Synthesis, characterization and reactivity ratios of poly N - CyberLeninka

0 downloads 0 Views 712KB Size Report
70 ± 1 °C and benzoyl peroxide as a free radical initiator. The copolymerization ... on the synthesis of N-monosubstituted methacrylamides. [24,8,7] and their ...
Journal of Saudi Chemical Society (2013) xxx, xxx–xxx

King Saud University

Journal of Saudi Chemical Society www.ksu.edu.sa www.sciencedirect.com

ORIGINAL ARTICLE

Synthesis, characterization and reactivity ratios of poly N-(p-bromophenyl)-2-methacrylamide-Co-N-vinyl-2pyrrolidone D. Thirumoolan a, K. Anver Basha a,*, Tapan Kanai b, S. Mohammed Safiullah a, K. Vetrivel a, K. Abdul Wasi a, B. Ranjithkumar a a b

P.G. & Research Department of Chemistry, C. Abdul Hakeem College, Hakeem Nagar, Melvisharam 632 509, Tamil Nadu, India Polymer Division, Naval Materials Research Laboratory, Add. Ambernath 421 506, India

Received 23 July 2013; revised 5 September 2013; accepted 10 September 2013

KEYWORDS N-(p-bromophenyl)-2-methacrylamide; N-vinyl-2-pyrrolidone; Monomer reactivity ratios; Thermo gravimetric analysis

Abstract The methacrylamide monomer, N-(p-bromophenyl)-2-methacrylamide (PBPMA) was synthesized by reacting p-bromoaniline dissolved in ethylmethylketone (EMK) with methacryloyl chloride in the presence of triethylamine. The copolymers of PBPMA with N-Vinyl-2-pyrrolidone (NVP) were synthesized by free radical solution polymerization using EMK as a solvent at 70 ± 1 C and benzoyl peroxide as a free radical initiator. The copolymerization behavior was studied in a wide composition interval with the mole fractions of PBPMA ranging from 0.20 to 0.80 in the feed. The copolymers were characterized by FT-IR, 1H NMR, 13C NMR and GPC. The solubility was tested in various polar and non polar solvents. The thermogravimetric analysis of the polymers showed that the thermal stability of the copolymer increases with PBPMA content. The copolymer composition was determined by elemental analysis. The monomer reactivity ratios were determined by the application of conventional linearization methods such as Fineman–Ross, Kelen–Tudos methods, and a non-linear error-in-variable model (EVM) method using a computer program. ª 2013 Production and hosting by Elsevier B.V. on behalf of King Saud University.

1. Introduction

* Corresponding author. Tel.: +91 9894066901; fax: +91 04172 269487. E-mail address: [email protected] (K. Anver Basha). Peer review under responsibility of King Saud University.

Production and hosting by Elsevier

Polymers with reactive functional groups are now being synthesized, characterized and used not only for their macromolecular properties, but also for their applications and the functional groups [1]. Several studies have been carried out on the synthesis of N-monosubstituted methacrylamides [24,8,7] and their radical copolymerization with commercial monomers. This study clearly shows that the nature as well as position of the substituent had a large effect on monomer

1319-6103 ª 2013 Production and hosting by Elsevier B.V. on behalf of King Saud University. http://dx.doi.org/10.1016/j.jscs.2013.09.003

Please cite this article in press as: D. Thirumoolan et al., Synthesis, characterization and reactivity ratios of poly N-(p-bromophenyl)-2methacrylamide-Co-N-vinyl-2-pyrrolidone, Journal of Saudi Chemical Society (2013), http://dx.doi.org/10.1016/j.jscs.2013.09.003

