation (Fig. 4b-i versus Fig. 4b-ii), whereby the greater charge imbalances in (=) 11 dictating the viscosity values.
 |
| Scheme 3 Interchain H-bonding leading to increased hydrodynamic volume. | |
3.5. Basicity constants
The pH vs. log[(1 − α)/α] linear regression (eqn (2), Scheme 2) furnished the ‘ni’ as the slope and log
Koi as the intercept. The basicity constants (Ki) of the anionic centers in the polymers 9–11 is given by eqn (3) (Scheme 2) where log
Koi = pH at α = 0.5 and ni = 1 for sharp basicity constants. In salt-free water, log
K1 of the amine group in (=) 11, which is the pK1 of its conjugate acid (±−) 10, and log
K2 of the terminal CO2− in (±−) 10 (i.e. pK2 of its conjugate acid (±) 9) were determined to be 9.33 and 5.45, respectively, in salt-free water (Table 2). log
K3 (i.e. pK3) involving the equilibrium: 9 (ZH2±) + H+
(ZH3+) 8 cannot be determined owing to the insolubility of zwitterionic (±) 9 (vide supra). Note that log
K of a base is the pKa of its conjugate acid.
The nis, which reflect a measure of polyelectrolyte effect, are found to be greater than 1 thereby indicating a gradual decrease of basicity constants (K) with increasing degree of protonation (α) (Table 2). The n1 and n2, associated with K1 and K2, respectively, were determined to be 2.07 and 1.44, thereby reflecting a greater polyelectrolyte effect i.e. greater changes in K1 values during the transformation of (=) 11 to (±−) 10. In salt-free water, charge centers in (=) 11 are expected to be more hydrated than in (±−) 10. The higher values of log
K1 and n1 than log
K2 and n2, respectively, are the consequences of the entropy driven protonation step.40 During protonation, a repeating unit (RU) of (=) 11 would lose greater number of hydrated water molecules than that of (±−) 10. Note that polyzwitterion (±) 9, being the most compacted and least hydrated as confirmed by its insolubility in water. As a consequence, during protonation of (=) 11 and (±−) 10, the respective polymer backbones would consist of (=) 11/(±−) 10 and (±−) 10/(±−) 9 with the former having greater number of water molecules in the hydration shell of each RU. The variations of log
Kis with α, shown in Fig. 4c, reflect their “apparent”41,42 nature since instead of remaining constant, they decrease with the increase in α. It has been established40 that the protonation process in similar cases is entropy-driven as a result of the release of water molecules from the hydration shell of the repeating unit that is being protonated. However, the exothermic enthalpy changes (ΔHos) remain constant with increasing α.40 Likewise, in both cases of protonation involving K1 and K2 for the current work, the exothermic enthalpy changes (ΔHos) are expected to remain constant with increasing α, and the ΔGos become less negative (i.e. less favorable) as a result of progressive decrease in the (+)ve ΔSos since ΔGo = ΔHo − TΔSo. With each protonation, the (−)ve charge density and the hydrated water molecules per RU decrease; as a consequence, a RU being protonated releases less water molecules from its hydration shell than that of the unit protonated in the previous step.40 In other words, the continuous decrease of positive ΔSo with α leads to lesser negative ΔGos, hence a gradual decrease in log
K values.
3.6. Viscometric titrations
The viscosity data of a 0.00858 M (i.e. 0.25 g dL−1) PZA 9 in salt-free water in the presence of various concentration of added NaOH at 23 °C is reported using a solid line in Fig. 4d. The pH values and basicity constants (vide supra) were used to construct the distribution curves (dashed lines) of the specie ZH2± (PZA 9), ZH±− (PZAN 10), and Z= (PDE 11). The reduced viscosity, as expected, increases with the addition up to 1 equivalent of NaOH as a consequence of the transformation of ZH2± to ZH±−. Maximum viscosity is achieved after the addition of 1 equivalent of NaOH whereby the concentration of ZH2± and ZH±− reaches the minimum and maximum values, respectively. Further addition of NaOH, which initiates the transformation of ZH±− to Z=, led to a gradual decrease in the reduced viscosities accompanied by an increase in the concentration of the ionic species Z=. This unusual viscosity behavior of dianionic 11 (Z=) having lower viscosity than its zwitterionic–anionic counterpart 10 (ZH±−) has also been observed in Fig. 4a and a rationale for the finding is given (vide supra).
