You are on page 1of 38

Accepted Manuscript

Title: Optimization of biodiesel production from castor oil by


Taguchi design

Authors: Bisheswar Karmakar, Sumit H. Dhawane, Gopinath


Halder

PII: S2213-3437(18)30195-7
DOI: https://doi.org/10.1016/j.jece.2018.04.019
Reference: JECE 2315

To appear in:

Received date: 12-2-2018


Revised date: 28-3-2018
Accepted date: 7-4-2018

Please cite this article as: Bisheswar Karmakar, Sumit H.Dhawane, Gopinath Halder,
Optimization of biodiesel production from castor oil by Taguchi design, Journal of
Environmental Chemical Engineering https://doi.org/10.1016/j.jece.2018.04.019

This is a PDF file of an unedited manuscript that has been accepted for publication.
As a service to our customers we are providing this early version of the manuscript.
The manuscript will undergo copyediting, typesetting, and review of the resulting proof
before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that
apply to the journal pertain.
Optimization of biodiesel production from castor oil by Taguchi design

Bisheswar Karmakara, Sumit H. Dhawanea, Gopinath Haldera*


a
Department of Chemical Engineering, National Institute of Technology Durgapur, India

T
Corresponding Author:

IP
*Dr. Gopinath Halder

R
Associate Professor,

SC
Department of Chemical Engineering,

National Institute of Technology Durgapur,

M.G. Avenue, Durgapur- 713209 U


N
West Bengal, India.
A
Telephone: +91-94347-88189
M

E-mail address: gopinathhaldar@gmail.com


ED

Highlights
PT

 The inedible castor oil was used as a feedstock for synthesis of RCME.

 The L16 Taguchi approach was adopted to optimize the esterification process.
E

 Methanol to oil ratio emerged as a most influential parameter affecting FFA conversion.
CC

 Esterification process follows pseudo-first order kinetics and requires low activation

energy.
A

 Fuel characterization results revealed the acceptability of the RCME.

1
Abstract

Development of a renewable substitute for petroleum fuels is needed to address the current
scenario of the rising fuel crisis affecting transportation, growing concerns regarding
environmental pollution as well as exhaustion of natural oil reserves. Keeping this in mind,
an experimental investigation was undertaken to synthesize biodiesel from inedible
commercial grade castor (Ricinus communis) oil. The feedstock was subjected to
esterification under the experimental conditions suggested by L16 Taguchi orthogonal

T
approach. Compared to heterogeneous sulfonated carbon catalysis, maximum FFA

IP
conversion of 90.83% was obtained with homogeneous acid catalyst under the following
optimized conditions: reaction temperature of 50 °C, reaction duration of 1 hour, catalyst

R
concentration of 1% w/w, methanol to oil ratio of 20:1 and agitation speed of 700 rpm.

SC
Identification of noteworthy parameters were carried out employing Taguchi approach and
emerged significant parameters were: molar ratio of oil to methanol, agitation speed, reaction
temperature and catalyst concentration with contribution factors of 59.6%, 16.95%, 12.59%
U
and 10.82% respectively. Statistical analysis using ANOVA (analysis of variance) showed
N
that the obtained results are in reasonable agreement with the predicted values. Kinetic
A
studies showed that the esterification follows pseudo-first order kinetics with reaction rate
constant of 0.0392 min-1 and activation energy of 8.184 kJ/mol. Fuel characterization results
M

confirmed that the RCME (Ricinus communis methyl ester) properties are comparable to
petro-diesel and within ASTM limits. Hence, the optimization employing L16 Taguchi
ED

approach yielded the best parametric conditions for the cost-effective synthesis of biodiesel
from castor oil through single step acid esterification.
PT

Keywords: Ricinus communis; Biodiesel; Acid esterification; Optimization study; Taguchi


E

approach; Kinetics.
CC

1. Introduction

The increasing damage to the environment from the combustion of petro-fuels, their
A

fluctuating costs due to the energy crisis and safety hazards in transportation make renewable,
biofuels a necessary and fitting substitute for the present and the predicted future scenario of
the global economy [1-2]. As a result, the procuring of alternative energy sources as a
replacement to fossil fuels is receiving an increasing interest since the past few decades.
Biodiesel has the following advantages compared to petro-diesel: (i) Being a type of

2
renewable energy it can be prepared from commonly available biological sources such as
vegetable oils, which include both edible and non-edible oils. This eliminates the need for
importing fuels, thereby increasing a country’s self-sufficiency on the fuel market. (ii) The
low sulphur content and lower aromatic content in the biodiesel contributes to the fuel’s low
greenhouse gas emission profile as well as having high flash point and fire point which
makes it safer for transportation. (iii) Biodiesel is also usable in its entirety or as a blend with
petro-diesel in modified or unmodified engines [3-6].

T
One of the well-known applications of esterification is the production of fatty acid methyl

IP
and ethyl esters (FAME and FAEE) using oils and fats. These have been thoroughly
investigated globally as a potential replacement for conventional diesel. Biodiesel refers to

R
fatty acid alkyl esters (FAAE) that can be derived by reacting the free fatty acids (FFA) or

SC
triglycerides present in the oils with alcohol in the presence or absence of an acid or base
catalyst. The properties of the biodiesel produced depend strongly on the type and
concentration of free fatty acids present, which vary depending on the feedstock used.
U
Vegetable oil feedstock includes edible oils from soybean, sunflower, rapeseed, palm,
N
linseed, etc. [7-8]. However, the use of edible oils for biodiesel production is undesirable
A
since they would lead to a sharp decline in the availability of oils meant for human
consumption. Hence, the use of non-edible oils is the next best idea. Yet, these crops face
M

strong criticism, since they can compete with food crops, as well as extensive land use. Also,
the major holdup in the production of biodiesel is the cost of feedstock, which accounts for
ED

75% to 85% of the price of biodiesel [9]. An ideal non-edible oilseed plant should not
compete with food crops, have high regional climatic adaptability, availability, high oil
PT

content, low pastoral needs (fertilizers, water, pesticides), and the capacity to grow on
undesirable land [10]. Some non-edible oilseed plants are: Jatropha curcas [11-12],
E

Pongamia pinnata, Ricinus communis (castor) [13-14], Hevea brasiliensis (rubber seed) [15-
CC

17]. In this study, we focussed on using castor oil as feedstock for biodiesel production.
Castor plant establishes itself easily in areas with a suitable climate and can also grow on
wastelands due to its relative resistance to hydric stress [18]. It is usually grown as a
A

decorative plant in public areas, thereby reducing the need for cultivation space [19]. Castor
oil on a global scale being grown in 43 countries is cultivated in 30 countries on a
commercial scale. From FAO (2012) statistical reports, the average global production of
castor seeds stood at 1.9 million metric tonnes on a yearly basis, cultivated on roughly 1.6
million hectares (ha), the valued average yield being at 1161.3 kg/ha. Due to climatic

3
favourability, the top 3 castor producing countries are India, China and Brazil, which account
for around 90% of the global production and meet the major part of the demands on the
global market [49]. Additionally, Figure 1 represents the global production of castor oil seed
(by country) from a period of 2013-2016, according to FAO statistical reports [50].