2 reactivity ratios, glass transition temperature, thermal stability and antimicrobial properties. Generally polymers and macrocyclic compounds having functional groups such as –NH2, –SH, and –COOH, are often required for applications such as sensors, actuators, smart windows, corrosion inhibitors, gene delivery, drug delivery or some medical applications. This is due to the heteroatom having a lone electron pair on its structure. Recently aromatic amines such as copolyaniline and pheylenediamines are used as sensors in various kinds of aqueous media in the environment [16,11]. N-vinyl-2-pyrrolidone (NVP) polymer has good properties such as biocompatibility, low toxicity, and filmforming and adhesive characteristics [24,4,17]. The literature also showed that polymers having ketone groups are used as photodegradable packing materials [3,4] and photoresist for microlithography. The understanding of copolymerization kinetics has gained great importance in recent decades. Because of this fact, the prediction of monomer reactivity ratios becomes a valuable quantitative aspect. Moreover, copolymerization is an important and useful way to develop new materials. Copolymerization modulates both intramolecular and intermolecular forces exercised between like and unlike polymer segments. Therefore, properties such as glass transition temperature, melting point, solubility, crystallinity, permeability, adhesion, elasticity, and chemical reactivity may be varied within wide limits [15]. Most existing procedures for calculating reactivity ratios can be classified as linear least-squares (LLS) and non-linear least-squares (NLLS) methods. It is accepted that LLS methods such as those proposed by [9], and by Kelen–Tudos [12,14] can only be applied to experimental data at sufficiently low conversion, because the calculation is based on the differential copolymerization equation [2,17]. The only LLS method, as an exception, is an extended Kelen–Tudos method [13,25], which involves a rather more complex calculation. In recent years some comprehensive work has been published on functional monomers and their polymers [1,10,18]. Functional polymers produced either by chemical modification of performed nonfunctional polymers, or by direct copolymerization of the desired functional monomers with suitably chosen structural and cross linking monomers [20,22,23]. The copolymerizability of vinyl monomers with a bulky substituent that also carries a highly electronegative atom like nitrogen and electronegative halogen like bromine is not discussed as yet. The possible effect of such a pendant group as the backbone with respect to reactivity is significant. The present paper deals mainly with the synthesis and characterization of copolymers of different compositions by free radical polymerization and reactivity ratio determined by Finemann–Ross method, Kelen–Tudos method, and a non-linear error-in-variable model (EVM) method using a computer program.

D. Thirumoolan et al. Benzoyl peroxide (BPO) was recrystallised from chloroform– methanol (1:1). p-bromoaniline (Hi-Media) was used without further purification, whereas AR grade benzene, hexane, chloroform, and methanol were distilled before being used. 2.2. Measurements FT-IR spectrum of the copolymer was recorded on Perkin-Elmer FT-IR spectrophotometer with KBr pellets. 1H NMR spectra and 13C NMR spectra were recorded in d6-DMSO with tetramethyl silane (TMS) as an internal reference on BRUKER 400 MHz spectrometer. Thermal stability of the polymers was determined using NETZSCH STA 409 C/CD thermal analyzer. The thermo gram was recorded with 5– 10 mg samples at a heating rate of 10 C per minute in nitrogen atmosphere. The molecular weights (Mn and Mw) of the homopolymer and copolymer were measured by employing gel permeation chromatography (GPC) (Waters 2690) using THF as eluent. Calibration was done by using polystyrene as a standard. 2.3. Copolymerization of p-bromophenylmethacylamide-co-Nvinyl-2-pyrrolidone (poly(PBPMA-Co-NVP)) Required quantities of the monomers N-vinyl-2-pyrrolidone and N-(p-bromophenyl)-2-methacrylamide, along with BPO were dissolved in 25 mL of Benzene placed in a standard reaction tube to obtain a homogeneous solution. The mixture was flushed with oxygen free dry nitrogen gas. The inlet and outlet of the reaction tube were closed by means of rubber tubing and pinch cork. The reaction vessel is then immersed in a thermostatic water bath maintained at 70 C. The copolymerization reaction was allowed to proceed for an appropriate duration. After the reaction, the reaction vessel was removed from the thermostat and cooled under the tap. Then the solution was poured in ice-cold excess hexane to precipitate the copolymer. The copolymers were purified by repeated precipitation by hexane from solution in chloroform. It was then dried in vacuum oven for 24 h. The Synthesis of copolymers (PBPMA-CoNVP) is shown in Scheme 1. White solid, IR [m, cm1, KBr]: 3318, 3088, 2993, 2878, 1615, 1502, 1479, 1735, 1693, 1392, 883, 771, 562. 1H NMR [400 MHz, d, ppm, d6-DMSO]: 9.43 (s, 1H, sec. amide), 7.97(dd, 2H, J = 8.6 Hz aromatic), 7.62 (dd, 2H, J = 7.9 Hz, aromatic), 1.99 (t, 2H, J = 6.6 Hz, –N–CH2–), 2.09 (m, 2H, –CH2–CH2–CH2–), 3.95 (t, 1H, J = 8.2 Hz, tertiary), 1.68 (s, 2H, bridging –CH2–), 1.62 (s, bridging –CH2–) and 1.30 ppm (s, 3H, a-CH3–). 13C NMR [400 MHz, d, ppm, d6-DMSO]: 172.33 (C‚O of PBPMA), 175.49 (C‚O of NVP), 138.82, 130.07, 121.88, 119.02 (PBPMA Ar–C), 23.63 (CH3–, a-methyl group of PBPMA unit), 44.22, 34.71 and 17.65 ppm (–CH2– NVP ring), 47.79 ppm (CH of NVP), 43.21 and 35.43 ppm (–CH2– bridging).

