You are on page 1of 9

Journal of King Saud University – Engineering Sciences (2018) 30, 22–30

King Saud University

Journal of King Saud University – Engineering Sciences


www.ksu.edu.sa
www.sciencedirect.com

ORIGINAL ARTICLE

Modelling and optimization of process variables for


the solution polymerization of styrene using
response surface methodology
Rasheed Uthman Owolabi *, Mohammed Awwalu Usman, Abiola John Kehinde

Department of Chemical Engineering, University of Lagos, Akoka, Yaba, Lagos 101017, Nigeria

Received 14 February 2015; accepted 25 December 2015


Available online 6 January 2016

KEYWORDS Abstract A satisfactory model for predicting monomer conversion in free radical polymerization
Free radical polymerization; has been a challenge due to the complexity and rigors associated with classical kinetic models. This
Polystyrene; renders the usage of such model an exciting endeavour in the academia but not exactly so in indus-
Acetone; trial practice. In this study, the individual and interactive effects of three processing conditions
Monomer conversion; (reaction temperature, reaction time and initiator concentration) on monomer conversion in the
Response surface solution polymerization of styrene using acetone as solvent was investigated in a batch reactor
methodology through the central composite design (CCD) model of response surface methodology (RSM) for
experimental design, modelling and process optimization. The modelled optimization conditions
are: reaction time of 30 min, reaction temperature of 120 °C, and initiator concentration of
0.1135 mol/l, with the corresponding monomer conversion of 76.82% as compared to the observed
conversion of 70.86%. A robust model for predicting monomer conversion that is very suitable for
routine industrial usage is thus obtained.
Ó 2016 The Authors. Production and hosting by Elsevier B.V. on behalf of King Saud University. This is
an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).

1. Introduction high molecular weight polymers (Krzysztof, 1998). Its essential


feature is that polymer chains are initiated, propagated, and
Free radical polymerization (FRP) is a chain growth polymer- terminated almost instantaneously. Modelling of polymeriza-
ization technique that is widely used in the industries due to its tion reaction is crucially vital in designing and optimizing reac-
versatility in monomer selection and relative insensitivity to tion conditions for tailoring product properties in addition to
impurity (Mastan et al., 2015), especially in the synthesis of enabling the prediction of conversion profile. Yong et al.
(2015) opined that ‘‘it is of technological importance to
* Corresponding author. develop an effective computational model for FRP, and
E-mail address: uthmanrash642@yahoo.com (R.U. Owolabi). thereby, aid in optimizing experimental conditions to achieve
Peer review under responsibility of King Saud University. the desired polymerization products”. Kinetic modelling of
free radical polymerization is characterized by intricacies, inac-
curacies and mathematical rigors which render its routine
usage, especially in practice, a serious challenge. The model
Production and hosting by Elsevier enables the prediction of such crucial parameters as monomer
http://dx.doi.org/10.1016/j.jksues.2015.12.005
1018-3639 Ó 2016 The Authors. Production and hosting by Elsevier B.V. on behalf of King Saud University.
This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).
Modelling and optimization of process variables 23