3.7. Scale inhibition by the synthesized polymer
Scaling of inorganic salts like CaSO4 and CaCO3 in reverse osmosis (RO) process damages membranes, thereby impeding their smooth functioning. Excessive presence of the relevant ions in the reject brine leads to their supersaturation and scaling. The percent scale inhibition (PSI) is calculated using eqn (4): |
 | (4) |
where [Ca2+] represents the concentrations in the presence and absence of antiscalant 9 at time zero (t0) and t.
In the current work, a supersaturated solutions of CaSO4 containing Ca2+ (2600 ppm) and SO42− (6300 ppm) in the presence of 0 (blank), 5, 10, 20, and 30 ppm of 9 as an antiscalant were examined by following the conductivity of the solutions versus time. The results of the investigation are given in Table 3. In the absence of 9, a sudden drop in conductivity indicates the precipitation of CaSO4 (Fig. 5-iv: blank). To our satisfaction, the presence of 9 at a concentration of 30 ppm imparted a 99% scale inhibition for about 1000 min, while it was 100% at the time of 500 min as calculated using eqn (4) using conductivity values as proportional to the ionic concentrations. Note that the presence of 5, 10, 20, and 30 ppm of the antiscalant was able to register PSI of 94, 98, 99, and 100%, respectively, at a time of 30 min. These are indeed efficient PSI values since the feed water usually resides for ≈30 min in the osmosis chamber. The time, at which a sharp drop in conductivity happens, signaling the onset of quick precipitation, is known as an induction period (IP). The IPs of 60, 90 and 600 min were observed in the presence of 5, 10 and 20 ppm of 9, respectively. Note that at the concentration of 30 ppm and the time scale of 1000 min, no induction period was observed; it imparted 99% inhibition. The new antiscalant has demonstrated its efficacy in scavenging metal ions and disrupting the nucleation and crystallization processes,43,44 hence it could be used as a potential antiscalant to minimize membrane fouling by CaSO4 scale.
 |
| Fig. 5 Precipitation behavior of a supersaturated solution of CaSO4 in the presence (5, 10, 15, 30 ppm) and absence of PZA 9 (treated with NaHCO3). Inset showing the precipitation behavior in a shorter time scale of 100 min. | |
3.8. Phase diagrams: [PZA 9 + 2 equivs NaOH]–PEG–H2O (0.6 M NaCl)
The polymer compositions in the phases, determined by 1H NMR spectroscopy, were used to construct the tie-lines in the phase diagram (Table 4). A small portion of each phase was taken, and after exchanging H2O with D2O the NMR spectra were recorded in the presence of K2CO3 which helped to minimize the overlap of PEG signal at δ 3.65 with the signals of PDE 11 (i.e. 9 + 2 equivs NaOH). The proton spectrum of the bottom phase (system 1, Table 4) is displayed in Fig. 6a; the four Hs marked e, f of PDE 11 appeared in the range δ 1.6–2.2 ppm having an area integration of A, while the remaining eleven protons marked a–d in the range δ 2.4–3.6 ppm integrated for an area of B. The mole ratios of the polymers 11/PEG was determined as A/C where C is the area of the four-proton signals for PEG at δ 3.65 ppm. For the top phase (Fig. 6b), where the excessive presence of PEG led to overlapping of its signal with that of PDE 11, mole ratio of 11/PEG was calculated as [(A + C) − (11 × B/4)]/B, where (A + C) accounts for the combined areas of four-proton PEG and eleven-proton PDE 11 while C is equated to (11 × B/4). The respective molar masses of 291.3 and 44.05 g mol−1 for each RU of 9 and PEG were used for calculation of the weight fractions (w) which were used to construct the tie lines as described30 (Fig. 7a).