2. Experimental section

2.4. Solubility test

2.1. Materials

Solubility of the copolymers was tested in various polar and non polar solvents. About 5–10 mg of the copolymer was added to about 2 mL of different solvents in a test tube and kept overnight with the test tube tightly closed. The solubility of the copolymers was noted after 24 h.

N-(p-bromophenyl)-2-methacrylamide (PBPMA) monomer was synthesized as a reported procedure [21]. N-vinyl-2-pyrrolidone (Sigma–Aldrich) was freed from inhibitor by distillation.

Please cite this article in press as: D. Thirumoolan et al., Synthesis, characterization and reactivity ratios of poly N-(p-bromophenyl)-2methacrylamide-Co-N-vinyl-2-pyrrolidone, Journal of Saudi Chemical Society (2013), http://dx.doi.org/10.1016/j.jscs.2013.09.003

Synthesis, characterization and reactivity ratios of poly N-(p-bromophenyl)-2-methacrylamide-Co-N-vinyl-2-pyrrolidone

Scheme 1

Synthesis scheme of copolymer poly(PBPMA-Co-NVP).

Table 1 Molecular weights by GPC for the copolymers of poly(PBPMA-Co-NVP). Polymer

Mn

Mw

Mw/Mn

poly(PBPMA-Co-NVP) 20:80 poly(PBPMA-Co-NVP) 50:50 poly(PBPMA-Co-NVP) 80:20

2058 5636 5804

3591 7623 7233

1.74 1.35 1.24

3. Results and discussion 3.1. Characterization of poly(PBPMA-Co-NVP) The IR spectrum of poly(PBPMA-Co-NVP), shows the characteristic band of both monomer units. This confirmed the structure of polymers in all aspects. The sharp band at 3318 cm1 is the most characteristic due to the –NH band of N-(p-bromophenyl)-2-methacrylamide (PBPMA) unit. The peak at 3088 cm1 corresponds to the C–H stretching of the aromatic system. The symmetrical and asymmetrical C–H stretching due to the methyl and methylene groups are observed at 2993 and 2878 cm1. The peak at 1735 cm1 is attributed to the amide carbonyl stretching. The peak at 1693 cm1 is due the C‚O stretching of the NVP. The peak at 1392 cm1 corresponds to the C–N–C of imide group in the NVP unit. The C–H and C‚C out of plane bending vibrations of the aromatic nuclei are observed at 883, 771 and 562 cm1, respectively. The above IR data confirms the polymer formation. The 1H NMR Spectrum of poly(PBPMA-Co-NVP) indicates the significant changes in the chemical shifts of backbone CH protons of PBPMA and NVP linkages, when these groups form the backbone of the copolymer macromolecules. The resonance signal at 9.43 ppm corresponds to –NH– proton of the PBPMA unit. The aromatic ring protons show