conversion, number average molecular weight (Mn), weight starch, reaction time, reaction temperature and initiator con-
average molecular weight (Mw) and polydispersity index centration to determine their individual and interactive effects
(PDI). The classical quasi-steady state approximation fails to on the grafting percentage. They obtained satisfactory results
provide accurate results especially at high conversion owing as the actual experimental yield at optimized conditions was
to an increased viscosity of the reaction mixtures in the face very close to the value predicted by their derived model.
of continuous occurrence of initiation throughout the poly- Aroonsingkarat and Hansupalak (2013) studied the effect of
merization. This is further compounded by an increased tem- processing conditions on monomer conversion in the graft
perature for this inherently exothermic reaction, especially in copolymerization of polystyrene and rubber using response
the absence of adequate heat removal. This dynamics of the surface methodology via CCD. The reaction temperature,
host solution engenders phenomena such as gel, glass and cage time, percentage of deproteinized rubber, and amount of chain
effect which complicate kinetic modelling. Several workers transfer agent were the four variables investigated. In a related
have attempted incorporating each of these occurrences in study, Sresungsuwan and Hansupalak (2013) investigated the
their models to varying degree of success (Achilias and influence of processing conditions on the mechanical proper-
Kiparissides, 1988, 1992; Venkateshwaran and Kumar, 1992; ties of compatibilized styrene/natural rubber blend using
Frounchi et al., 2002; Keramopoulos and Kiparissides, 2002, CCD. To the best of our knowledge, no such study has been
2003; Achilias, 2007; Verros and Achilias, 2009). However, reported in the literature for solution polymerization of
these efforts largely result in models with a limited range of styrene.
application. In our previous contributions, we established the various
Recently, Garg et al. (2014a–d) derived an analytical solu- factors influencing monomer conversion in the solution poly-
tion for free radical polymerization and validated same for merization of styrene (solvent polarity, nature and concentra-
various possible scenarios to establish its general applicability. tion of initiator, monomer concentration, reaction temperature
The effectiveness of the Garg model was further corroborated and time) and their limiting values (Kehinde et al., 2013;
by its good prediction of monomer conversion in the solution Owolabi et al., 2014, 2015). In the present study, we explore
polymerization of styrene using acetone as solvent (Owolabi the effect of initiator concentration, reaction time and reaction
et al., 2014). However, the model is still fraught with the inher- temperature on monomer conversion in the solution polymer-
ent complexity and mathematical rigors of other kinetic mod- ization of styrene with acetone as solvent. Response surface
els. Another recent effort by Yong et al. (2015) to model free methodology via CCD was used to design the experiment, gen-
radical polymerization using dissipative particle dynamics erate a model and optimize the process variables. The overall
hardly address the enumerated shortcomings. There is there- objective is to obtain a model equation for routine determina-
fore the need for a robust model that is valid for all range of tion of monomer conversion as a function of these processing
conversion and yet simple enough for routine industrial appli- conditions.
cation. Such model should be based on process parameters
rather than live and dead chains as obtained in the kinetic 2. Materials and methods
models. To this end, the response surface methodology proves
invaluable. 2.1. Materials
Response surface methodology (RSM) is a statistical exper-
imental design that enables simultaneous varying of process
The chemicals used are styrene (99%) inhibited by 10–15 ppm
variables, unlike what obtains in the conventional experimen-
4-tertbutylcatechol, benzoyl peroxide (75%), methanol
tation, thereby eliciting the interaction between such variables.
(CH3OH) (99.8%), sodium sulphate (Na2SO4) (99%), sodium
It is a faster and more economical method for gathering
hydroxide (NaOH) (98%), and acetone (99.9%). All the
research results than the classic one-variable at a time or
reagents were of analytical grade, purchased from Sigma
full-factor experimentation (Krishnaiah et al., 2015). It also
Aldrich in Germany and used as received except for styrene
provides a model equation relating the response parameter
monomer which was de-stabilized as previously described
to the process variables and optimization of the same. It is a
(Kehinde et al., 2013; Owolabi et al., 2014, 2015).
veritable tool that has been deployed in a wide range of fields
namely; transesterification (Betiku et al., 2015; Muppaneni
2.2. Polymerization
et al., 2013), solvent extraction (Rai et al., 2016;
Mohammadi et al., 2016), adsorption (Ahmed and Theydan,
2014; Ezechi et al., 2015), Fenton process (Kumar and Pal, The polymerization was conducted in a 62 mm diameter round
2012), drying operations (Krishnaiah et al., 2015), carrageenan bottom pressure double neck reaction flask equipped with a
production (Bono et al., 2014) etc. reflux system. The vapourized acetone was collected by a reflux
In polymer and related fields, RSM has found application condenser. In each run, specific amount of BPO and styrene
in some reported studies (Ghasemi et al. 2010; Lee et al., monomer concentration (8.612 M) were dissolved in a desired
2011; Nasef et al., 2011; Banerjee et al., 2012; Chieng et al., volume of acetone to maintain a monomer to solvent ratio of
2012; Zheng et al., 2015; Rojo et al., 2015; Razak et al., 1:1. The reactor was maintained at the specified temperature
2015; Fattahpour et al., 2015; Hirzin et al., 2015; and time as shown in Table 2. Thereafter, the reactor was
Davoudpour et al., 2015). Razali et al. (2015) used RSM to opened up, and cooled to collect the resulting polymer solu-
study the grafting of polydiallydimethylammonium chloride tion. The clear polymer solution was added to about 3 ml of
(PolyDADMAC) to cassava starch using potassium persul- methanol in a beaker with continuous stirring to precipitate
phate (KPS) as a free radical initiator. Four variables were the polymer. The top clear solvent was decanted while the bot-
investigated via central composite design (CCD) namely; mole tom polymer samples were air-dried to remove excess solvent
ratio of diallydimethylammonium chloride (DADMAC) to and dried for 2 weeks at room conditions until a constant
24 R.U. Owolabi et al.

weight was reached. Each run was conducted in triplicates and model that correlated the response to the three process vari-
the average value taken. Monomer conversion was determined ables using a second-degree polynomial as shown in Eq. (2).
by gravimetric method as described in a previous study
(Kehinde et al., 2013). X
n X
n n1 X
X n
Y ¼ bo þ bi Xi þ bii Xii þ bij Xi Xj ð2Þ
i¼1 i¼1 i¼1 j¼iþ1
2.3. Design of experiment
where Y is the predicted response, bo is a constant coefficient,
In this study, polystyrene samples were prepared via solution bi is a linear coefficient, bii is the quadratic equation, bij is an
polymerization of styrene using acetone as solvent and benzoyl interaction coefficient, and Xi and Xj are the coded values of
peroxide (BPO) as initiator. The variables studied were the the polymerization variables. Table 2 shows the run order,
reaction temperature (X1 ), initiator concentration (X2 ), and experimental design and the observed response (% monomer
reaction timeðX3 Þ: All variables and their respective range were conversion) for the three variables and 20 experimental runs
chosen based on our earlier studies (Kehinde et al., 2013; generated. Design Expert software (minitab 16.1) is used for
Owolabi et al., 2014, 2015) as shown in Table 1. A five-level- RSM regression analysis and optimization of monomer con-
three-factor central composite design (CCD) was employed, version data with input parameters. The statistical testing of
requiring 20 experimental runs (calculated based on Eq. (1)) the model, which includes linear, quadratic and interaction
which consist of 8 factorial runs, 6 axial runs and 6 replicate coefficient, is performed by ANOVA analysis with F-test to
runs at the centre. obtain the empirical correlation between input and output
parameters. To examine the goodness of fit of the model, each
N ¼ 2n þ 2n þ Nc ¼ 23 þ 2x3 þ 6 ¼ 20 ð1Þ
term of model is tested statistically which confirmed the signif-
where N is the total experimental runs and n is the number of icance of F-values with p 6 0.05. The values of R2, adjusted R2,
variables. and predicted R2, lack of fit and adequate precision of models
The a-value for this design was fixed at 1 (face-centred), are obtained to check the quality of the suggested polynomial.
and the response for this experiment was the percent monomer The response surface plot and contour plot are drawn to visu-
conversion ðYÞ. The response was used to develop an empirical alize the input–output relationships.

