Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Fuel Processing Technology 178 (2018) 78–87

Contents lists available at ScienceDirect

Fuel Processing Technology


journal homepage: www.elsevier.com/locate/fuproc

Research article

Purification of crude glycerol derived from biodiesel production process: T


Experimental studies and techno-economic analyses

Chol G. Chola, Ravi Dhabhaia, Ajay K. Dalaia, , Martin Reaneyb,c
a
Department of Chemical and Biological engineering, University of Saskatchewan, SK S7N 5A8, Canada
b
Department of Plant Sciences, University of Saskatchewan, SK S7N 5A8, Canada
c
Guangdong Saskatchewan Oilseed Joint Laboratory, Department of Food Science and Engineering, Jinan University, Guangzhou, Guangdong 510632, China

A R T I C LE I N FO A B S T R A C T

Keywords: In the present work, crude glycerol was purified by a combined strategy of physicochemical treatment and semi-
Glycerol purification continuous membrane filtration using a 5 kDa ultrafiltration tubular membrane. Three parameters – tempera-
Physico-chemical treatment ture, pressure, and flow rate were studied to see the effect of membrane filtration on glycerol purity. A maximum
Membrane filtration glycerol purity of 93.7% was obtained from crude glycerol of 40% purity after the physicochemical treatment
Techno-economic analysis
and membrane filtration at the temperature, pressure, and flow rate of 50 °C, 700 kPa, and 50 mL/min, re-
spectively. Most of the purification occurred during physicochemical treatment. Techno-economic analysis based
on a scenario where all the purified glycerol is converted to value added chemicals – solketal and glycerol
carbonate - showed that the glycerol purification process is economically feasible. In this scenario (scenario 3),
the required capital investment was $2.1 million and the net present value of the project were $6 million (with
10% discounting rate) or $3.65 million (with 15% discounting rate), respectively, over 10 years of operation
after start-up with capital investment in the initial three-year period with no returns. The unit cost and revenue
of crude glycerol purification was $50.85/kg and $80.36/kg, respectively, making it a promising undertaking.
The results of the present work can also be useful for the purification or recovery of other valuable biodiesel by-
products such as free fatty acids, soaps, and solvents.

1. Introduction process, presence of other impurities in the feedstock, and the recovery
of biodiesel, solvents, and catalyst [10]. Most biodiesel production
Glycerol is the main co-product in the biodiesel production process, processes use methanol and sodium or potassium methoxide or hy-
as it is produced in significant quantities, about 10 wt% [1–5]. Glycerol droxide (as catalysts). Accordingly, crude glycerol contains impurities
(C3H8O3) is a sugar alcohol and also known as glycerine, 1,2,3-propa- such as inorganic salts, matter organic non-glycerol (MONG) and water
netriol, glyceritol, glycyl alcohol, or 1,2,3-trihydroxypropane [6,7]. It is [10]. MONG consists of fatty acid methyl esters (FAME), tri-, di- and
a colourless, odorless, and viscous liquid at room temperature and is mono-glycerides, several types of free fatty acids (FFAs), and methanol
biodegradable, hygroscopic, and non-toxic [6,7]. The composition of or ethanol [7].
crude glycerol varies widely (about 30–80%) according to the method The biodiesel industry alone generated about 200,000 t of crude
of biodiesel production [5]. There are four major processes for produ- glycerol in 2003, which exponentially increased to 2 million tonnes in
cing biodiesel and its co-product glycerol. These processes are: trans- 2012. A sizable chunk of this supply was surplus, posing a problem of
esterification for biodiesel production, saponification for manufacturing disposal, since crude glycerol is of little economic value and use due to
of soap, hydrolysis for the fatty acid production, and microbial fer- the presence of inorganic salts and other impurities [11,12]. For ex-
mentation. Table 1 shows the compositions of crude glycerol produced ample, the crude glycerol price in 2014 was US$240/tonne while for
by different processes [8,9]. pure glycerol (USP grade), it was US$900/tonne [12]. Purification of
Homogeneous catalytic transesterification is one of the major glycerol increases its economic and applicable value and makes bio-
sources for most crude glycerol production [7]. The transesterification diesel production more viable [7]. Pure glycerol is a renewable com-
reaction is shown in Fig. 1 [9]. The chemical composition of crude modity and feedstock for biorefineries (food, chemical, and pharma-
glycerol is affected by a number of factors such as – the type of catalyst ceutical industries) and used in the production of fuels or fuel additives
used in biodiesel production, the efficiency of the transesterification [6,7].