Table 2

3

a signal between 7.62 and 7.97 ppm. The signals at 2.09 and 1.99 ppm are due to the presence of –CH2– protons of the pyrrolidone ring. The –CH2– group attached to the N atom in the pyrrolidone ring was merged with the solvent peak. The peak at 1.30 ppm shows a-methyl protons of the polymer. Finally the signals observed at 1.68 and 1.62 ppm corresponds to the backbone methylene groups which confirm the polymer formation. The proton-decoupled 13C NMR spectra of poly (PBPMA-Co-NVP) shows that the amide carbonyl carbon of PBPMA appeared at 172.33 ppm. The NVP carbonyl carbon appeared at 175.49 ppm. Thus the two signals clearly explain the two carbonyl carbon present in our polymer. The aromatic carbons of PBPMA unit in copolymer appeared at 138.82, 130.07, 121.88 and 119.02 ppm, respectively. The ring carbons of NVP unit in the copolymer observed at 44.22, 34.71 and 17.65 ppm, respectively. The signals at 43.21 and 35.43 ppm, are due to the presence of backbone methylene carbons. The a-methyl carbon atom of PBPMA unit appeared at 23.63 ppm. The above 13C NMR spectral data confirms the formation of the copolymer. 3.2. Solubility test The solubility of the newly prepared copolymers in various solvents was tested in room temperature. The polymers were easily soluble in various polar and non-polar solvents, namely toluene, benzene, chloroform and acetone. The solubility test clearly shows that there is a wide possibility for using different solvents for the copolymers to be used in coating applications. 3.3. Molecular weights of the copolymers of poly(PBPMA-CoNVP) The molecular weights of the copolymers were determined by gel permeation chromatography (GPC) with polystyrene and

Thermo gravimetric analysis of copolymers of poly(PBPMA-Co-NVP).

Copolymer composition m1

m2

0.20 0.50 0.80

0.80 0.50 0.20

IDT

171 252 224

Temp (C) of weight loss (%) of copolymers 10%

30%

50%

70%

90%

208 302 280

264 322 294

283 353 338

298 390 393

312 – –

Please cite this article in press as: D. Thirumoolan et al., Synthesis, characterization and reactivity ratios of poly N-(p-bromophenyl)-2methacrylamide-Co-N-vinyl-2-pyrrolidone, Journal of Saudi Chemical Society (2013), http://dx.doi.org/10.1016/j.jscs.2013.09.003

4

D. Thirumoolan et al. Table 3

Monomer compositions in feed and in copolymers.

Feed composition in mole fraction PBPMA(M1)

NVP(M2)

0.20 0.35 0.50 0.65 0.80

0.80 0.65 0.50 0.35 0.20

Figure 1

Conversion (%)

4.4 6.3 7.2 9.8 8.4

Copolymer composition in mole fraction PBPMA(m1)

NVP(m2)

0.1946 0.2886 0.4101 0.4564 0.6083

08054 0.7114 0.5899 0.5436 0.3917

Composition curves of (PBPMA-Co-NVP). Figure 2

tetrahydrofuran as the standard and solvent, respectively. The weight-average (Mw), number-average (Mn) molecular weights and the polydispersity indices (Mw/Mn) of poly (PBPMA-Co-NVP) are presented (Table 1). The polydispersity index of the polymers ranged between 1.24 and 1.74. 3.4. Thermogravimetric analysis of poly(PBPMA-Co-EMA) Thermogravimetric analysis results of copolymers are presented in Table 2. The data clearly indicates that all polymers undergo a single-step decomposition. Initially the initial decomposition temperature (IDT) increases, later it decreases with increase of PBPMA content in the copolymer. The initial decomposition temperatures (IDT) were determined from the TGA thermogram and were found to be in the range 171– 252 C. The first decomposition may be due to the rupture of weak linkages and volatilization of low molecular weight species. The decomposition of copolymers at high temperature may be due to the breakage of main chain accompanied by volatilization of the cleaved products.

Fineman–Ross (F–R) plot for (PBPMA-Co-NVP).