The castor beans are the source for castor oil. Zhang et al. [9] conclusively reported in their
study that only feedstock with FFA content below 2% w/w can be used for alkali-catalyzed
transesterification of triglycerides for obtaining FAAE (fatty acid alkyl esters). Above this

T
threshold, transesterification of oils with high amounts of FFA (free fatty acids) leads to

IP
saponification which is an undesirable reaction leading to the formation of unusable by-
products. Therefore, for the oils with FFA greater than 2% w/w, acid-catalyzed esterification

R
of the FFA present is a good choice [20].

SC
In simple terms, the main difference between optimizing a single factor at a time and
experimental designing is that in the first case, the next experiment to be conducted was

U
based on the results obtained from the previous experiments, while the second technique
clearly covers the entire set of experiments in a grid-like design, which covers the entire
N
experimental range. The optimization of experimental conditions while altering one variable
A
at one time (OVAT) is disadvantageous because it requires maximum number of
M

experimental runs to identify the significant parameters. Also, design of experiments (DOE)
for optimization has the advantage of reducing the number of experimental runs, the outcome
ED

of which is lower usage of chemicals and feedstock [41-43]. One technique to design an
experiment using suitable software is Response surface methodology (RSM) comprises a
system of analysis and elucidation of the relationship between the result (fed as a response)
PT

and independent parameters. Types of experimental designs include Central Composite


Design (CCD), Plackett – Burman Design, and Box–Behnken Doehlert (BBD) [44-46].
E

In the esterification process, FFA present in the oil reacts with methanol to give FAME (fatty
CC

acid methyl esters), in the presence of an acid catalyst which forms a homogenous mixture.
Agra et al. [21] performed ethanolysis of castor oil using H2SO4 (98%), obtaining 60%
A

conversion of the FFA present in the 1st step and 82% conversion in the 2nd step respectively,
for a runtime of 2 hours. S. Halder et. al. [13] carried out the esterification of castor oil with
methanol in the presence of H2SO4 (98%) and obtained an FFA% reduction of 73.27% under
optimum conditions via RSM 24 full factorial CCD (central composite design). To the best of
our knowledge, there are no studies reporting the significant parameters affecting FFA
conversion of castor oil and their optimization using Taguchi approach. Therefore, an

4
identification of the best parametric conditions for RCME production using orthogonal arrays
is required. Former studies reported that RCME meets the quality standards essential for
commercialization [23-24].

This study is aimed at optimizing the process parameters of the acid catalysed esterification
reaction for obtaining maximum FFA conversion and identifying the significance of each
parameter that affects the yield. Parameters such as reaction temperature, duration of
reaction, concentration of acid catalyst, molar ratio of oil to methanol and agitation speed

T
were considered for their impacts on the process. Taguchi method was utilized for designing

IP
an experimental matrix with a far less number of runs compared to other optimization
techniques. A regression model equation was developed to check the relevance of the

R
parameters; they were statistically analyzed to define relationships using analysis of variance

SC
(ANOVA) [13]. Under the established optimized conditions giving maximum FFA
conversion, a study was conducted with a view to establishing the kinetics of the acid
esterification. The fuel obtained was analysed using gas chromatography (GC) to check for
U
various methyl ester percentage compositions and characterized following ASTM test
N
procedures in order to determine its physico-chemical properties and compared with the
A
acceptable standards for commercial biodiesel, thereby establishing its suitability to be used
as a substitute to (or as a blend with) petro-diesel. Since the process is optimized, the process
M

can be scaled up for commercial production with minor adjustments (if required), thereby
providing a high yield of biodiesel per unit quantity of raw feedstock used. This in turn can
ED

increase the availability and reduce the price of commercial biodiesel.

2. Materials and Procedure


PT

2.1. Materials required


E

Castor (Ricinus communis) oil of commercial grade was purchased from the local market to
CC

be used as raw material for the production of biodiesel. The oil was subjected to filtration and
drying in a hot air oven at a temperature range of 100 – 110 °C to remove impurities and
moisture. All the reagents purchased for use in this study were of analytical grade. Solvents
A

used were Propan-2-ol (Qualigens, Mumbai, India) and Methanol (99.50%; Merck, India).
Sulphuric acid (98%) and Potassium Hydroxide (anhydrous) were acquired from Merck,
Mumbai, India. Arium 611 DI ultra-pure water system of Sartorius A. G., Gottingen,
Germany was used for obtaining deionized water.

2.2. Analysis of castor oil

5
The castor oil was characterized following ASTM standardized experiments in our
laboratory. Acid value (AV) gives an idea of the amount of free fatty acids (FFA) present in
the oil used for the experiment. It can be quantified by the amount of potassium hydroxide
required to nullify the fatty acids in a known quantity of the sample [27]. Acid value of castor
oil was estimated by ASTM Method D974.

For determining the acid value and primary FFA content, a solution of 0.1 N KOH (aqueous)
was prepared to be used as a standard solution. 0.8 g sample of castor oil was taken and 40 ml

T
of propan-2-ol (used as solvent) was added. Two drops of phenolphthalein solution to be used

IP
as an indicator was then added. The solution was titrated by drop-wise addition of aqueous
KOH with continuous shaking until a faint yet stable pink colour was observed to appear.

R
Addition of KOH was continued, intended for colour persistence upto 30 minutes. The

SC
quantity of 0.1 N aqueous KOH required was noted [28]. Performing the titrations in
triplicates ensured that the results obtained were precise. From these analyses, acid value
(AV) and free fatty acid content (FFA %) can be calculated from Equation (1) and (2):
5 6 .1  N  t U
N
AV  (1)
m
A
2 8 .2  N  t
FFA%  (2)
m
M

where t is the titre value, N = 0.1 denotes the strength of the KOH solution (normality), m is
the mass of the sample of oil taken, while 56.1 is the molar weight of KOH and 28.2 denotes
ED

the presence of oleic acid present in the oil sample and its weight.

2.3. Experimental matrix designing by L16 orthogonal array


PT

In the acid catalyzed esterification of commercial grade castor oil, the reduction in FFA%
relies on a multitude of parameters such as: the molar ratio of oil to methanol, duration of
E

reaction, reaction temperature, the concentration of acid catalyst and speed of agitation [15].
CC

In optimization approaches like the conventional ‘one-factor-at-a-time’ or central composite


design (CCD), a large number of experiments are usually required to be run in order to
establish the best parametric conditions for any process. The Taguchi technique developed by
A

Dr. Genichi Taguchi has the very specialty of not investigating every possible combination of
parameters, essentially reducing the cumbersome optimization procedure to only a few
combinations. For the elucidation of factors with the significant influence on the yield of a
desirable product, a minimum number of experiments need to be carried out, making the
collection of data easier as well as reducing precious time and resources. Based on the effect,

6
the most significant parameter can thus be estimated [29-30]. For the present research, the
number of parameters P and the number of variation levels L of each parameter dictating the
number of runs N necessitate the selection of an L16 orthogonal array. The Taguchi design of
experiments (DOE) was done using Design Expert 10.0. Table 1 provides a comprehensive
summary as recognized using Equation (3):

N   L  1 P 1 (3)

2.4. Experimental Procedure

T
Castor oil esterification was carried out in a three-necked flat round-bottomed flask using

IP
either homogeneous acid catalyst (H2SO4) or heterogeneous sulfonated carbon catalyst. For