Table 1 Experimental variables and their coded levels for central composite design.
Variables Units Coded variables level
a 1 0 1 a
X1 °C 39.55 60 90 120 140.45
X2 mol/l 0.0304 0.0515 0.0825 0.1135 0.1346
X3 min 16.36 30 50 70 83.64

Table 2 Experimental design matrix for the preparation of the styrene polymerization.
Run Coded factor Actual factor % Conversion
X1 X2 X3 X1 (°C) X2 (mol/l) X3 (min) Observed response Predicted response
1 0 0 0 90 0.0825 50 47.46 48.23
2 0 0 0 90 0.0825 50 48.12 48.23
3 1 1 1 60 0.0516 30 33.55 37.53
4 0 0 1.6818 90 0.0825 83.6359 48.34 50.53
5 0 0 0 90 0.0825 50 48.79 48.23
6 1 1 1 60 0.1135 30 45.92 48.44
7 0 1.68179 0 90 0.1346 50 57.84 58.40
8 1 1 1 60 0.1135 70 44.37 45.16
9 1 1 1 120 0.1135 30 76.82 76.41
10 1.6818 0 0 39.546 0.0825 50 45.25 40.59
11 1 1 1 60 0.0516 70 34.88 36.02
12 1 1 1 120 0.0516 70 54.31 52.53
13 0 0 0 90 0.0825 50 49.00 48.23
14 1 1 1 120 0.0516 30 55.19 55.14
15 1 1 1 120 0.1135 70 75.28 72.04
16 0 0 0 90 0.0825 50 47.24 48.23
17 0 0 0 90 0.0825 50 48.57 48.23
18 0 0 1.6818 90 0.0825 16.3641 58.72 55.48
19 0 1.6818 0 90 0.0304 50 34.44 32.83
20 1.6818 0 0 140.454 0.0825 50 74.39 78.00
Modelling and optimization of process variables 25

the coefficient of determination R2, adjusted R2, and predicted


Table 3 Coefficient of the model.
R2. The quadratic models in terms of coded and actual value of
Factor Coefficient variables are shown in Eqs. (3) and (4), respectively.
Coded Uncoded (actual)
Y ¼ 48:2265 þ 11:1217X1 þ 7:6016X2  1:4716X3
Constant 48.2265 53.2398
þ 3:9144X21  0:9223X22 þ 1:6905X23  0:2750X1 X3
X1 11.1217 0.619582
X2 7.6016 188.92  0:4425X2 X3 þ 2:5925X1 X2 ð3Þ
X3 1.4716 0.395978
X21 3.9144 0.00434928 Y ¼ 53:2398  0:619582X1 þ 188:92X2  0:395978X3
X22 0.9223 962.788
X23 1.6905 0.00422626 þ 0:00434928X21  962:788X22 þ 0:00422626X23
X1 X3 0.2750 0.000458333 þ 2:7921X1 X2  0:000458333X1 X3  0:714863X2 X3 ð4Þ
X2 X3 0.4425 0.714863
X1 X2 2.5925 2.79214 The positive signs in the models signify synergetic effects of
R-square 96.81% factor while the negative sign indicates antagonistic effect.
Adjusted R-square 76.30% The analysis of variance (ANOVA) of the regression model
Predicted R-square 93.94% (Table 3) revealed R2 value of 0.9681, indicating that the model
can explain 96.81% of the data variation and only 3.19% of
the total variations were not explained by the model. For a
model to be adequate, R2 value should not be less than 0.75
2.4. Characterization (Le Man et al., 2010). However, Koocheki et al. (2009) posited
that a large value of R2 does not always imply that the regres-
The sample prepared with optimum processing conditions sion model is a good one and such inference can only be made
(optimized sample) was characterized to establish its true nat- based on a similarly high value of adj R2. The value of the
ure. The IR spectra of the sample dispersed in KBr discs were adjusted determination coefficient (Adj R2 = 0.9394) therefore
done using Perkin Erlmer spectroscopy. DSC/DTA scanning confirmed that the model was highly significant, which indi-
of the sample was carried out on a DSC Q200 machine. The cated a good agreement between the experimental and pre-
sample was pulverized for morphological analysis using elec- dicted values of monomer conversion. Thus the model is
tron microscopy. adequate for prediction in the range of experimental variables.
According to Rai et al. (2016), adj R2 and pred R2 should be
within 20% to be in good agreement. This requirement is sat-
3. Results and discussion
isfied in this study with a pred. R2 value of 0.7630. The model
therefore offers 76.3% variability in prediction monomer con-
3.1. Development of regression model equation
version beyond the experimental range of process conditions.
Table 4 shows the ANOVA of each term of the quadratic
The observed percent monomer conversions for the 20 exper- model. A term is significant if the F-value is large and
imental runs are presented in Table 2. These data are used to P < 0.05. From the table, the linear terms X1 and X2 are sig-
determine the coefficient of the polynomial equation as earlier nificant, only one quadratic term X21 and one interaction term
explained in Section 2.3. These estimated coefficients for both X1 X2 are significant. The other terms have no significant effect
the coded and actual values are shown in Table 3 along with on monomer conversion.