3.5. Copolymer composition and monomer reactivity ratios The monomer reactivity ratios for the copolymerization of PBPMA with NVP were determined from the monomer feed ratios and the copolymer compositions. The reactivity ratio is useful in understanding the copolymerization behavior of the comonomers. The classical approach for acquiring copolymer data was to isolate the copolymers from each of the five feed compositions at less than 10% conversions and analyze the copolymer compositions by elemental analysis. The monomer compositions in feed and in copolymers are tabulated in Table 3. The plot of mole fractions of PBPMA (M1) in the feed versus that in the copolymer (m1) is shown in Fig. 1, which indicates the formation of an azeotrope at a PBPMA: NVP = 35.5:64.5 mol% composition. The reactivity from the copolymer composition of PBPMA: NVP was determined by the application of Finemann–Ross (F–R) and Kelen–Tudos (K–T) method Figs. 2 and 3. The F–R and K–T parameters

Please cite this article in press as: D. Thirumoolan et al., Synthesis, characterization and reactivity ratios of poly N-(p-bromophenyl)-2methacrylamide-Co-N-vinyl-2-pyrrolidone, Journal of Saudi Chemical Society (2013), http://dx.doi.org/10.1016/j.jscs.2013.09.003

Synthesis, characterization and reactivity ratios of poly N-(p-bromophenyl)-2-methacrylamide-Co-N-vinyl-2-pyrrolidone

5

Table 5 Comparison of the monomer reactivity ratios of PBPMA with NVP various methods.

Figure 3

System

Methods

r1

r2

Poly(PBPMA-Co-NVP)

F–R K–T RREVM

0.21 0.23 0.19

0.79 0.77 0.92

Kelen–Todus (K–T) plot for (PBPMA-Co-NVP). Figure 4

for the copolymers are presented in Table 4. In poly(PBPMACo-NVP), the value of r1 is 0.23 and r2 is 0.77 (K–T method), similarly the value of r1 is 0.21 and r2 is 0.79 (F–R method). The r1 and r2 values together indicate that NVP is more reactive than PBPMA, hence the copolymers contain a higher proportion of NVP units. Moreover the product of r1 and r2 is less than one, which indicates that the system follows a random distribution of monomeric unit. Now-a-days a number of non-linear methods [19,5,6] have been proposed to obtain correct values of monomer reactivity ratios. Notable among them is the non-linear error in variable model (EVM) method using a computer program (RREVM), which gives accurate results. In this method, one starts from the r1 and r2 values obtained by the method of K–T, r1 and r2 values estimated were generated using errors of 1% for the monomer composition in feed and of 5% for the copolymer composition. The r1 and r2 values from methods such as F–R, K–T and RREVM are presented in Table 5. For the first time a computer program published [26] allows rapid data analysis of the non-linear calculations. It also permits the calculation of the validity of the reactivity ratios in a quantitative fashion. The computer program produces reactivity ratios for the monomers in the system with a 95% joint confidence limit determination. These methods allow us to

Table 4

RREVM plot for (PBPMA-Co-NVP).

take accurately into account all the sources of experimental error. The r1 and r2 values obtained by the RREVM method are 0.19 and 0.92, respectively. The 95% joint confidence region for the determined r1 and r2 values using RREVM is shown in Fig. 4. 4. Conclusions The new methacrylate based copolymer has been successfully synthesized by free radical polymerization. Characterizations of the copolymer were performed by FT-IR, 1H NMR, 13C NMR and GPC. From GPC, the polydispersity indices of the copolymer were calculated. It suggests that the copolymer has a strong tendency for chain termination by disproportionation. The TGA analysis results support the copolymer having good thermal stability. The copolymer composition was obtained by 1H NMR analysis of the copolymer. The reactivity ratios were determined by F–R, K–T methods as well as by a non-linear EVM method using a computer program RREVM. The r1 values of all these methods are lesser than 1 and r2 values are also lesser than 1. This indicates that the composition of the system follows a random distribution of monomeric units.

F–R and K–T parameters of PBPMA and NVP copolymer system.

S. No

F = M1/M2

f = m1/m2

G = (F(f1))/f

H = F2/f

g = G/(a + H)

e = H/(a + H)

1. 2. 3. 4. 5.