R
the preparation of heterogeneous catalyst, carbonized and steam activated Mesua ferrea linn
(MFL) seed shells prepared by Bora et. al. were used [16] following the sulfonation process

SC
used by them. One neck fitted with a thermometer was used to measure the temperature of
the mixture, while a reflux condenser was fixed on the other neck to decrease the evaporative

U
loss of methanol by refluxing it back to the reactor. The middle neck was used for the
N
addition of reagents. After heating the oil to the pre-set temperature, methanol was
transferred to the reactor. The addition of methanol leads to a drop in temperature. A
A
magnetic stirrer hotplate allows the adjustment of the plate temperature to keep the mixture
M

temperature constant and set the agitation speed in accordance with the experimental
conditions selected [9, 31]. After attainment of the desired temperature, an acid catalyst or
ED

sulfonated carbon was added in the chosen weight%. The reaction was allowed to take place
for the specified duration in the experimental design matrix represented in Table 2. The
PT

mixture was then kept in ice-cold water for a few minutes in order to arrest the reaction, since
esterification cannot occur at temperatures lower than 20°C. For removal of sulfonated
carbon catalyst, Whatman series 40 filter paper is used. The mixture is then distilled to
E

recover the unused methanol in pure form for subsequent use; this reduces the overall cost. It
CC

is then emptied into a separating funnel to be left standing overnight to let the mixture
separate into two distinct layers. The top layer is the FAME rich phase while the lower layer
A

is an aqueous phase, which was drained out. After recovering the FAME formed, it was
washed with deionized water several times to remove acid catalyst and residual alcohol. The
refining is necessary in order to purify the crude biodiesel to be within the acceptable
international standards [22]. The sample was then analyzed for the reduction in FFA%.

2.5. Analysis of variance (ANOVA)

7
A statistical analysis of variance (ANOVA) can be useful for determining the process
parameters which are significant from a statistical point of view. The calculation of S/N
(signal to noise) ratio depends on the results obtained from experiments. By S/N ratio
analysis, the optimum level of each parameter along with the optimized set of parameters
achieving the highest yield of FAME can be ascertained. Then the optimal combination of the
parameters controlling the process can be predicted. However, it is unable to explain the
contribution of distinct parameters on the FFA conversion. This can be accomplished by
analysis of variance by estimating the statistical parameters [38]. In ANOVA study, Fischer

T
test (F-value) determines the significance of the chosen model and individual parameter

IP
affecting the response. However, p-value gives an idea about the probability of getting the F-

R
value of this extent. Sum of squares obtained for the model and individual parameter was
used to calculate the contribution factor of each parameter to check its influence on biodiesel

SC
yield. A final experiment was conducted in triplicates for validation of predictions at the
optimal process parameters and to check for any disparities between actual and predicted

U
response [15]. The contribution factor for every parameter provides an idea about its impact
N
on the process and can be calculated by Equation (4):
A
(4)
M

Where, SSf = the sum of squares for the fth variable and SST = sum of squares of all
variables. This analysis is aimed at increasing the percentage conversion of FFA. Therefore,
ED

while choosing condition for optimization, a target was selected as maximum. A linear
regression model equation can correlate the predicted response with the actual response
PT

(obtained by experimental runs) for all process variables, which can be useful in validating
the model [15].
E

2.6. Characterization of methyl esters


CC

Gas chromatography (GC) analysis of the biodiesel obtained helped in understanding the
composition of the fuel by confirming the presence of the different types of methyl esters.
A

Gas chromatography was performed using Agilent GC system 7820A. Physico-chemical


properties of the RCME (Ricinus communis methyl ester) produced can be accessed through
ASTM (American Society for Testing and Materials) test methods. Viscosity of a fuel can be
related to atomization ease and combustion efficiency of a fuel. Kinematic viscosity was
estimated with a Redwood viscometer. Properties like flash point and fire point are useful in
estimating the lowest temperature at which a fuel gives off vapours, which form an ignitable

8
mixture with air. Ease of storage and transportation become predictable from these values
[40]. Flash point and fire point was assessed with closed cup Pensky Martens apparatus. Acid
value and residual FFA% of RCME obtained can be estimated through titrimetric assay [25].
Aniline point is useful in estimating the amount of aromatic hydrocarbons (such as paraffins)
in the RCME. A low aniline point indicates that the biodiesel contains a high amount of
aromatic compounds. Cetane number of biodiesel indicates a fuel’s ignition delay. A higher
cetane number indicates better combustion efficiency of a fuel.

T
3. Kinetics of castor oil Esterification

IP
A kinetic model equation for FFA conversion through esterification depends on assumptions
such as:

R
(a) The esterification reaction is reversible in nature.

SC
(b) The rate of non-catalysed reaction is insignificant in comparison to the acid catalyzed
reaction, which occurs in the oil phase.

U
(c) The methanol added to the oil was high enough for it to remain constant during the
N
process.
A
Experiments were run under optimized conditions established by Taguchi approach. Samples
withdrawn at fixed time intervals in centrifuge tubes were immediately placed in ice-cold
M

water to arrest the reaction. Samples were then washed with an equal volume of water in
order to remove the residual acid catalyst and other undesirable by-products formed. Oil
ED

sample forms the upper layer which was analyzed for the formation of RCME. The residual
FFA value obtained can be converted into a derivative using finite differentiation [17]. The
PT

samples were analyzed in triplicates for 3 different temperatures to ensure that the results
obtained are accurate.
E

A simplified form of the homogeneous esterification reaction when a surplus of methanol is


CC

provided is demonstrated by Equation (5):


R1 – COOH + R2 – OH R1 – COOR2 + H2O (5)
A

This reaction is catalyzed by H2SO4. R1 is usually a linear chain of carbon atoms containing a
variety of unsaturations, while R2 represents a methyl group. The reaction is reliant on the
concentrations of both FFA and methanol. This suggests that the esterification ought to
follow second order kinetics. Nevertheless, in reality, a large excess of methanol is provided
to shift the equilibrium towards the formation of RCME [26]. Thus, the reaction follows

9
pseudo-first order kinetic law, since the concentration of methanol is high enough not to
affect the stoichiometry of the reaction.
The rate of reaction (r) apropos to time was evaluated by the use of Equation (6):

d C FFA
r    k T  (6)
dt

Here k denotes the pseudo – first order reaction rate constant, which depends on temperature
(T) and can be also used in Arrhenius Equation (7):

T
EA

k T   kr  Ae RT
(7)

IP
where EA is the activation energy, A is the frequency factor and R is the universal gas

R
constant (molar). Concentration of FFA apropos to time at distinct temperatures was

SC
d C FFA
converted to their corresponding derivatives by finite differentiation for calculating at
dt

each time interval. Integrating Equation (6) within limits gives Equation (8).