Corresponding author.
E-mail address: Ajay.Dalai@usask.ca (A.K. Dalai).

https://doi.org/10.1016/j.fuproc.2018.05.023
Received 31 December 2017; Received in revised form 15 May 2018; Accepted 16 May 2018
0378-3820/ © 2018 Elsevier B.V. All rights reserved.
C.G. Chol et al. Fuel Processing Technology 178 (2018) 78–87

Table 1 challenges during membrane filtration [15,16]. Ceramic membranes


Different compositions of crude glycerol produced by different processes [8,9] are emerging as a good alternative to conventional polymeric mem-
Component Transesterification (%) Saponification (%) Hydrolysis (%) branes because of their great thermal, chemical, and mechanical sta-
bility properties, higher permeability, and easy cleaning process
Glycerol 30–60 83–84 88–90 [17,23]. These characteristics are important for membrane because of
Ash 10–19 8.5–9.5 0.7–1.0
the nature of the crude glycerol solution treated by the membrane. In
Water ≤10 6–7 8–9
MONG (matter ≤40 3–4 0.7–1.0
addition, robust membrane is also an absolute requirement for in-
organic non- dustrial scale up of the process, which has been explored in the techno-
glycerol) economic analysis. Among the finishing steps, activated carbon ad-
Trimethylene 1 0.1 0.2 sorption removes colour in crude glycerol and removes some free fatty
glycol
acids (lauric and myristic acids) by adsorbing them along with other
molecular compounds [6,7,17].
Glycerol purity is defined by its grade – 95% purity glycerol is There are several studies in literature for purification of crude gly-
classified as technical grade; 96–99% purity glycerol as USP grade cerol employing one or more physicochemical treatment steps such as
glycerol; and 99.7% purity glycerol as Kosher glycerol [12]. Technical neutralization, acidification, filtration, extraction, and adsorption
grade glycerol is used as a building block in chemicals but is not used in [7,14,18–21], ion-exchange [20,22], and vacuum distillation [23].
food production or drug formulation. USP (United States Pharmaco- However, there are handful of studies for glycerol purification by
peia) grade glycerol is employable in production of foods and phar- membrane filtration, and to the best of our knowledge, no study has
maceuticals while Kosher glycerol is suitable for use in Kosher foods been done for the crude glycerol purification using tubular membrane
production [12]. in semi-continuous mode.
For the purification of crude glycerol, techniques such as – dis- In a previous study, it was described that the combined physico-
tillation, ion exchange, and sequential physicochemical treatments, chemical treatment and dead-end membrane filtration was successful in
including saponification, acidification, phase separation, solvent ex- purifying crude glycerol to a purity of > 95% [7]. As it was a dead-end
traction, neutralization, and activated carbon or yeast adsorption are filtration, the system was limited in its volume handling capability and
employed [6,7,12,13]. Acidification of crude glycerol at low pH can the membrane needed to be replaced or required frequent cleaning.