Table 4 Composition of the phases of the [POEa + PZAb] system (2.0 equiv. NaOH, 0.6 M NaCl) at 296 K shown in Fig. 7a
NMR method |
System |
Total system |
Top phase |
Bottom phase |
Volume ratioc |
PEG w × 100 |
PZA w × 100 |
PEG w × 100 |
PZA w × 100 |
PEG w × 100 |
PZA w × 100 |
Poly(oxyethylene) of molar mass 35 000 g mol−1. 9. Volume ratio of top and bottom phase. |
1 |
3.07 |
4.46 |
7.70 |
0.101 |
0.128 |
7.14 |
0.634 |
2 |
3.68 |
3.15 |
6.76 |
0.182 |
0.254 |
6.43 |
1.06 |
3 |
2.48 |
3.63 |
5.56 |
0.282 |
0.354 |
5.83 |
0.683 |
4 |
2.67 |
2.65 |
4.84 |
0.282 |
0.454 |
5.02 |
1.01 |
5 |
1.94 |
2.81 |
3.66 |
0.557 |
0.831 |
4.13 |
0.619 |
Turbidity method |
System |
Binodal data |
System |
Binodal data |
PEG w × 100 |
PZA w × 100 |
PEG w × 100 |
PZA w × 100 |
a |
0.332 |
4.25 |
h |
2.48 |
0.823 |
b |
0.457 |
3.55 |
i |
2.92 |
0.657 |
c |
0.603 |
2.97 |
j |
3.32 |
0.553 |
d |
0.897 |
2.37 |
k |
3.88 |
0.451 |
e |
1.38 |
1.74 |
l |
4.58 |
0.380 |
f |
1.70 |
1.40 |
m |
6.06 |
0.254 |
g |
2.05 |
1.09 |
|
|
|
 |
| Fig. 6 1H NMR spectrum of (a) bottom layer and (b) top layer of (System 5, Table 4) in D2O (+K2CO3). | |
 |
| Fig. 7 (a) Phase diagram [■ and □ represent data obtained by respective NMR and turbidimetric method] at 296 K of 0.6 M NaCl containing PZA 9 (treated with 2 equivs NaOH)–PEG at 296 K; (b) correlation of phase diagram of PZA 9 (2 equiv. NaOH)–PEG–H2O (0.6 M NaCl). | |
The binodal, constructed using turbidity method, is shown in Fig. 7a. The tie lines are drawn by connecting Atotal, Atop and Abottom which represents the composition of the total system, PEG-rich top phase and PZA-rich bottom phases, respectively. The ratio of the length of the tie-line segments (Atotal − Abot)/(Atotal − Atop) is equated to Vtop/Vbottom where V represents the volume of the phases.21 The composition on a tie-line therefore determines the ratio of the phase volumes.
A single- and two-phase region is demarcated by a binodal curve whose position with respect to the axes determines the economic viability for industrial application. As shown in Fig. 7a, the binodal for the ATPS is very much closer to the axes requiring only a total polymer concentrations of ≈5% for the separation of the phases. The most beneficial aspect of the environmentally friendly ATPS is the solubility of PDE (=) 11 at a higher pH, while it can be recycled21 at acidic pH by precipitating it in the form of (±) 9. The current ionic ATPS containing the polymer component having pH-triggerable functionalities (N and CO2−) of glutamic acid is anticipated to have pH responsive behavior that would impart selectivity in bioseparation.
3.9. Correlation of the phase diagram
Diamond and Hsu45 developed eqn (5) and (6) based on Flory–Huggins theory to check the consistency of the tie-lines. |
ln K1 = A1(w′′1 − w′1)
| (5) |
and |
ln K2 = A2(w′′2 − w′2)
| (6) |
where w′′ and w′ represent the wt% of polymer 1 (PZA 9 + NaOH) and polymer 2 (PEG), in the respective top and bottom phase. The ln
K − (w′′i − w′i) plots with zero intercept gave the slopes A1 and A2 which reflect the effects of molar mass of the polymers and their interactions with water. The partition coefficients K1 and K2 of polymer 1 and polymer 2 were determined by their concentration ratio (Ct/Cb) in the top and bottom layer. The straight-line fits in Fig. 7b certifies the ability of the model to describe the phase behavior. Eqn (7)46 describes the relationship between the root mean-square deviation (rmsd) with the Kexp and Kcal: |
 | (7) |
where N represent the number of tie lines (Fig. 7a). The values of the parameters Ai and (rmsd)i in Table 5 ascertain the usefulness of equations in correlating the experimental data. In this model, an extensive phase equilibrium determination is avoided, since a single phase composition would suffice to determine A1 and A2.