𝐶𝐹𝐹𝐴(0)
ln 𝐶 = kt U (8)
N
𝐹𝐹𝐴(𝑡)
A
The value of k is the slope of a plot of ln C FFA(t) against t. The values for EA and A for acid
esterification can be evaluated using Equation (7) and by taking the natural log of Equation
M

(7).
EA
ED

ln k  ln A  (9)
RT

Using plots of ln k versus reciprocal of temperature (T) provides the values for EA and A. The
PT

EA of the acid esterification was calculated from the slope (EA/R), and A was calculated from
antilog of the intercept.
E

4. Results and Discussions


CC

4.1. Feedstock analysis results

Castor oil before esterification and the RCME obtained after esterification are shown in
A

Figures 2a-2d. Analysis of the castor oil obtained for biodiesel production helped in
estimating its characteristic properties prior to utilization for acid esterification. Values
obtained from the feedstock analysis are summarized in Table 3. The FFA content of the
castor oil is greater than 2% w/w, which means that the feedstock needs to be acid esterified
in order to obtain RCME.

10
4.2. Analysis of Variance (ANOVA) of esterification

The esterification process of Ricinus communis oil was performed in accordance with the set
of parametric conditions obtained by experimental matrix design using L16 orthogonal array
approach. The reduction in FFA concentration was taken to be the performance characteristic
and the result obtained after each run was fed as a response and reported in Table 2.The drop
in FFA content of RCME after esterification was analyzed statistically for an estimation of
the significance of the model chosen for optimization and the effects of individual process

T
parameters on the response by means of ANOVA studies. ANOVA is a vigorous method

IP
used to examine the legitimacy and prominence of an individual parameter and chosen
optimization model for the development of a mathematical model equation. The ANOVA

R
results of the esterification carried out are tabulated in Table 4. It focuses on the analysis of

SC
the variance around the mean of the performance characteristics and is accomplished by
estimating the Fischer’s test value (F-value). The impact of any parameter is explained by its
F-value and corresponding sum of squares. Higher F-value and sum of squares of any
U
parameter indicates its relative importance in the process of the response. Conversely, the
N
extents of obtained values of these tests are exclusively owed to response signals, and are
A
assured by a p-value. The p-value clarifies the probability of obtaining an F-value of this
order owing to noise; values below 0.05 or 5% attest the importance of the particular process
M

parameter. In the present study, the obtained sum of squares for the model is 2067.65, and is
thus high enough to be considered as significant. The significance of the model is implied by
ED

the model F-value of 58.92 and there is only a 0.32% chance (p value = 0.0032) that an F-
value this large could occur due to noise. This verifies the model to be significant in the
PT

optimisation of the esterification of Ricinus communis oil in the selected levels of process
parameters as all test values attained have greater magnitude and those are chiefly due to
E

adequate response signals. The ANOVA study also showed that amidst the five selected
CC

parameters, only four parameters significantly affect the response. As seen, duration of
reaction has a negligible effect on the FFA conversion for the present study. Among the
significant parameters, molar ratio of oil to methanol has shown massive influence on the
A

FFA conversion with F-value and sum of squares 140.55 and 1233.03 respectively. There
exists only a 0.32% chance (p value = 0.0032) of obtaining these higher values due to
noise.Hence it shows the importance of molar ratio of oil to methanol in the esterification
procedure. Besides, agitation speed, reaction temperature and catalyst concentrations have

11
impacts on the FFA conversion but they are less significant than the molar ratio of oil to
methanol with F-values of 39.95, 29.68 and 25.51 respectively.

The impacts of individual parameters are again established by the assessment of the
contribution factor (calculated by Equation 4 and tabulated in Table 6) of each parameter in
the process. The results obtained demonstrates that molar ratio of oil to methanol has the
highest contribution percentage of 59.6%, thereby affecting the process largely which is
followed by agitation speed which has a contribution percentage of 16.95%, reaction

T
temperature which has a contribution percentage of 12.59% and lastly, catalyst concentration

IP
affects the FFA reduction by the lowest of 10.82%. The experimental results affirm that the
ANOVA results are in accordance with the contribution factor results and attest that molar

R
ratio of oil to methanol has the highest effect on the FFA reduction.

SC
Mathematical exploration of the acid esterification helped in an evaluation of the importance
and applicability of the model chosen. The experimental responses obtained are used in

U
estimation of the correlation coefficient (R2) and perfect fit are evident from the R2 value of
0.9958. Adjusted R2 has a value of 0.9789, the predicted R2 being 0.8798. The values are in
N
reasonable agreement with one another, the difference is less than 0.2, which shows that the
A
variance around the mean of the response elucidated from the model and the ability of this
M

model in prediction of response is appreciable. The desired precision value of a model is


4.The present study has a precision value of 29.367 which shows that the model is apt in
ED

response prediction and optimization of the esterification. Standard deviation (SD) is termed
as the root mean square error; for the present study, it is 1.71, while the coefficient of
variance is 2.96%. A low value of SD and coefficient of variance shows that this model can
PT

predict the best conditions with high precision [5, 39]. Stated results are presented in Table 5.
E

4.3. Regression model equation


Model and individual parameter analysis were done by ANOVA and parameters significantly
CC

affecting the process were identified. A regression analysis conducted with the significant
parameters while omitting insignificant parameters gave coefficients of each significant
A

parameter (obtained with 95% assurance) which were taken into account in the mathematical
model equation for the FFA conversion in Equation 10:
FFA conversion% = 57.87 - 3.76 * A[1] + 6.80 * A[2] – 1.45 * A[3] + 0.26 * C[1] + 6.02*
C[2] – 3.16 * C[3] – 10.85 * D[1] – 4.90 * D[2] + 3.28 * D[3] – 5.20 *
E[1] – 3.80 * E[2] + 6.17 * E[3] (10)

12
Where, A[1], A[2] and A[3] are reaction temperatureat the first, second and third level
respectively, C[1], C[2] and C[3] are catalyst concentration at the first, second and third level
respectively,D[1], D[2] and D[3] are molar ratio of oil to methanol at the first, second and
third level respectivelyand lastly, E[1], E[2] and E[3] are agitation speed at the first, second
and third level respectivelyin accordance with the values presented in Table 1. The regression
model equation can predict the FFA conversion within the given set of parametric conditions.
Predicted FFA conversion values apropos to the experimentally obtained results were plotted
and depicted in Figure 3. The plot shows that the results are mutually agreeable and the

T
significance of the model equation for an accurate calculation of the response [25].

IP
4.4. Optimisation and model validation

R
Castor oil esterification must be optimized for an understanding of the best parametric

SC
conditions that can maximize FFA conversion. Observations from ANOVA are: duration of
reaction is an insignificant parameter, thus lowest value of 1 hour is sufficient for maximum

U
FFA conversion. For the process to be economic and sustainable, every significant parameter
should be at its lowest level. However, achieving the best condition for maximum FFA
N
reduction requires each parameter such as reaction temperature, catalyst concentration, oil to
A
methanol molar ratio and agitation speed to have a suitable numeric value. The optimum
M

conditions obtained for the present study are: reaction temperature 50°C, catalyst loading 1%
w/w, molar ratio of oil to methanol as 1:20 and agitation speed of 700 rpm which gives
ED

maximum conversion of 89.34%. Optimum conditions obtained from the model were again
authenticated by carrying out the confirmatory experiments thrice; the experimental value
(90.83%) is comparable to the model predictions. Clearly, the chosen model is able to predict
PT

the optimum conditions of castor oil esterification precisely. Optimum conditions for the
esterification are summarized in Table 7. A comparative test study between esterification at
E

optimized conditions using H2SO4 as homogeneous acid catalyst and sulfonated carbon as
CC

heterogeneous catalyst was conducted, which showed that homogeneous catalysis is able to
provide better conversion of the FFAs present in castor oil, as evident from Table 8.
A

4.5. Effects of individual process parameter on FFA conversion

Process variables can have variable effects on FFA reduction in oil. They are: reaction
temperature, reaction time, catalyst concentration and molar ratio of oil to methanol. Studies
were carried out to derive prime prerequisites for the esterification reaction between Ricinus
communis oil and methanol using H2SO4 as a catalyst for maximum FFA reduction.