increase glycerol content and reduce the amount of ash in recovered Owing to these shortcomings of the system, in the present work, semi-
glycerol [14]. It might however, lead to higher MONG content in the continuous membrane filtration was coupled with the physicochemical
resultant enriched glycerol due to formation of free fatty acids (FFAs) treatment to evaluate the efficiency of the process and to see the effect
from acidification of saponified fatty acids (SFA) by mineral acid (H+) of the parameters on the purification process. In addition, as glycerol
[7]. purification involves multiple unit operations, a detailed techno-eco-
For purification of crude glycerol, vacuum distillation, ion ex- nomic analysis of the process was carried out to determine economic
change, membrane separation, and activated carbon adsorption are feasibility and the profitability of the process.
regarded as deep refining technologies [6]. Out of these, vacuum dis-
tillation and ion exchange are energy- and cost-intensive processes for 2. Materials and methods
purifying glycerol resulting in high energy input requirement [6].
Furthermore, the salt content in crude glycerol originating from 2.1. Material
homogeneous catalytic processes usually ranges from 5 to 7%, making
the conventional purification techniques costly [10]. Crude glycerol samples were supplied by Milligan Biofuels, Foam
Membrane filtration is an emerging technology in glycerol pur- Lake, SK, Canada. ACS grade glycerol (99.5 wt% purity) was purchased
ification and is highly appealing because of its simple operability, low from Fisher scientific, Canada. Ceramic tubular membranes (diameter
energy requirement, and therefore low cost and good purification 4.7 cm, area 443 cm2, molecular weight cut-off 5 kDa) were composed
performance (> 90 wt% glycerol output) and environmental friendli- of ZiO2-TiO2 with TiO2 support. The membranes and tubular membrane
ness [15,16]. Membrane filtration can be incorporated as an assisting holder was purchased from Tami Industries, France. All other chemicals
technology to other purification processes such as ion-exchange, dis- were of analytical grade unless otherwise stated.
tillation, or adsorption after an initial treatment which removes sig-
nificant level of impurities considering the limitations of membrane
filtration. This will increase the purification efficiency and product 2.2. Glycerol purification process
consistency, make process operation easier, and save the plant opera-
tional time. The overall process for glycerol purification is summarized in Fig. 2
Membrane performance is dependent on several factors including and the steps consist of saponification, acidification, overnight phase
membrane material and porosity, feed temperature, pH, trans-mem- separation, and neutralization, followed by membrane filtration under
brane pressure, tangential velocity, and interaction between membrane varying temperature, pressure, and flow rate conditions. Solvent and
and feed [15,16]. Caking of membrane material and fouling remain real water evaporation and activated charcoal treatment are the final fin-
ishing steps.