Table 5 Values for parameter Aa in eqn (5) and (6) along with the rmsdb of the model from the experimental data of the partition coefficients
Entry |
System |
PZAc |
PEGd |
A1a |
rmsd |
A2a |
rmsd |
Ai is the slope of the linear regression of ln Ki versus w′′i − w′i, with zero intercept value (eqn (5) and (6)). Root mean-square deviation. PZA 9. Poly(oxyethylene) of molar mass 35.0 kg mol−1. |
1 |
POE + PZA + 2.0 equiv. NaOH |
−0.581 |
0.00587 |
0.537 |
3.28 |
3.10. A comparison of homopolymer 5 and copolymer 9
Several properties of poly(diallylglutamic acid) (5) and copolymer poly(diallylglutamic acid-alt-SO2) (9) are compared in Table 6. The lower basicity constant K1 of the copolymer 11 is attributed to the electron-withdrawing ability of the SO2 in its RU. Since log
K1 = pKNH+, the conjugate acid 10 has thus greater acidity than that of the corresponding homopolymer. Copolymer 11 is found to have much higher intrinsic viscosity (1.03 dL g−1) than its corresponding homopolymer (0.160 dL g−1) (Table 6). Homopolymer 5 is found to be a more effective antiscalant than copolymer 9; in the presence of 20 ppm each of 5 and 9, the scale inhibition was found to be 100 and 77%, respectively, after the elapse of 500 min. The better performance of the homopolymer 5 may be attributed to its lower molar mass; polymer with a smaller size may interfere more severely with the growth of scale by efficient poisoning of its active sites.47
Table 6 Basicity constants Ki and scale inhibition efficiency of homo- and co-polymer
Polymer |
pKNH+aor log K1 |
pKCO2Ha or log K2 |
[η]b (dL g−1) |
Scale inhibitionc (%) |
CSCd |
NaCl (M) |
NaBr (M) |
NaI (M) |
HCl (M) |
In salt-free water at 23 °C. Intrinsic viscosity of 1–0.0625% polymer 5 (ref. 20) and 9 (Table 1, entry 2) treated with 2 equivalents NaOH in 0.1 M NaCl at 30 °C was measured with an Ubbelohde Viscometer (K = 0.005317 mm2 s−2). After 500 min using 20 ppm polymer in a supersaturated solution of CaSO4 (aq.) at 40 °C. Critical salt concentration required to promote solubility at 23 °C. Insoluble in the presence of any concentration of NaCl. |
 |
10.9 |
5.25 |
0.160 |
100 |
0.548 |
0.271 |
0.133 |
0.0104 |
 |
9.93 |
5.45 |
1.03 |
77 |
Insolublee |
2.16 |
0.581 |
8.59 |
The solubility of the polymers differs greatly; the copolymer with much higher CSC values demonstrated stronger zwitterionic interactions (Table 6). The greater dispersion of the positive charges in the zwitterionic dipoles of 9 by electron-withdrawing SO2 makes it less hydrated and thus more attracted to the negative end of the dipole. The solubility behavior of polymer 9 makes it an attractive component in a pH-controlled recyclable ATPS.
4. Conclusions
Monomer 4 containing residues of glutamic acid and SO2 on alternate cyclopolymerization afforded CPE 8 [i.e. poly(4-alt-SO2)] in excellent yields. CPE 8 on depletion of HCl during dialysis gave water-insoluble PZA 9 whose solubility is promoted by the addition of NaI, NaBr and HCl. PZA (±) 9 upon neutralization with 1 and 2 equivalents of NaOH was converted to water-soluble (±−) 10 and (=) 11, which has been used in developing an environmentally friendly recyclable ATPS. The basicity constants Ki for the ligand centres in PDE (=) 11 have been determined. The polymer has demonstrated antiscalant properties and imparted excellent CaSO4-scale inhibition. Currently we are working on the synthesis of cross-linked resins containing residues of metal chelating glutamic acid for its use in the purification of wastewater.