13
4.5.1. Effect of reaction temperature

The effect of temperature on FFA conversion was studied by a variation of the reaction
temperature at four different levels set at values 40, 50, 60 and 70 ̊C. In the present study,
under reaction conditions of reaction time = 1 hr, catalyst concentration = 1% w/w, methanol
to oil molar ratio = 20:1 and agitation speed = 700 rpm, maximum FFA conversion was
observed at the temperature of 50 ̊C, as demonstrated by Figure 4a. The increasing trend in
FFA conversion apropos to temperature was noted up to 50 °C. However, additional rise in

T
temperature demonstrates a reduction in the FFA conversion.

IP
Optimization of the process at low temperatures is favourable because it requires a lower
amount of thermal energy consumption for the reduction of FFAs. Up to 50 °C, increased

R
molecular interactions (due to rise in reaction temperature) between the reactants seems to

SC
escalate the rate of reaction. Also, at high temperatures, acid esterification unfavourable as
the forward reaction is impeded due to evaporative loss of methanol, which disrupts the

U
methanol to oil molar ratio [32] and also shifts the reaction equilibrium away from product
formation, resulting in saponification. Results reported by S.H. Dhawane et. al [16], S.
N
Halder et. al. [13] and Leung and Guo [33] are in agreement with our study.
A
4.5.2. Effect of methanol to oil molar ratio
M

Molar ratio of oil to methanol is the most important factor, as established from the
esterification of castor oil. Under reaction conditions of reaction temperature = 50 °C,
ED

reaction time = 1 hr, catalyst concentration = 1% w/w and agitation speed = 700 rpm,
maximum FFA conversion was obtained at molar ratio of 1:20, as demonstrated by Figure 4b.
PT

Theoretically one mole of oil reacts with 3 moles of alcohol to yield biodiesel and water.
However, the alcohol required is always in large excess to drive the equilibrium of the
E

reaction towards the formation of products; also, the optimum use of alcohol is quite crucial;
CC

this is achievable by proper adjustment of the molar ratio. It can be conclusively stated that
addition of surplus amounts of methanol drives the reaction towards FFA conversion [17].
Pradhan et al. (2012) reported that temperatures above the boiling point of methanol are
A

unproductive because the evaporation of methanol leads in the decrease in biodiesel yield
[47-48].

4.5.3. Effect of reaction time

Under reaction conditions of reaction temperature = 50 °C, catalyst concentration = 1% w/w,


methanol to oil molar ratio = 20:1 and agitation speed = 700 rpm, maximum FFA conversion
14
was obtained at reaction time = 1 hour. With increase in time from 1 hour to 2.5 hours, no
significant improvement in FFA conversion was observed.

Mass transfer between oil and alcohol is enhanced when adequate residence time is provided
for the reactants to interact [17]. Observations suggested that the duration of reaction is an
insignificant parameter during castor oil acid esterification. This could be due to a vigorous
reaction between the reactants in the presence of the catalyst. Kinetic studies that were
carried out later, allowed for a better understanding of this performance characteristic.

T
Without any improvement in yield, an extended reaction time is unnecessary as production

IP
cost gets high [34] Results are in agreement with M. Chai et. al.(2014), who also reported
other works in their article that state on a large scale, most researchers hold the view that for

R
economical production of biodiesel with high yields, a balance between reaction time and

SC
temperature needs to be maintained.

4.5.4. Effect of catalyst concentration

U
Under reaction conditions of reaction temperature = 50 °C, reaction time = 1 hour, methanol
N
to oil molar ratio = 20:1 and agitation speed = 700 rpm, catalyst concentration was varied
from 0.5% to 2%. maximum FFA conversion was obtained at catalyst concentration = 1%
A
w/w. The maximum reduction in FFA% was obtained at an acid concentration of 1%, as
M

illustrated by Figure 4c.

Ricinus communis oil used in this study contains considerable amounts of free fatty acids
ED

(4.58%). Therefore, the oil is unsuitable for trans-esterification with base catalysts and needs
to be acid-esterified for biodiesel production. Elevation in catalyst concentration beyond 1%
PT

w/w had a negative impact on FFA conversion. This is because the acid catalyst cannot
provide catalytic activity beyond a certain concentration in the esterification process [31].
E

With the cost of production in mind, 1% w/w H2SO4 or sulfonated carbon catalyst is
CC

preferred. Results reported by Ridha et al. [31], Agra et al. [21] and Goyal et al [35] are in
agreement with our findings.

4.5.5. Effect of agitation speed


A

Maximum FFA conversion was observed at an agitation speed of 700 rpm, when catalyst
concentration was at 1% w/w, reaction temperature was at 50 ̊C, duration of reaction was 1
hour and oil to methanol ratio was at 1:20, during experiments run according to L16 matrix
recommendations.

15
Even if one of the reactants is viscous, as is the case with castor oil, agitation speed will
affect the reaction. Agitation speed proves to be an important parameter in the FFA
conversion, as illustrated by Figure 4d. FFA conversion increases with increasing agitation
speed, implying that turbulence caused due to high agitation speed leads to increased
frequency interaction of the reactants at a molecular level. High agitation also diminishes the
resistance encountered during mass transfer, hence the rate of reaction and FFA conversion
increase. Beyond 700 rpm, FFA conversion is observed to decrease slightly, probably due to
insufficient interaction time between the reactants upon extreme agitation [15]. Observations

T
made by Dhawane et. al., who in one of their works, optimized the synthesis of Hevea

IP
brasiliensis (rubber seed) oil and in another optimized the synthesis of iron (II) doped

R
carbonaceous catalyst, conform to our findings.

SC
4.6. Characterization of Biodiesel

The obtained biodiesel was analysed for the compositional variance in the methyl esters

U
contained by Gas chromatography (GC). The analysis revealed that the fatty acids present in
the oil had been successfully converted into various methyl esters, as supported by Figure 5.
N
A summary of the results obtained from the chromatography analysis has been summarized
A
in Table 9. Characteristics of the RCME obtained are assessed and compared with the
M

properties of petro-diesel. Characteristics assessed in accordance with ASTM standards are


summarized in Table 10. The properties were found to be within the suggested range for
ED

biodiesel and comparable to petrol-diesel. The FFA% of the biodiesel obtained is very close
to 0.5 which makes it suitable for use in unmodified engines without problems such as
corrosion of the metallic parts [16]. Moreover, the aniline point noted is high, which means
PT

that the fuel has low aromatic hydrocarbons. The RCME is also safer for storage and
transportation compared to petro-diesel, as evident from the high flash point and fire point.
E

Cetane number is within the acceptable range for commercially usable biodiesel and is quite
CC

higher than the values for petro-diesel, indicating that the biodiesel obtained has a low
ignition delay and is a superior alternative to petro-diesel.
A

4.7. Kinetic Studies

4.7.1. FFA Concentration versus Time Plot

The kinetic studies were performed with the aim of establishing the order of kinetics of the
homogeneous acid esterification process. The process was carried out from the lowest level
of temperature considered in the experimental matrix up to the temperature at which

16
maximum FFA conversion was noted, to understand the effect of rising temperature on FFA
conversion. According to previously established optimized conditions, the change in FFA
conversion was determined at temperature intervals of 5 ̊C within the range 40-50 ̊C by the
withdrawal of samples at time intervals of 5 minutes during the esterification. The reactions
were carried out beyond the optimized duration by an additional 30 minutes to determine
reaction equilibrium. The chosen range of temperatures to study the kinetics was obtained
from the optimisation study and the rate of esterification is found to be prominent in this
range.