Table 4 Analysis of variance (ANOVA) for response surface quadratic model.


Source Sum of squares DF Mean square F-value T-value P-value
Regression model 2839.34 9 315.48 33.74 38.673 0.000
Linear 2507.97 3 835.99 89.42 13.442 0.000
X1 1689.25 1 1689.25 180.69 9.187 0.000
X2 789.15 1 789.15 84.41 1.779 0.000
X3 29.57 1 29.57 3.16 0.106
Square 275.42 3 91.81 9.82 4.860 0.003
X21 233.57 1 220.81 23.62 1.145 0.001
X22 12.26 1 12.26 1.31 2.099 0.279
X23 29.60 1 41.18 4.41 0.062
Interaction 55.94 3 18.65 1.99 0.254 0.179
X1 X3 0.60 1 0.60 0.06 0.409 0.802
X2 X3 1.57 1 1.57 0.17 2.398 0.691
X1 X2 53.77 1 53.77 5.75 0.037
Residual error 93.49 10 9.35
Lack of fit 90.89 5 18.18 34.95
Pure error 2.60 5 0.52
Total 2932.83 19
26 R.U. Owolabi et al.

3.2. Analysis of response surface

Three-dimensional response surfaces were plotted to investi-


gate the interaction among the variables and to determine
the optimum condition of each factor for maximum styrene
conversion. The effect of initiator concentration and reaction
temperature on styrene conversion at a constant reaction time
of 50 min is presented in Fig. 2a. As the reaction temperature
and initiator concentration increase, the styrene monomer con-
version increases. Maximum styrene conversion was obtained
at initiator concentration 0.1346 mol/l. The optimum styrene
conversion could be obtained at about 120 °C and at initiator
concentration 0.1346 mol/l. In overall, there is a net positive
Figure 1 Conversion of styrene monomer observed in the interactive effect between the two process variables. This is a
experiment versus predicted values by the model. strong indication of the dependence of the styrene monomer
conversion on both the temperature and initiator
concentration.
The interactive effect of initiator concentration and reac-
tion time on styrene conversion at a constant reaction temper-
ature of 90 °C is shown in Fig. 2b. The conversion was
observed to rapidly increase with an increase in initiator con-
centration compared to that of reaction time. There is a nega-
tive significant interaction between the initiator concentration
and reaction time. This shows that the styrene conversion
reduces with an increase in initiator concentration and reaction
time. There is probably an indication of lower solvent cage
effect below this temperature at particular reaction time which
allows the generated radicals to escape into the bulk medium
and grow.
The plot on Fig. 2c is similar to that of Fig 2b. The interac-
tive effect of reaction temperature and reaction time on
styrene conversion at a constant initiator concentration of
Figure 2a Response surface plot of the interactive effect of 0.0825 mol/l is shown in Fig. 2c. The conversion was observed
initiator concentration and reaction temperature on % styrene to rapidly increase with an increase in reaction temperature
conversion holding reaction time constant. compared to that of reaction time. There is a negative signifi-
cant interaction between the reaction temperature and reaction
time. This shows that the styrene conversion reduces with an
increase in reaction temperature and reaction time above the
Fig. 1 shows the plot of predicted conversion by the devel- optimum value of 0.0825 mol/l initiator concentration.
oped model against experimental values. The model success- Fig. 2d shows the contour plots of styrene conversion as a
fully captured the correlation between the process conditions function of reaction temperature and time. It can be seen that
and monomer conversion because the predicted values were an increased temperature and lower reaction time (between 20
very close to the observed values. and 30 min) would increase the styrene conversion. This may
be due to the formation of radical population at the early