Fig. 1. Transesterification reaction to produce fatty acid methyl ester (FAME) and glycerol.

79
C.G. Chol et al. Fuel Processing Technology 178 (2018) 78–87

Fig. 2. The steps in crude glycerol purification process.

2.2.1. First step: physicochemical treatment Table 2


Physicochemical treatment includes a sequential process of sapo- Central composite design (CCD) of experiment.
nification, acidification, phase separation and extraction as described Experimental run Temp (°C) Pressure Flow Rate Response (glycerol
previously [7]. The slightly modified process is as follows: Crude gly- (kPa) (mL/min) purity wt%)
cerol (glycerol content 40 wt%) was mixed with methanol to obtain
1 40 862 125 87.3
working glycerol concentration of 20 wt% to reduce glycerol viscosity
2 25 1380 50 88.2
and improve the flow behavior. It was then heated to 60 °C followed by 3 40 862 125 88
saponification to pH 12.0 using 12.5 M KOH with constant stirring for 4 25 345 50 83.8
30 min to convert free fatty acid (FFA) to soap (saponified fatty acids). 5 50 345 200 89.9
Acidification was then done with 1 M concentrated HCl (98%) to pH 1.0 6 50 345 50 85.8
7 40 862 125 89
and stirred for 30 min at room temperature (25 °C) before overnight
8 40 862 125 88.1
separation via gravity sedimentation in separatory funnels. Acidified 9 50 1380 200 79.4
glycerol separates into two distinct layers: top layer of FFA and gly- 10 50 1380 50 91
cerol-rich bottom layer and a tiny layer of organic salts at the bottom of 11 40 862 125 88.1
12 25 1380 200 83.4
the glycerol phase. The upper layer was slowly decanted off and the
13 25 345 200 82.3
bottom inorganic salt layer run off to leave the middle glycerol-rich 14 40 862 125 88.1
layer to be extracted with volumes of petroleum ether and toluene
solvents. The separated glycerol was extracted through solvent extrac-
tion using equal volume each of petroleum ether and toluene and then LabVIEW software via USB. The flow pattern in feed tanks was such
with half volume each of petroleum ether and toluene in a second round that the purified liquid can either go to the same tank or different tank
to remove residual FFAs. The resultant glycerol was neutralized to after membrane filtration as both the tanks were independently con-
pH 7.0 with 12.5 M KOH to obtain the treated glycerol, an enriched nected to the pump. To study membrane filtration of treated feed,
form of glycerol, which was used in the membrane filtration experi- transmembrane temperature, pressure, and feed flow rate were varied
ments. in the range of 25–50 °C, and 350–1380 kPa, and 50–200 mL/min, re-
spectively. The details of experimental design are presented in Table 3.
Treated glycerol from the physicochemical step (Section 2.2.1) was
2.2.2. Second step: membrane filtration in semi-continuous mode filled in the feed tank up to over the half-way mark. Then the tank(s)
Treated glycerol was further enriched through membrane filtration and membrane module were heated, and the module was pressurized to
using a 5 kDa ultrafiltration tubular membrane at varying temperature, the desired pressure before the valves were opened to allow feed flow to
flow rate, and pressure to obtain the final enriched glycerol. the membrane module for filtration. About 15 mL filtrate was collected
Temperature was varied from 25 to 50 °C, flow rate from 50 to 200 mL/ at each set of conditions with the process lasting about an hour. After
min, and pressure from 345 to 1380 kPa. The experimental combina- the process, membrane was thoroughly flushed with methanol to clean
tions determined employing a central composite design CCD, (Table 2). the foulants and to prepare for the next round of experiments.
The schematic for the membrane filtration set-up is shown in Fig. 3(a)
and the actual semi-continuous apparatus in shown in Fig. 3(b).
The apparatus consisted of two feed tanks connected with the 2.2.3. Finishing steps: solvent vaporization and activated carbon treatment
membrane module (cross flow filtration) through a feed pump. The third stage of glycerol purification involved adsorption of en-
Temperature and pressure inside the feed tank and membrane module riched glycerol with activated carbon as one of the finishing steps in
were monitored constantly and controlled by thermocouples (K type, glycerol purification to remove colour and residual FFA. A ratio 1:10
Omega) and pressure transducers (Honeywell) connected to tempera- was used as per previous study [7]. The final stage was water and
ture and pressure monitors which were connected to PC using interface solvent (methanol, residual petroleum ether, and toluene) evaporation

80
C.G. Chol et al. Fuel Processing Technology 178 (2018) 78–87

helium as the carrier. FFA, mono-, di-, and triglycerides, and esters
concentration in crude, treated, and membrane filtered glycerol were
determined by HPLC using tetrahydrofuran as mobile phase at 1 mL/
min flow rate using a refractive index detector. Water content was
determined using an automated Karl Fisher coulometric titrator (Metler
Toledo DL32) using methanol for dilution as the titrator was sensitive to
a water content of 5 wt% maximum. Crude, treated, and membrane
enriched glycerol were characterized by Fourier-Transform infrared
(FTIR) spectrometer (Bruker Vertex 70, MA, USA) with ATR module.
Each spectrum was the average of 16 co-addition scans with a total scan
time of 15 s in the range of 400–4000 cm−1 at 4 cm−1 resolution.