T
IP
A plot of ln CFFA (t) versus time t (Figure 6) illustrates the effect of reaction temperature on
the conversion of the FFA present in oil to be significant. Results obtained revealed the

R
increase in reduction of FFA concentration from 40 to 50 ̊C. At lower temperatures, the rate

SC
of acid esterification was lower, as evident from the increasing k values obtained at higher
temperatures. It can be inferred that an increase in the temperature of reaction shifts the
equilibrium to the right, increasing the FFA conversion. Reaction rate constants for different
temperatures are summarized in Table 11. U
N
4.7.2. Activation Energy and Frequency Factor
A
The estimation of EA and A was done from Figure 7, which depicts the negative logarithm
M

plot of rate constants (-ln k) of the reaction apropos to the inverse of reaction temperature
(1/T). A straight line was obtained which substantiated the reaction rate constants to be
ED

concurrent with Arrhenius’ law. The observations made are briefed in Table 12. The
procedure of acid esterification has lower activation energy [36-37]. Low EA is favourable as
PT

the reaction tends to occur faster, thereby reducing the reaction time and increasing the
conversion of FFA. The low requirement of activation energy signifies the effective catalysis
of esterification at lower energy expenditure.
E
CC

5. Conclusion
In this study, the objective was to develop a cost-effective procedure of biodiesel synthesis
from commercial grade castor oil, which was subjected to acid esterification under the
A

experimental conditions suggested by L16 Taguchi orthogonal array. Maximum FFA


conversion of 90.83% was obtained under the following conditions: reaction temperature of
50 °C, reaction duration of 1 hour, catalyst (H2SO4) concentration of 1% w/w, methanol to oil
ratio of 20:1 and agitation speed of 700 rpm. Under the optimized conditions, a comparative
test run using H2SO4 as homogeneous catalyst and sulfonated carbon (obtained from

17
carbonised and steam activated Mesua ferrea linn seed shells) as heterogeneous catalyst
showed that homogeneous catalyst gives better conversion of the FFAs in castor oil. Molar
ratio of oil to methanol has the highest contribution to the FFA conversion, while the duration
of reaction was found to be insignificant. The ANOVA study and statistical analysis results
showed that the 3FI model is able to optimize the process efficiently and accurately with
precise prediction. The validity of the chosen model was confirmed by conducting
experiments at the predicted optimum conditions and the results obtained were in good
agreement with the predicted response. The L16 orthogonal array approach chosen was

T
therefore effective in identifying the significant parameters of castor oil esterification and

IP
optimizing them.Kinetic studies performed showed that the esterification reaction for FFA

R
conversion follows pseudo-first order kinetics with a very fast conversion rate of 0.0392 min-
1
and low activation energy of 8.184 kJ/mol which signifies that the reaction occurs very

SC
vigorously with a minimum energy requirement. The biodiesel obtained was subjected to GC
analysis in order to determine the compositional percentages of the methyl esters obtained

U
after esterification of the castor oil. Standard fuel characterization experiments confirmed that
N
the physic-chemical properties of the RCME produced fall within the acceptable range of
A
commercially usable biodiesel, being both low in aromatic hydrocarbon content and stable
with a high flash point and fire point, making it suitable for storage and transportation.
M

Cetane number tests confirmed that the fuel is of acceptable quality for use as commercial
biodiesel with lower ignition delays than petro-diesel. Hence it can be blended or used in its
ED

entirety as an alternative to petro-diesel.


PT

Acknowledgement

The authors would like to express their sincere gratitude for the monetary support from the
E

Ministry of Science and Technology, Govt. of India through research project no.
CC

EEQ/2016/000243.
A

References:

[1] Marchetti J.M. Miguel V.U., Errazu A.F., Heterogeneous esterification of oil with high
amount of free fatty acids, Fuel (2010), 86: 906–910.

[2] Keskin A., Gürü M., Altiparmak D., Aydin K., Using of cotton oil soapstock biodiesel–
diesel fuel blends as an alternative diesel fuel, Renewable Energy (2008), 33: 553–557.
18
[3] Li W., Du W., Liu D., Optimization of whole cell-catalyzed methanolysis of soybean oil
for biodiesel production using response surface methodology, Journal of Molecular Catalysis
A: Chemical (2007), 45: 122–127.

[4] Vicente G., Coteron A., Martinez M., Aracil J., Application of the factorial design of
experiments and response surface methodology to optimize biodiesel production, Industrial
Crops and Products (1998), 8: 29–35.

[5] KılıçM., Uzun B.B., Pütün E., Pütün A.E., Optimization of biodiesel production from

T
castor oil using factorial design, Fuel Processing Technology (2013) 111: 105–110

IP
[6] Janaun J, Ellis N., Perspectives on biodiesel as a sustainable fuel, RenewableSustainable

R
Energy Rev (2010), 14:1312-1320.

SC
[7] Saloua F, Saber C, Hedi Z., Methyl ester of [Maclurapomifera (Rafin.) Schneider] seed
oil: biodiesel production and characterization, BioresourTechnol (2010), 101: 3091-3096.

U
[8] Ramezani K., Rowshanzamir S., Eikani M.H., Castor oil transesterification reaction: A
kinetic study and optimization of parameters, Energy (2010), 35: 4142-4148.
N
[9] Zhang Y., Dube M.A., McLean D.D., Kates M., Biodiesel production from waste cooking
A
oil: Economic assessment and sensitivity analysis, BioresourTechnol (2003), 90: 229–40.
M

[10] Bart J., Palmeri N., Cavallero S., Biodiesel science and technology: from soil to oil.
Florida. EstadosUnidos: CRC Press; (2010).
ED

[11] Zou H., Lei M., Optimum process and kinetic study of Jatropha curcas oil pre-
esterification in ultrasonical field, J Taiwan Inst Chem Eng. (2012), 43:730-735.
PT

[12] Berchmans HJ, Hirata S., Biodiesel production from crude Jatropha curcas L. seed oil
with a high content of free fatty acids, Bioresour Technol. (2012), 99:1716-1721.
E

[13] Halder S., Dhawane S.H., Kumar T., Halder G., Acid-catalyzed esterification of castor
CC

(Ricinus communis) oil: optimization through a central composite design approach, Biofuels
(2015), 6:191-201.
A

[14] Torrentes-Espinoza G., Miranda B.C., Vega-Baudrit J., Mata-Segreda Julio F., Castor oil
(Ricinuscommunis) supercritical methanolysis, Energy (2017), 140: 426-435.