Figure 2b Response surface plot of the interactive effect of Figure 2c Response surface plot of the interactive effect of
initiator concentration and reaction time on % styrene conversion reaction temperature and reaction time on % styrene conversion
holding reaction temperature constant. holding initiator concentration constant.
Modelling and optimization of process variables 27

period of the reaction. Specifically, a conversion of above


80%, between 70–80%, 60–70%, 50–60% and less than 50%
are obtainable at reaction temperatures of 140, 120, 100 and
72 °C, respectively at a constant initiator concentration of
0.0825 mol/l. The reaction temperature and reaction times
demonstrated a suppressive effect on each other.
Fig. 2e shows the contour plots of styrene conversion as the
function of initiator concentration and reaction time. From the
plots, operating at high initiator concentration and low reac-
tion time as experienced in earlier case is favorable to increase
conversion while holding temperature at 90 °C. A suppressive
effect or interplay was also observed by the two process vari-
ables (initiator concentration and reaction time). Least conver-
sion of less than 35% conversion is obtainable at operating
conditions 30–75 min reaction time, 0.04 mol/l initiator con-
Figure 2d Contour plot of % conversion vs reaction tempera-
centration and 90 °C.
ture, reaction time.
A synergetic effect of process variables (initiator concentra-
tion and reaction temperature) was observed in Fig. 2f. High
styrene conversion (greater than 90%) is obtained at higher
reaction time and higher initiator concentration while low styr-
ene conversion reverse happens (less than 35%) at reversed
operating conditions.

3.3. Optimization and validation

In order to obtain the maximum response that jointly satisfies


all process conditions, optimization was carried out using the
RSM software. The optimum conditions obtained from this
study are as follows: reaction time of 30 min, reaction temper-
ature of 120 °C and initiator concentration of 0.1135 mol/l.
The corresponding optimized monomer conversion is
76.82%. Validation experimental runs were conducted using
the optimum conditions in duplicate and the average value
of conversion obtained is 70.86%. In comparison with the pre-
dicted value there is an error of about 8.41%. There is there-
fore a good agreement between the experimental value and
Figure 2e Contour plot of % conversion vs initiator concentra- the predicted value based on the model. It is also pertinent
tion, reaction time. to mention that the optimized conversion obtained in this
study is remarkably comparable with a conversion of 70% at
a similar temperature of 120 °C reported to be frequent in
industrial sites for the production of general purpose polystyr-
ene resins (Mermier et al., 2015).

3.4. Characterization of the polystyrene

The polystyrene sample prepared with optimum conditions


was subjected to various analyses and the findings are pre-
sented and discussed in this section.

3.4.1. Spectral analysis


The IR spectra of the optimized sample and another sample
prepared by bulk polymerization indicate that the aromatic
ring and alkene that characterize polystyrene dominate the
spectrogram (1448.74–1695.18 cm1). The peaks at 1599.01
and 1448.74 cm1 are assigned to C‚C stretching of phenyl
group. The CAH deformation vibration band of benzene ring
hydrogen was seen at 751.64 cm1 and ring deformation was
observed at 693.61 cm1 (Naghash et al., 2007; Kaniappan
and Latha, 2011). This confirmed that the PS retained the
Figure 2f Contour plot of % conversion vs initiator concentra- alkene or benzene rings of the styrene from which it was
tion, reaction temperature. formed. Both the reference and Sample spectrograms are
28 R.U. Owolabi et al.