2.4. Techno-economic feasibility analysis

For the techno-economic analysis, glycerol purification process was


modeled using Aspen Hysys software and the associated capital and
operating costs estimated from the simulated process flow diagram,
PFD (Fig. 4). The process in this simulation follows closely the experi-
mental process that has been described in Section 2.2. Equipment are
sized by the software based on the process material and energy balances
and the operating conditions around an equipment. Because of the
polar nature of most components in this process, non-random two li-
quid (NRTL) model was chosen as the fluid simulation package that can
accurately simulate the process. Another equally good option is uni-
versal quasi-chemical (UNIQUAC) model. NRTL was used together with
the UNIFAC LLE (liquid-liquid equilibrium) to estimate the missing
coefficients, particularly for missing components that were created and
added to the Hysys library [4,24].
Among the missing components created were KOH, KCl, potassium
oleate (K-Oleate), glycerol carbonate, and solid catalyst Amberlyst-15,
used in ketalization of glycerol. Other chemicals and equipment had
their equivalents used to represent them in the simulation. Oleic acid
was used to represent free fatty acid component of the crude glycerol, n-
pentane and n-hexane to represent petroleum ether by being combined
in equal volumes, diethyl ether was used to represent dibutyltin oxide
catalyst for glycerol carboxylation, and paraldehyde represented solk-
Fig. 3. (a) Schematic of the membrane filtration apparatus for glycerol pur- etal as they have the same chemical formula and comparable physical
ification. and chemical properties. A splitter was used to represent a membrane
(b) Semi-continuous membrane filtration apparatus. separator with chemical components in the feed being separated using
split ratios from literature [16,25]. Pressure drop was assumed to be
Table 3 zero for all equipment mostly because most of the process steps are
Enrichment of glycerol after each stage of physicochemical treatment. taking place at atmospheric pressure. Jacketed reactors were chosen,
Stage of physic-chemical Glycerola % (w/ Methanol % Water %
where heating or cooling was required. Where acids were involved,
treatment w) (w/w) (w/w) process equipment and their piping systems were made of stainless steel
(ss). The rest of the equipment and piping system were made of regular
Crude glycerol 40.0 30.0 5.0 carbon steel (cs), unless stated otherwise.
Saponification-acidification 75.4 ± 2.3 0 2.5–4.0
Economic evaluation refers to the evaluation of capital and oper-
and Overnight separation
Petroleum ether extraction 82.5 ± 1.8 0 2.5–4.0 ating costs associated with the construction and operation of a chemical
Toluene extraction 86.4 ± 1.9 0 2.5–4.0 process [26,27]. The criteria for economic evaluation in this work is
Neutralizationb 83.5 ± 2.4 0 4.0–8.0 based on capital cost, manufacturing cost, and project profitability as
a
measured by net present value or worth (NPV or NPW) of the project.
Average of glycerol percentage for > 40 batches of treated glycerol after
Three different hypothetical scenarios were studied in this work. In
solvent evaporation.
b scenario 1, all the purified glycerol is to be sold directly as it is. In
Neutralization slightly reduces the glycerol purity as water is generated in
scenario 2, 50% of the purified glycerol is to be sold directly, while 50%
acid-base neutralization reaction.
is to be converted to value-added fuel additives – 25% to solketal, and
the other 25% to glycerol carbonate. In scenario 3, all purified glycerol
from treated and enriched glycerol to obtain a final volume of 3–4 mL of
is to be converted to value-added fuel additives with 50% being con-
glycerol sample. Solvent and water evaporation was done by heating
verted to solketal and the other 50% into glycerol carbonate.
glycerol to temperatures above 100 °C with constant stirring.
3. Results and discussions
2.3. Analytical methods
3.1. Effectiveness of the physicochemical treatment
The concentration of glycerol and methanol (in wt%) were de-
termined by a gas chromatograph (GC Agilent 7890 series) using The crude glycerol used in this study had the same composition and
StabilWax column (dimensions 30 m × 250 μm × 0.5 μm; Restek Corp., physical characteristics as described previously [7]. The mechanism of
USA) at 250 °C with an FID detector at 300 °C, nitrogen at 159 kPa, and physicochemical treatment was also described in the previous study

81
C.G. Chol et al. Fuel Processing Technology 178 (2018) 78–87

Fig. 4. Process flow diagram for techno-economic analysis.