[15] Dhawane S.H., Kumar T., Halder G., Biodiesel synthesis from Hevea brasiliensis oil
employing carbon supported heterogeneous catalyst: Optimization by Taguchi method,
Renewable Energy (2016), 89:506-514.

19
[16] Dhawane S.H., Bora A.P., Kumar T., Halder G., Parametric optimization of biodiesel
synthesis from rubber seed oil using iron doped carbon catalyst by Taguchi approach,
Renewable Energy (2017), 105:616-624.

[17] Bokhari A., Chuah L.F. et. al., Optimisation on pretreament of rubber seed (Hevea
brasiliensis) oil via esterification reaction in a hydrodynamic cavitation reactor, Bioresource
Technology (2016).

[18] Bueno A.V., Bento Pereira M.P., de Oliveira Pontes J.V., Tavares de Luna F.M.,

T
Cavalcante Jr. C.L., Performance and Emissions Characteristics of Castor Oil Biodiesel Fuel

IP
Blends, Applied Thermal Engineering (2017)

R
[19] Philips R., Rix M., Annuals and Biennials, London: Macmillan (1999), 106.

SC
[20] Felizardo P., Correia M.J.N., Paposo I., Mendes J.F., Berkemeier R., Bordado J.M.,
Production of biodiesel from waste frying oils. Waste Manage (2006), 26: 487–94.

U
[21] Agra IB, Warnijati S, Wiratni, Two steps ethanolysis of castor oil using sulphuric acid as
catalyst to produce motor oil, Werc. (1996), 9: 1025-1028.
N
[22] Marchetti J.M., Miguel V.U., Errazu A.F., Possible methods for biodiesel production.
A
Renew. Sustain. Energy Rev. (2007), 11: 1300–1311.
M

[23] Dias, J.M., Araújo, J.M., Costa, J.F., Alvim-Ferraz, M.C.M., Almeida, M.F., Biodiesel
production from raw castor oil. Energy (2013), 53: 58–66.
ED

[24] Encinar, J.M., González, J.F., Pardal, A., Transesterification of castor oil under
ultrasonic irradiation conditions. Preliminary results.Fuel Process. Technol. (2012), 103: 9–
PT

15.

[25] Dhawane S.H., Kumar T., Halder G., Central composite design approach towards
E

optimization of flamboyant pods derived steam activated carbon for its use as heterogeneous
CC

catalyst in transesterification of Heveabrasiliensis oil, Energ Convers Manage. (2015),


100:277-287.
A

[26] Ghaly A.E, Dave D., Production of Biodiesel by Enzymatic Transesterification: Review,
American Journal of Biochemistry and Biotechnology (2010).

[27] Methods for testing Commercial fatty acids,


https://www.aocs.org/Documents/TechnicalPDF/Methods/6th/S1_64.pdf

20
[28] EN 14104:2003., Fat and oil derivates. Fatty Acid Methyl Esters (FAME).Determination
of acid value.

[29] Kumar R.S., Kumar K.S., Velraj R., Optimization of biodiesel production from
Manilkarazapota (L.) seed oil using Taguchi method, Fuel (2015), 140: 90-96.

[30] Phadke M.S., Quality Engineering Using Robust Design, Pearson Education, New Delhi
(2008).

[31] Banani R., Youssef S., et al., Waste Frying Oil with High Levels of Free Fatty Acids as

T
one of the prominent sources of Biodiesel Production, J. Mater. Environ. Sci., 6 (2015), 4:

IP
1178-1185.

R
[32] Arora R., Kapoor V., Toor A.P., Esterification of free fatty acids in waste oil using a

SC
carbon-based solid acid catalyst, ICETET (2014), 196-199.

[33] Leung D.Y.C., Guo Y., Transesterification of neat and used frying oil: optimization for

U
biodiesel production”, Fuel Process Technol. (2006), 87:883-890.
N
[34] Chai M, Tu Q., Lu M., et al., Esterification pretreatment of free fatty acid in biodiesel
production, from laboratory to industry, Fuel Process Technol. (2014), 125: 106-113.
A

[35] Goyal P., Sharma M.P., Jain S., Optimization of esterification and transesterification of
M

high FFA Jatropha curcas oil using response surface methodology, J Petro Sci Res. (2012), 1:
36-43.
ED

[36] Rattanaphra D., Harvey A.P., Thanapimmetha A., Srinophakum P., Kinetic of myristic
acid esterification with methanol in the presence of triglycerides over sulfated zirconia,
PT

Renew. Energy (2011), 36: 2679–2686.

[37] Marchetti J.M., Pedernera M.N., Schbib N.S., Production of biodiesel from acid oil
E

using sulphuric acid as catalyst: kinetics study, Int. K. Low Carbon Technol. (2012), 6: 38–
CC

43.

[38] Kumar S., Chary G.H.V.C., Dastidar M.G., Optimization studies on coal–oil
A

agglomeration using Taguchi (L16) experimental design, Fuel (2015), 141:9–16.

[39] Dhawane S.H., Kumar T., Halder G., Parametric effects and optimization on synthesis of
iron (II) doped carbonaceous catalyst for the production of biodiesel, Energy Conversion and
Management (2016), 122:310–320.

21
[40] Canoira L., Garcıa Galea´n J., Alca´ntara R., Lapuerta M., Garcıa-Contreras R., Fatty
acid methyl esters (FAMEs) from castor oil: Production process assessment and synergistic
effects in its properties, Renewable Energy (2010) 35:208–217.

[41] Bezerra M.A., Santelli R.E., Oliveira E.P., Villar L.S., Escaleira L.A., Response surface
methodology (RSM) as a tool for optimization in analytical chemistry, Talanta (2008),
76:965–977.

[42] Leardi R., Experimental design in chemistry: a tutorial, Anal. Chim. Acta (2009),

T
652:161–172.

IP
[43] Akalın M.K., Karagöz S., Akyüz M., Supercritical ethanol extraction of bio-oils from

R
German beech wood: Design of experiments, Industrial Crops and Products (2013), 49:720–
729.

SC
[44] Ferreira S.L.C., Bruns R.E., da Silva E.G.P., dos Santos W.N.L., Quintella C.M., David
J.M., de Andrade J.B., Breitkreitz M.C., Jardim I.C.S.F., and Neto B.B., Statistical designs

U
and response surface techniques for the optimization of chromatographic systems, J.
N
Chromatogr. A. (2007), 1158:2–14.
A
[45] Chintala R., Clay D.E., Schumacher T.E., Malo D.D., and Julson J.L., Optimization of
M

oxygen parameters determination of carbon and nitrogen in biochar materials, Anal. Lett.
(2013), 46:532–538.
ED

[46] Akalın M.K., Karagöz S., Optimization of Ethanol Supercritical Fluid Extraction of
Medicinal Compounds from St. John's Wort by Central Composite Design, Analytical Letters
(2014), 47:11, 1900-1911.
PT

[47] Pradhan S., Madankar C., Mohanty P., Naik S., Optimization of reactive extraction of
E

castor seed to produce biodiesel using response surface methodology, Fuel (2012), 97:848-
855.
CC

[48] Ngoya T., Aransiola E.F., Oyekola O., Optimisation of Biodiesel Production from Waste
Vegetable Oil and Eggshell Ash, South African Journal of Chemical Engineering (2017).
A

[49] FAOSTAT. 2012. Food and Agriculture Organization of United Nations.