Acknowledgements
The authors gratefully acknowledge the financial support provided by King Abdulaziz City for Science and Technology (KACST) through project No. AR-32-99 and the facilities provided by King Fahd University of Petroleum and Minerals.
References
- B. S. Meldrum, J. Nutr., 2000, 130, 1007–1015 Search PubMed.
- R. Sapolsky, Biology and Human Behavior: The Neurological Origins of Individuality, The Teaching Company, See pages 19 and 20 of Guide Book, 2nd edn, 2005 Search PubMed.
- W. J. McEntee and T. H. Crook, Psychopharmacology, 1993, 111, 391–401 CrossRef CAS PubMed.
- M. G. Erlander and A. J. Tobin, Neurochem. Res., 1991, 16, 215–226 CrossRef CAS PubMed.
- I. Ali, W. A. Wani, K. Saleem and M. F. Hsieh, RSC Adv., 2014, 4, 29629–29641 RSC.
- A. Ogunleye, A. Bhat, V. U. Irorere, D. Hill, C. Williams and I. Radecka, Microbiology, 2015, 161, 1–17 CrossRef CAS PubMed.
- I. Bajaj and R. Singhal, Bioresour. Technol., 2011, 102, 5551–5561 CrossRef CAS PubMed.
- I. L. Shih and Y. T. Van, Bioresour. Technol., 2001, 79, 207–225 CrossRef CAS PubMed.
- C. Li, D. F. Yu, A. Newman, F. Cabral, C. Stephens, N. R. Hunter, L. Milas and S. Wallace, Cancer Res., 1998, 58, 2404–2409 CAS.
- T. Tsujimoto, J. Kimura, Y. Takeuchi, H. Uyama, C. Park and M. H. Sung, J. Microbiol. Biotechnol., 2010, 20, 1436–1439 CrossRef CAS PubMed.
- C. Park, Y. H. Choi, H. J. Shin, H. Poo, J. J. Song, C. J. Kim and M. H. Sung, J. Microbiol. Biotechnol., 2005, 15, 855–858 CAS.
- M. R. Vakili, S. Zahmatkesh, M. V. Panahiyan and T. Jafarizadeh, Des. Monomers Polym., 2015, 18, 315–322 CrossRef CAS.
- S. Murakami and N. Aoki, Biomacromolecules, 2006, 7, 2122–2127 CrossRef CAS PubMed.
- X. Wen, E. F. Jackson, R. E. Price, E. Edmund Kim, Q. Wu, S. Wallace, C. Charnsangavej, J. G. Gelovani and C. Li, Bioconjugate Chem., 2004, 15, 1408–1415 CrossRef CAS PubMed.
- Z. Yu, C. Peng, Y. Luo, J. Zhu, C. Chen, M. Shen and X. Shi, RSC Adv., 2015, 5, 76700–76707 RSC.
- Y. C. Rajan, B. S. Inbaraj and B. H. Chen, RSC Adv., 2015, 5, 15126–15139 RSC.
- G. B. Butler, Cyclopolymerization and Cyclocopolymerization, Marcel Dekker, New York, 1992 Search PubMed.
- S. Kudaibergenov, W. Jaeger and A. Laschewsky, Adv. Polym. Sci., 2006, 201, 157–224 CrossRef CAS.
- P. K. Singh, V. K. Singh and M. Singh, e-Polym., 2007, 030, 1–34 Search PubMed.
- Z. A. Jamiu, H. A. Al-Muallem and S. A. Ali, Des. Monomers Polym., 2016, 19, 128–137 CrossRef CAS.
- P. Albertsson, Partition of Cell particles and Macromolecules, Wiley, N.Y., 3rd edn, 1986 Search PubMed.
- M. Waziri, B. F. Abu-Sharkh and S. A. Ali, Biotechnol. Prog., 2004, 20, 526–532 CrossRef PubMed.