0.2500 0.5385 1.0000 1.8571 4.0000

0.2263 0.3912 0.5627 0.7352 1.3998

0.8542 0.8380 0.7771 0.6688 1.1424

0.2761 0.7412 1.7771 4.6909 11.4302

0.4174 0.3337 0.2190 0.1035 0.0865

0.1349 0.2951 0.5010 0.7260 0.8659

a=

p

0.2761 · 11.4302 = 1.77.

Please cite this article in press as: D. Thirumoolan et al., Synthesis, characterization and reactivity ratios of poly N-(p-bromophenyl)-2methacrylamide-Co-N-vinyl-2-pyrrolidone, Journal of Saudi Chemical Society (2013), http://dx.doi.org/10.1016/j.jscs.2013.09.003

6

D. Thirumoolan et al.

Acknowledgements Authors would like to acknowledge the financial support by the University Grants Commission (UGC) (F. No. 39-808/ 2010 (SR)), New Delhi, India and also thank the Management, C. Abdul Hakeem College, Melvisharam, Vellore, Tamil Nadu, India for the encouragement to carry out this work.

[14]

References

[16]

[1] A. Akelah, A. Moel, Functionalized Polymer and Their Applications, Thompson, New Delhi, 1990. [2] T. Alfery, G.J. Goldfinger, The mechanism of copolymerization, J. Chem. Phys. 12 (1944) 205–209. [3] M. Coskun, H. Erten, E. Ozdemir, M.A. Ahmedov, 3Cyclohexyloxy-2-hydroxypropyl acrylate-styrene copolymers: synthesis, characterization, and reactivity ratios, J. Macro. Sci. Pure Appl. Chem. A34 (1997) 91–98. [4] K. Demirelli, M. Coskun, I. Erol, Copolymerization and monomer reactivity ratios of 2-(3-mesityl-3-methylcyclobutyl)2-hydroxyethyl methacrylate with acrylonitrile, Eur. Polym. J. 36 (2000) 83–88. [5] K.F. Driscoll, P.M. Reilly, Determination of reactivity ratios in copolymerization, Macromol. Chem. Macromol. symp. 10–11 (1987) 355–374. [6] M. Dube, S.R. Amin, A. Penlidis, A microcomputer program for estimation of copolymerization reactivity ratios, J. Polym. Sci. Polym. Chem. 29 (1991) 703–708. [7] I. Erol, C. Soykan, H. Turkmen, Y. Tufan, Synthesis and characterization of novel methacrylates derived from morpholine and pyrrolidine: the determination of kinetic parameters with thermogravimetric analysis, J. Macro. Sci. Pure Appl. Chem. A40 (2003) 1213–1225. [8] I. Erol, C. Soykan, Free-radical-initiated copolymerization of 2(2-naphthylamino)-2-oxo-ethyl methacrylate with methyl methacrylate and styrene, Polym. Int. 53 (2004) 1235–1244. [9] M. Finemann, S.D. Ross, Linear method for determining monomer reactivity ratios in copolymerization, J. Polym. Sci. 5 (1950) 259–265. [10] G.G.A. Gordon, C.S.J. Selvamalar, A. Penlidis, S. Nanjundan, Homopolymer of 4-propanoylphenyl methacrylate and its copolymers with glycidyl methacrylate: synthesis, characterization, reactivity ratios and application as adhesives, React. Funct. Polym. 59 (2004) 197–209. [11] M.R. Huang, Y.B. Ding, X.G. Li, Lead-ion potentiometric sensor based on electrically conducting microparticles of sulfonic phenylenediamine copolymer, Analyst 138 (2013) 3820–3829. [12] T. Kelen, F. Tudos, B. Turcsanyi, Confidence intervals for copolymerization reactivity ratios determined by the Kelen– Tudos method, Polym. Bull. 2 (1980) 71–76. [13] T. Kelen, F. Tudos, B. Turcsanyi, J.P. Kennedy, Analysis of the linear methods for determining copolymerization reactivity ratios. IV. A comprehensive and critical reexamination of