(http/faostat.fao.org).

[50] Available at: http://www.fao.org/faostat/en/#data/QC/visualize

22
T
R IP
SC
U
N
A
M

Figure 1 Global production of castor oil seeds from 2013- 2016 (FAO statistics)
ED
E PT
CC
A

23
T
R IP
SC
Figure 2a – 2b Commercial grade castor oil used as feedstock (side view and top view)

U
N
A
M
ED
E PT
CC
A

Figure 2c – 2d Biodiesel (RCME) obtained after acid esterification of castor oil (side view
and top view)

24
T
R IP
SC
U
N
A
M

Figure 3 Predicted FFA Conversion% vs. Actual FFA Conversion%


ED
E PT
CC
A

25
T
R IP
SC
U
N
A
M

Figure 4a Effect of reaction temperature on FFA Conversion%


ED
EPT
CC
A

26
T
R IP
SC
U
N
A
M

Figure 4b Effect of methanol to oil molar ratio on FFA Conversion%


ED
E PT
CC
A

27
T
R IP
SC
U
N
A
M

Figure 4c Effect of catalyst concentration on FFA Conversion%


ED
E PT
CC
A

28
T
R IP
SC
U
N
A
M

Figure 4d Effect of agitation speed on FFA Conversion%


ED
EPT
CC
A

29
T
R IP
SC
U
N
A
M

Figure 5 Gas chromatography analysis of the RCME obtained


ED
EPT
CC
A

30
T
R IP
SC
U
N
A
M

Figure 6 Influence of reaction temperature through time on the reduction


of FFA% present in castor oil
ED
E PT
CC
A

31
T
R IP
SC
U
N
A

Figure 7 Plot of –ln k vs. 1/T (× 100) for castor oil


M
ED
E PT

Table 1 Process parameters and their values at different levels


CC

Levels
Parameter (unit)
1 2 3 4
A

Reaction Temperature (°C) 40 50 60 70

Reaction Time (hours) 1 1.5 2 2.5

Catalyst Concentration%
0.5 1 1.5 2
(w/w)

32
Methanol : oil molar ratio 5:1 10:1 15:1 20:1

Agitation speed (rpm) 500 600 700 800

Table 2 L16 Experimental Matrix developed by Taguchi method

Methanol : Agitation
Run Temperature Time Catalyst FFA
oil molar Speed
No. (°C) (hours) (w/w) Conversion%
ratio (rpm)

T
1 50 2 2 05:01 600 46.6

IP
2 50 1.5 0.5 20:01 700 84.66

R
3 70 2 1 20:01 500 69.27

SC
4 60 1.5 2 15:01 500 52.46

5 60 2.5 1 05:01 700 56.8

U
N
6 40 2.5 2 20:01 800 65.35
A
7 60 2 0.5 10:01 800 54.3
M

8 40 1 0.5 05:01 500 38.48

9 40 2 1.5 15:01 700 60.1


ED

10 60 1 1.5 20:01 600 62.1

11 70 1.5 1.5 05:01 800 46.18


PT

12 70 1 2 10:01 700 54.6


E

13 70 2.5 0.5 15:01 600 55.06


CC

14 40 1.5 1 10:01 600 52.5

15 50 1 1 15:01 800 76.97


A

16 50 2.5 1.5 10:01 500 50.46

Table 3 Proximate analysis of castor oil

33
Parameter (unit) Value Testing Procedure

Relative density (g/cm3) 0.961 ASTM D287

Acid Value (mg KOH/g) 9.12 ASTM D974

Free Fatty Acid % 4.58 -

Molecular Weight (g/mol) 933.45 -

T
Kinematic viscosity (cm2/s) 2.7 ASTM D445

IP
Table 4 Analysis of variance (ANOVA) of model and process parameters

R
SC
Sum of Mean F p-value
Source df
Squares Square Value Probability > F

Model 2067.65 12 172.30


U
58.92 0.0032 Significant
N
A-Temperature 260.33 3 86.78 29.68 0.0099
A
C-Catalyst
223.80 3 74.60 25.51 0.0123
M

Concentration

D-Methanol to oil
1233.03 3 411.01 140.55 0.0010
ED

molar ratio

E-Agitation speed 350.49 3 116.83 39.95 0.0064


PT

Residual 8.77 3 2.92


E

Cor Total 2076.42 15


CC

Table 5 Statistical parameters estimated from ANOVA study


A

Std. Dev. 1.71 R-Squared 0.9958

Mean 57.87 Adj R-Squared 0.9789

C.V. % 2.96 Pred R-Squared 0.8798

PRESS 249.54 Adeq Precision 29.367

34
-2 Log Likelihood 35.79 BIC 71.83

AICc 243.79

Table 6 Percentage contribution factor for each parameter on FFA conversion%

Parameter (unit) Contribution Factor (%)

Methanol : oil molar ratio 59.6

T
IP
Agitation speed (rpm) 16.95

R
Temperature (°C) 12.59

SC
Catalyst Concentration% (w/w) 10.82

U
Table 7 Optimum conditions for maximum RCME production
N
Parameter (unit) Optimum Value
A

Temperature (°C) 50
M

Reaction duration (hour) 1


ED

Methanol : oil molar ratio 20:1

Agitation speed (rpm) 700


PT

Catalyst Concentration% (w/w) 1


E
CC

Table 8 Optimized yield of acid esterification (homogeneous vs heterogeneous)


A

Method FFA Conversion%

Homogeneous acid esterification 90.83

Heterogeneous acid esterification 71.23

Table 9 Gas chromatography analysis results

35
Retention Area Area % Height Height % Component
Time (min)

12.64 1084884 1.74 459654 4.25 Cis-9-oleic acid methyl


ester

15.122 850284 1.36 252548 2.33 Methyl cis-11-


eicosanoate

T
15.545 2220440 3.55 612804 5.66 Methyl linolenate

IP
16.401 3175050 5.08 872461 8.06 Methyl heneicosanoate

R
23.692 55146844 88.27 8626380 79.70 12R-hydroxy-9Z-

SC
octadecenoic acid methyl
ester

Totals 62477502 100 10823847


U 100 ______
N
Table 10 Fuel characterization of biodiesel (RCME)
A

Parameter (unit) Value Testing Procedure


M

Kinematic viscosity (cm2/s) 2.21 ASTM D445


ED

Acid Value (mg KOH/g) 1.01 ASTM D974

Free Fatty Acid % 0.55 -


PT

Flash Point (°C) 148 ASTM 6450


E

Fire Point (°C) 160 ASTM 6450


CC

Aniline Point (°C) 48 ASTM D611

Cetane Number 47 ASTM D613


A

Table 11 Reaction rate constants at different temperatures

Temperature ̊C Reaction rate constant (k) min-1 R2

40 0.0239 0.9780

36
45 0.0303 0.9517

50 0.0392 0.9339

Table 12 Activation energy and frequency factor

Kinetic Parameters Values

Activation energy EA (kJ/mol) 8.184

T
Frequency Factor A (min-1) 16.62

R IP
SC
U
N
A
M
ED
E PT
CC
A

37

You might also like