similar polymerization though under different control and


Table 5 Polystyrene major peaks.
activation environments. Differential scanning calorimetry
S/N Wave numbers in cm1 Wave numbers in cm1 (DSC)/Digital thermal analyser (DTA) performed on the
(Nicholas, 2012) (this study) optimized PS samples detects no exothermic peak. The absence
1 538.8 537.76 of the peak is an indication of the formation of atactic
2 623.3 693.61 polystyrene (aPS) in which the phenyl groups are randomly
3 694.1 – distributed on both sides of the polymer chain (Maria et al.,
4 752.1 751.64 2008). This random positioning prevents the chains from
5 905.9 906.84 aligning with sufficient regularity to achieve any crystallinity.
6 966.4 –
In a similar vein, Brun et al. (2011) carried out DSC/DTA
7 1025.7 1024.50
8 1068.4 1068.35
analysis on both aPS and syndiospecific polystyrene (sPS)
9 – 1270.03 and obtained no peak for aPS.
10 1367.6 1367.42
11 1448.8 1448.74 4. Conclusion
12 1491.0 1492.98
13 1599.3 1599.01
The individual and combined effects of three processing condi-
14 – 1695
15 2849.9 – tions (reaction time, reaction temperature, and initiator con-
16 2920.1 2849.12 centration) on monomer conversion, in the radical solution
17 – 2919.93 polymerization of styrene, were studied using CCD model of
18 3024.9 3025.14 RSM. The optimum factors were determined as reaction time
19 3059.5 3058.63 of 30 min, reaction temperature of 120 °C, and initiator con-
20 – 3743.77 centration of 0.1135 mol/l. Under these conditions, the opti-
mized (maximum) monomer conversion is 76.82%. The
obtained quadratic regression model is very adequate based
on ANOVA test. A robust model for predicting monomer con-
almost identical in peak position and intensity (Table 5)
version in styrene polymerization that is suitable routine indus-
excluding the peak above 3500 cm1 in the sample spectra
trial application has been developed. Characterization test (IR,
probably due to the presence of minor foreign materials.
DSC/DTA, molecular weight) confirmed that the optimized
sample possesses the necessary properties of polystyrene.
3.4.2. Molecular weight
The molecular weight of the optimized sample was determined Acknowledgement
through viscosity measurement as previously reported
(Kehinde et al., 2013). The value obtained was The authors acknowledge the support of University of Lagos
5.65738  105, when compared with the recommended values Mini-Research Grant (Ref. No.: M2014/02) for the completion
(5.5–20.5  105) reported by Wagner (1985) and Goldberg of this work.
et al. (2003), it is within acceptable range for easy processabil-
ity. The optimized sample therefore has useful commercial
References
applications and the ability to bear sufficient design loads.
Achilias, D.S., 2007. A review of modeling of diffusion controlled
3.4.3. Thermal properties polymerization reactions. Macromol. Theory Simul. 16 (4), 319–
Bindu et al. (2001) linked the melting point of materials to its 347.
thermal property. Park et al. (2001) similarly related melting Achilias, D., Kiparissides, C., 1988. Modeling of diffusion-controlled
point of polymers to their molecular weights, degree of cross free-radical polymerization reactions. J. Appl. Polym. Sci. 35 (5),
linking and polymer rigidity. Materials with sharp melting 1303–1323.
Achilias, D.S., Kiparissides, C., 1992. Development of a general
points exhibit sound thermal, morphological and molecular
mathematical framework for modeling diffusion-controlled free
properties. The melting point of optimized sample in this study radical polymerization reactions. Macromolecules 25 (14), 3739–3750.
was determined to be in the range of 215–220 °C. This is Ahmed, M.J., Theydan, S.K., 2014. Optimization of microwave
slightly different from values reported for polystyrene in liter- preparation conditions for activated carbon from Albizia lebbeck
ature such as 240 °C (Maria et al., 2008), 262.9–270.1 °C (Chen seed pods for methylene blue dye adsorption. J. Anal. Appl. Pyrol.
et al., 2009) and 275 °C (Brun et al., 2011). As observed in the 105, 199–208.
spectral analysis, the reduced melting point in this study may Aroonsingkarat, K., Hansupalak, N., 2013. Prediction of styrene
be as a result of foreign bodies likely to be present in the PS conversion of polystyrene/natural rubber graft copolymerization
(as evidenced by Peak 3500 cm1). However, the wide dis- using reaction conditions: Central composite design versus artificial
agreement in the melting points of PS obtained by Maria neural network. J. Appl. Polym. Sci. 128 (4), 2283–2290.
Banerjee, S., Joshi, M., Ghosh, A.K., 2012. Optimization of propy-
et al. (2008) with others is due to the nature of PS formed in
lene/clay nanocomposite processing using Box-Behnken statistical
each case. The later synthesized ordered syndiotactic polystyr- design. J. Appl. Polym. Sci. 123 (4), 2042–2051.
ene (through Ziegler–Natta polymerization) with the phenyl Betiku, E., Okunsolawo, S.S., Ajala, S.O., Odedele, O.S., 2015.
groups positioned on alternating sides of the hydrocarbon Performance evaluation of artificial neural network coupled with
backbone. This form is highly crystalline with a Tm of genetic algorithm and response surface methodology in modeling
270 °C (518 °F). Such PS is not commercially produced and optimization of biodiesel production from shea tree (vitellaria
because the polymerization is slow. The former performed paradoxa) nut butter. Renewable Energy 76, 408–417.
Modelling and optimization of process variables 29