- Z. A. Jamiu, T. A. Saleh and S. A. Ali, RSC Adv., 2015, 5, 42222–42232 RSC.
- S. A. Ali, O. C. S. Al-Hamouz and N. M. Hassan, J. Hazard. Mater., 2013, 248(249), 47–58 CrossRef PubMed.
- S. A. Ali, I. W. Kazi and N. Ullah, Ind. Eng. Chem. Res., 2015, 54, 9689–9698 CrossRef CAS.
- S. A. Ali, S. A. Haladu and H. A. Al-Muallem, J. Appl. Polym. Sci., 2014, 131, 40615 CrossRef.
- S. A. Ali, N. Y. Abu-Thabit and H. A. Al-Muallem, J. Polym. Sci., Part A: Polym. Chem., 2010, 48, 5693–5703 CrossRef CAS.
- S. A. Haladu and S. A. Ali, J. Polym. Sci., Part A: Polym. Chem., 2013, 51, 5130–5142 CrossRef CAS.
- F. H. Butt, F. Rahman and U. Baduruthamal, Desalination, 1995, 103, 189–198 CrossRef CAS.
- S. A. Ali, H. A. Al-Muallem and M. A. J. Mazumder, J. Chem. Eng. Data, 2013, 58, 2574–2585 CrossRef CAS.
- F. Rullens, M. Devillers and A. Laschewsky, Macromol. Chem. Phys., 2004, 205, 1155–1166 CrossRef CAS.
- J. F. Pearson and M. A. Slifkin, Spectrochim. Acta, Part A, 1972, 28, 2403–2413 CrossRef CAS.
- G. B. Butler and R. J. Angelo, J. Am. Chem. Soc., 1957, 79, 3128–3131 CrossRef CAS.
- R. M. Pike and R. A. Cohen, J. Polym. Sci., 1960, 44, 531–538 CrossRef CAS.
- W. Jaeger, J. Bohrisch and A. Laschewsky, Prog. Polym. Sci., 2010, 35, 511–577 CrossRef CAS.
- V. D. Vynck and E. J. Goethals, Macromol. Rapid Commun., 1997, 18, 149–156 CrossRef.
- T. A. Wielema and J. B. F. N. Engberts, Eur. Polym. J., 1987, 23, 947–950 CrossRef CAS.
- J. C. Salamone, W. Volksen, A. P. Olson and S. C. Israel, Polymer, 1978, 19, 1157–1162 CrossRef CAS.
- M. Yoshizawa and H. Ohno, Chem. Lett., 1999, 28, 889–890 CrossRef.
- R. Barbucci, M. Casolaro, P. Ferruti and M. Nocentini, Macromolecules, 1986, 19, 1856–1861 CrossRef CAS.
- R. Barbucci, M. Casolaro, N. Danzo, V. Barone, P. Ferruti and A. Angeloni, Macromolecules, 1983, 16, 456–462 CrossRef CAS.
- R. Barbucci, M. Casolaro, M. Nocentini, S. Correzzi, P. Ferruti and V. Barone, Macromolecules, 1986, 19, 37–42 CrossRef CAS.
- H. David, S. Hilla and S. Alexander, Ind. Eng. Chem. Res., 2011, 50, 7601–7607 CrossRef.
- R. J. Davey, The Role of Additives in Precipitation Processes, Industrial Crystallization, 81, ed. S. J. Jancic and E. J. de Jong, North-Holland Publishing Co., 1982, pp. 123–135 Search PubMed.
- A. D. Diamond and J. T. Hsu, AIChE Symp. Ser., 1992, 290, 105–111 CAS.
- S. Shahriari, S. G. Doozandeh and G. Pazuki, J. Chem. Eng. Data, 2012, 57, 256–262 CrossRef CAS.
- D. Hasson, H. Shemer and A. Sher, Ind. Eng. Chem. Res., 2011, 50, 7601–7607 CrossRef CAS.
|
This journal is © The Royal Society of Chemistry 2016 |
Click here to see how this site uses Cookies. View our privacy policy here.