[15]

[17]

[18]

[19]

[20]

[21]

[22]

[23]

[24]

[25]

[26]

carbocationic copolymerization data, J. Polym. Sci. Polym. Chem. 15 (1977) 3047–3074. J.P. Kennedy, T. Kelen, F. Tudos, Analysis of the linear methods for determining copolymerization reactivity ratios. II. A critical reexamination of cationic monomer reactivity ratios, J. Polym. Sci. Polym. Chem. 13 (1975) 2277–2289. S.W. Kuo, H.C. Kao, F.C. Chang, Thermal behavior and specific interaction in high glass transition temperature PMMA copolymer, Polymer 44 (2003) 6873–6882. X.G. Li, H. Feng, M.R. Huang, G.L. Gu, M.G. Moloney, Ultrasensitive Pb(II) potentiometric sensor based on copolyaniline nanoparticles in a plasticizer-free membrane with long life time, Anal. Chem. 84 (2012) 134–140. F.R. Mayo, F.M. Lewis, Copolymerization. I. A basis for comparing the behavior of monomers in copolymerization; the copolymerization of styrene and methyl methacrylate, J. Am. Chem. Soc. 66 (1944) 1594–1601. S. Nanjundan, C.S. Unnithan, C.S.J. Selvamalar, A. Penlidis, Homopolymer of 4-benzoylphenyl methacrylate and its copolymers with glycidyl methacrylate: synthesis, characterization, monomer reactivity ratios and application as adhesives, React. Funct. Polym. 62 (2005) 11–24. D. Radic, L. Gargallo, Synthesis, reactivity ratios, and solution behavior of vinylpyrrolidone-co-monoalkyl itaconate and vinylpyrrolidone-co-dialkyl itaconate copolymers, Macromolecules 30 (1997) 817–825. B.S.R. Reddy, R. Arshady, M.H. George, Copolymerization of N-vinyl-2-pyrrolidone with 2,4,5-trichlorophenyl acrylate and with 2-hydroxyethyl methacrylate: Reactivity ratios and molecular weights, Eur. Polym. J. 41 (1985) 511–515. S.M. Safiullah, D. Thirumoolan, K. Anver Basha, K.M. Govindaraju, D. Gopi, Tapan Kanai, A.B. Samui, Synthesis, characterization and corrosion protection properties of polyN(p-bromophenyl)-2-methacrylamide-co-glycidyl methacrylate on low nickel stainless steel, J. Polym. Eng. 31 (2011) 199–204. S. Soundararajan, B.S.R. Reddy, S. Rajadurai, Synthesis and characterization of glycidyl methacrylate-styrene copolymers and determination of monomer reactivity ratios, Polymer 31 (1990) 366–370. S. Soundararajan, B.S.R. Reddy, Glycidyl methacrylate and Nvinyl-2-pyrrolidone copolymers: Synthesis, characterization, and reactivity ratios, J. Appl. Polym. Sci. 43 (1991) 251–258. C. Soykan, S. Guven, R. Coskun, 2-[(5-methylisoxazol-3yl)amino]-2-oxo-ethyl methacrylate with glycidyl methacrylate copolymers: Synthesis, thermal properties, monomer reactivity ratios, and antimicrobial activity, J. Polym. Sci. Part A: Polym. Chem. 43 (2005) 2901–2911. F. Tudos, T. Kelen, B. Turcsanyi, J.P. Kennedy, Analysis of the linear methods for determining copolymerization reactivity ratios. VI. A comprehensive critical reexamination of oxonium ion copolymerizations, J. Polym. Sci. Polym. Chem. Ed. 19 (1981) 1119–1132. A.M. Van Herck, Least-squares fitting by visualization of the sum of squares space, J. Chem. Ed. 72 (1995) 138–140.

Please cite this article in press as: D. Thirumoolan et al., Synthesis, characterization and reactivity ratios of poly N-(p-bromophenyl)-2methacrylamide-Co-N-vinyl-2-pyrrolidone, Journal of Saudi Chemical Society (2013), http://dx.doi.org/10.1016/j.jscs.2013.09.003