Bindu, R.L., Nair, R.C.P., Ninan, K.N., 2001. Phenolic resins with Keramopoulos, A., Kiparissides, C., 2002. Development of a compre-
phenyl maleimide functions: thermal characteristics and laminate hensive model for diffusion-controlled free-radical polymerization
composite properties. J. Appl. Polym. Sci. 80 (10), 1664–1674. reactions. Macromolecules 35 (10), 4155–4166.
Bono, A., Anisuzzaman, S.M., Ding, O.W., 2014. Effect of process Keramopoulos, A., Kiparissides, C., 2003. Mathematical modeling of
conditions on the gel viscosity and gel strength of semi-refined diffusion-controlled free-radical terpolymerization reactions. J.
carrageenan (SRC) produced from seaweed (Kappaphycus alvar- Appl. Polym. Sci. 88 (1), 161–176.
ezii). J. King Saud Univ. Eng. Sci. 26, 3–9. Koocheki, A., Taherian, A.R., Razavi, S., Bostan, A., 2009. Response
Brun, N., Bourson, P., Margueron, S., Duc, M., 2011. Study of the surface methodology for optimization of extraction yield, viscosity,
thermal behavior of syndiotactic and atactic polystyrene by Raman hue and emulsion stability of mucilage extracted from Lepidium
spectroscopy. In: Journées d’Études des Équilibres entre Phases, perfoliatum seeds. Food Hydrocolloids 23 (8), 2369–2379.
EDP Sciences 0004. Krishnaiah, D., Bono, A., Sarbatly, R., Nithyanandam, R., Anisuz-
Chen, J., Biao, C., Ying-Mei, S., Chang-Wen, P., Zhao-Hua, J., 2009. zaman, S.M., 2015. Optimisation of spray drying operating
Syndiospecific polymerization of styrene with ar[o, o/nh]cp*ticl, conditions of Morinda citrifolia L. fruit extract using response
activated with MMAO. Chin. J. Polym. Sci. 27 (5), 659–665. surface methodology. J. King Saud Univ. Eng. Sci. 27, 26–36.
Chieng, B.W., Ibrahim, N.A., Wan Yunus, W.M.Z., 2012. Optimiza- Krzysztof, M., 1998. Controlled radical polymerization. In: ACS
tion of tensile strength of poly(lactic acid)/grapheme nanocompos- Symposium Series, 685. American Chemical Society (500).
ites using response surface methodology. Polym. Plast. Technol. Kumar, R., Pal, P, 2012. Response-surface optimized Fenton’s pre-
Eng. 51 (8), 791–799. treatment for chemical precipitation of struvite and recycling of
Davoudpour, Y., Hossain, S., AbdulKhalil, H.P.S., Mohammed water through downstream nanofiltration. Chem. Eng. J. 210, 33–44.
Haafiz, M.K., Mohd Ishak, Z.A., Hassan, A., Sarker, Z.I., 2015. Le Man, H., Behera, S.K., Park, H.S., 2010. Optimization of
Optimization of high pressure homogenization for the isolation of operational parameters for ethanol production from Korean food
cellulosic nanofibers using response surface methodology. Ind. waste leachate. Int. J. Environ. Sci. Technol. 7 (1), 157–164.
Crops Prod. 74, 381–387. Lee, K.S., Jeon, B., Cha, S.W., Jeong, K.Y., Han, I.S., Lee, Y.S., Lee,
Ezechi, E.H., Kutty, S.R.M., Malakahmad, A., Isa, M.H., 2015. K., Cho, S.M., 2011. A study on optimizing the mechanical
Characterization and optimization of effluent dye removal using a properties of glass fiber-reinforced polypropylene for automotive
new low cost adsorbent: equilibrium, kinetics and thermodynamic parts. Plast. Technol. Eng. 50 (1), 95–101.
study. Process Saf. Environ. Prot. 98, 16–32. Maria, C.A., Kuhn, J.L., Adriana, C.A.C., Osvaldo, L.C., 2008.
Fattahpour, S., Shamanian, M., Tavakoli, N., Fathi, M., Sheykhi, S. Styrene polymerization by nickel and titanium catalysts based on
R., Fattahpour, S., 2015. Design and optimization of alginate– tris(pyrazolyl)borate ligands. J. Braz. Chem. Soc. 19 (8), 1560–
chitosan–pluronic nanoparticles as a novel meloxicam drug deliv- 1566.
ery system. J. Appl. Polym. Sci. 132 (28). http://dx.doi.org/10.1002/ Mastan, E., Li, X., Zhu, S., 2015. Modeling and theoretical develop-
app. 42241. ment in controlled radical polymerization. Prog. Polym. Sci. 45,
Frounchi, M., Farhadi, F., Mohammadi, R.P., 2002. Simulation of 71–101.
styrene radical polymerization in batch reactor – a modified kinetic Mermier, N.R.J.D., Castor Jr, C.A., Pinto, J.C., 2015. Solution styrene
model for high conversion. Sci. Iran 9 (1), 86–92. polymerizations performed with multifunctional initiators. J. Appl.
Garg, D.K., Serra, C.A., Hoarau, Y., Parida, D., Bouquey, M., Polym. Sci. 132 (39). http://dx.doi.org/10.1002/app 42609.
Muller, R., 2014a. Analytical solution of free radical polymeriza- Mohammadi, R., Mohammadifar, M.A., Mortazavian, A.M., Rouhi,
tion: derivation and validation. Macromolecules 47, 4567–4586. M., Ghasemi, J.B., Delshadian, Z, 2016. Extraction optimization of
Garg, D.K., Serra, C.A., Hoarau, Y., Parida, D., Bouquey, M., pepsin-soluble collagen from eggshell membrane by response
Muller, R., 2014b. Analytical solution of free radical polymeriza- surface methodology (RSM). Food Chem. 190, 186–193.
tion: applications implementing gel effect using AK model. Muppaneni, T., Reddy, H.K., Ponnusamy, S., Patil, P.D., Sun, Y.,
Macromolecules 47, 7370–7377. Dailey, P., Deng, S., 2013. Optimization of biodiesel production
Garg, D.K., Serra, C.A., Hoarau, Y., Parida, D., Bouquey, M., from palm oil under supercritical ethanol conditions using hexane
Muller, R., 2014c. Analytical solution of free radical polymeriza- co-solvent: a response surface methodology approach. Fuel 107,
tion: applications implementing gel effect using CCS model. 633–640.
Macromolecules 47, 8178–8189. Naghash, J.H., Karimzadeh, A., Momeni, R.A., Massah, R.A., Alian,
Garg, D.K., Serra, C.A., Hoarau, Y., Parida, D., Bouquey, M., H., 2007. Preparation and properties of triethoxyvinylsilane-mod-
Muller, R., 2014d. Analytical solution of free radical polymeriza- ified styrene- butyl acrylate emulsion coploymers. Turk. J. Chem.
tion: applications implementing nonisothermal effect. Macro- 31, 257–269.
molecules 47, 8514–8523. Nasef, M.M., Aly, A.A., Saidi, H., Ahmad, A., 2011. Optimization of
Ghasemi, I., Karrabi, M., Mohammadi, M., Aziz, H., 2010. Evalu- reaction parameters of radiation induced grafting of 1-vinylimida-
ating the effect of processing conditions and organoclay content on zole onto poly(ethylene-co-tetrafluoroethene) using response sur-
the properties of styrene-butadiene rubber/organoclay nanocom- face method. Radiat. Phys. Chem. 80 (11), 1222–1227.
posites by response surface methodology. eXPRESS Polym. Lett. 4 Nicholas, O., 2012. Determining the identity of an unknown polymer
(2), 62–70. using infra-red spectroscopy. Chemistry 113 Laboratory Note-
Goldberg, A.I., Hohenstein, W.P., Mark, H., 2003. Intrinsic viscosity – book, 21–24, retrieved from www.psu.edu on 3 February 2015.
molecular weight relationship for polystyrene. J. Polym. Sci. 5, Owolabi, R.U., Usman, M.A., Kehinde, A.J., 2014. Styrene conver-
503–509. sion modeling and estimation of polydispersity index. NUST J.
Hirzin, R.S.F.N., Azzahari, A.D., Yahya, R., Hassan, A., 2015. Eng. Sci. 7 (1), 22–33.
Optimizing the usability of unwanted latex yield by in situ Owolabi, R.U., Usman, M.A., Kehinde, A.J., 2015. Rationalization of
depolymerization and fictionalization. Ind. Crops Prod. 74, 773– solvent effects in the solution polymerization of styrene. Andolu
783. University J. Sci. Technol. (in press)
Kaniappan, K., Latha, S., 2011. Certain Investigations on the Park, H., Yang, I., Wu, J., Kim, M., Hahm, H.K.S., Hee, H., 2001.
formulation and characterization of polystyrene/poly(methyl- Synthesis of siliconacrylic resins and their applications to super
methacrylate) blends. Int. J. Chem. Technol. Res. 3 (2), 708–717. weatherable coatings. J. Appl. Polym. Sci. 81 (7), 1614–1623.
Kehinde, A.J., Usman, M.A., Owolabi, R.U., 2013. Solvent–initiator Rai, A., Mohanty, B., Bhargava, R., 2016. Supercritical extraction of
compatibility and sensitivity of conversion of styrene homo- sunflower oil: a central composite design for extraction variables.
polymerization. J. Polym. Eng. 33 (9), 775–783. Food Chem. 192, 647–659.
30 R.U. Owolabi et al.