Chemical abbreviation [MeOH – methanol; Light Liq – light liquids; PE_solvent – petroleum ether solvent; PE_Rec – petroleum ether recovered; Liq_Waste – liquid
waste; KOH and KOH-2 – potassium hydroxide; Cat-1 and Cat-2 – catalysts 1 and 2; Acet – acetone; Liq_Waste – liquid waste; Rec_MeOH – recovered methanol;
Excess_CO2 – excess carbon dioxide].
Equipment abbreviation [MIX-100 and MIX-101 – mixers; R-100, R-101, R-102, R-103 and R-104 – conversion reactors for saponification, acidification, neu-
tralization, ketalization and carboxylation, respectively; V-100 – gravity separator; E-100, E-101, E-102, E-103 and E-104 – heat exchangers; V-101 – liquid-liquid
separator; V-102, V-103, V-104 – flash separators; L-100, L-101, L-102 – pumps; VLV-100 – feed control valve; M-100 – membrane separator; T-100 and T-101 – Tee-
splitters; RCY-1 – recycle symbol; and Q-100, Q-105, Q-110, Q-115, Q-120, Q125, Q-130, Q-135, Q-140, Q-145, Q-150, Q-155, Q-160, Q-165, Q-170 – cooling or
heating energy sources].

Table 4 [28]. Both the solvents were employed sequentially for two rounds of
Composition of crude, treated enriched glycerol. extraction and both were equally effective for enhancing the glycerol
Component Crude glycerol Treated glycerola Enriched glycerolb purity by removing FFA from the crude glycerol. The last step of the
(wt%) (wt%) (wt%) treatment, neutralization, led to slight decrease in glycerol purity due to
the formation of water and salt in the acid-alkali neutralization reaction
Glycerol 40.00 86.40 93.70
[7]. A maximum of 86.4 wt% glycerol purity was obtained after phy-
Methanol 30.00 0 0
Water 5.00 3.00 3.00
sicochemical treatment which is slightly lower than that obtained
Triglycerides 0.13 0.03 0.03 (88.6 wt%) in our previous study [7], than the results of Manosak et al.
Diglycerides 0.68 0.03 0.02 [21] who reported 95.74 wt% and similar to that of Xiao et al. (> 94 wt
Monoglycerides 1.90 0.25 0.12 %) [5]. In the present work, the glycerol yield reported was the average
Free fatty acids (FFA) 13.20 0 0
glycerol yield from > 40 batches, hence slight variation in yield was
Esters 10.00 1.52 0
Ash 4.90 2.80 1.2 obtained. Furthermore, Table 4 presents the overall composition of
Total 105.80 93.93 98.07 crude, treated, and enriched glycerol. Treated glycerol refers to the
purified glycerol obtained after the physicochemical treatment, while
a
After physicochemical treatment and solvent evaporation. enriched glycerol is obtained after membrane filtration and activated
b
Enriched glycerol is glycerol after membrane filtration, activated charcoal carbon adsorption of treated glycerol. In both the cases, solvent and
adsorption, and solvent and partial water evaporation. partial water evaporation was carried out before analysis. Enrichment
in glycerol purity was obtained after physicochemical treatment and
[7]. Table 3 shows the enrichment in glycerol purity after each stage of membrane filtration. 100% removal of FFA and methanol, and ~85%
physicochemical treatment which is the average of > 40 batches. The removal of esters was obtained, while there was significant reduction in
first combined step (saponification, acidification, and phase separation) ash, mono-, di-, and triglycerides content after the physicochemical
enhanced the glycerol purity to > 75 wt% from the initial 40 wt% treatment. The total composition of crude and treated glycerol is not
glycerol which is ~1.9 fold enhancement in glycerol purity. For the 100%, the reason can be the presence of minor unidentified colloidal
solvent extraction, two different solvents – petroleum ether and toluene particles formed during the treatment (in case of treated glycerol) or
were employed as the solvent type and ratio affects the glycerol yield biodiesel production (for crude glycerol).