Razak, J.A., Ahmad, S.H., Ratnam, C.T., Mahamood, M.A., Venkateshwaran, G., Kumar, A., 1992. Solution of free-radical
Yaakub, J., Mohammad, N., 2015. Effects of EPDM-g-MAH polymerization. J. Appl. Polym. Sci. 45 (2), 187–215.
compatibilizer and internal mixer processing parameters on the Verros, G.D., Achilias, D.S., 2009. Modeling gel effect in branched
properties of NR/EPDM blends: an analysis using response surface polymer systems: free radical solution homopolymerization of vinyl
methodology. J. Appl. Polym. Sci. 132 (27). http://dx.doi.org/ acetate. J. Appl. Polym. Sci. 111 (5), 2171–2185.
10.1002/app 41299. Wagner, H.L., 1985. The Mark-Houwink-Sakurada equation for the
Razali, M.A.A., Ismail, H., Ariffin, A., 2015. Graft copolymerization viscosity of linear polyethylene. J. Phys. Chem. 14 (2), 611–617
of polyDADMAC to cassava starch: evaluation of process variables (Ref. Data).
via central composite design. Ind. Crops Prod. 65, 535–545. Yong, X., Kuksenok, O., Balazs, A.C., 2015. Modeling free radical
Rojo, E., Alonso, M.K., Saz-Orozco, B.D., Oliet, M., Rodriguez, F., polymerization using dissipative dynamics. Polymer. http://dx.doi.
2015. Optimization of the silane treatment of cellulosic fibers from org/10.1016/j.polymer.2015.01.052.
eucalyptus wood using response surface methodology. J. Appl. Zheng, Y., Tang, Q., Wang, T., Wang, J., 2015. Molecular size
Polym. Sci. 132 (26). http://dx.doi.org/10.1002/app 42157. distribution in synthesis of polyoxymethylene dimethyl ethers and
Sresungsuwan, N., Hansupalak, N., 2013. Prediction of mechanical process optimization using response surface methodology. Chem.
properties of compatibilized styrene/natural-rubber blend by using Eng. J. 278, 183–189.
reaction conditions: central composite design versus artificial
neural networks. J. Appl. Polym. Sci. 127 (1), 356–365.

You might also like