82
C.G. Chol et al. Fuel Processing Technology 178 (2018) 78–87

1.2
Table 5
Representative results of the membrane filtration of treated feed.

1
Temperature (°C) Flow rate (mL/min) Pressure (kPa) Glycerol purity (wt%)

25 50 345 83.8
50 700 87.8
0.8 50 1380 88.2
Transmittance (a u)

40 50 345 92.4
50 700 91.0
0.6 100 700 92.2
125 862 88.1
50 50 345 85.8
50 700 93.7
0.4
50 1380 91.0
ACS Glycerol 100 345 86.6
Purified Glycerol 100 700 87.3
0.2 100 1380 86.8
Treated Glycerol
200 345 89.9
Crude Glycerol

0
0 500 1000 1500 2000 2500 3000 3500 4000 4500
can be observed that the highest glycerol purity results of 93.7 wt%
Wavenumber (cm-1)
were obtained at 50 °C temperature, 50 mL/min flow rate and 700 kPa
Fig. 5. FTIR spectra for crude, enriched (treated step 1 and purified step 2) and pressure. The next best glycerol purity results are 92.4 and 92.2 wt%
ACS grade glycerol. glycerol purity obtained at 40 °C, 50 mL/min and 345 kPa; and 40 °C,
100 mL/min and 700 kPa, respectively. These glycerol purities are very
3.2. Effect of semi-continuous membrane filtration close to technical grade. Glycerol purity after the initial sequential
physicochemical treatment was 82.5 wt%. Filtration at 25 °C (Table 5)
The best results for membrane filtered enriched glycerol are pre- using a 5 kDa ultrafiltration membrane improved glycerol purity
sented in Table 4. As compared to treated glycerol, further enrichment 93.7 wt% highest at 50 °C, 50 mL/min flow rate and 700 kPa pressure.
in glycerol purity was obtained after membrane filtration and a max- Highest glycerol purities from 92 to ~94 wt% were obtained at the flow
imum of ~94 wt% glycerol purity was obtained which is very close to rate and pressure ranges of 50–100 mL/min and 345–700 kPa which
technical grade (95%). Membrane treatment was successful in re- imply that glycerol purity during semi-continuous membrane filtration
moving colloidal particles, oil droplets, and some other components in this work is optimized at a temperature, pressure, and flow rate of
such as salt deposits, which were not completely removed during the 50 °C, 700 kPa, and 50 mL/min, respectively.
physicochemical treatment. Fig. 4 presents the comparative FTIR Organic molecules and colloids of larger molecular size (MWCO) are
spectra of crude, physicochemical treated, membrane enriched, and rejected by the ultrafiltration membrane of 5 kDa MWCO due to their
ACS grade (for reference) glycerol samples. The FTIR spectra for treated large molecular size and are thus unable to sieve through membrane
and membrane enriched glycerol (presented in Fig. 5) were very similar pores into the permeate [15,16,25]. This results in further enhancement
and were overlapping with that of ACS grade glycerol. The spectra for in glycerol purity. Relatively low glycerol purity results are obtained at
crude glycerol were identical to that of our previous study [7] as the 25 °C compared to 40 and 50 °C and this can be attributed to higher
same batch of crude glycerol was utilized for the present study. There glycerol viscosity at room temperature (~1.2 cP vs 0.8 cP from 40 to
were notable changes in the spectra for crude glycerol as compared to 50 °C) [7]. This affects the flow properties of glycerol and the ability to
other three spectra for pure or purified glycerol. At around 1550 and flow through the membrane and be separated from the impurities [30].
1740 cm−1, strong peaks were obtained for crude glycerol which de- Glycerol purification by the combination of physicochemical treatment
note eC]O bond indicating the presence of carboxylic acid or fatty and membrane filtration works best at lower glycerol viscosity as this
acid esters in crude glycerol which were almost negligible in purified result in more solute coming into contact with membrane surface and
glycerol spectra [7,29]. Intense peak at 2900 cm−1 indicate un- increasing the chances of the smaller components being filtered through
saturated C]C compound (s) in crude glycerol which was absent or the pores by sieving, as per the primary filtration mechanism in Darcy's
below detection limit in purified or ACS glycerol spectra [14]. law governing ultrafiltration [16,25]. With improved feed flow due to
Membrane performance in enhancing glycerol purity was studied in low glycerol dynamic viscosity as a result of elevated temperatures,
terms of the effects of parameters – temperature, pressure, and feed bigger particles, mostly organics, are also easily rejected by membrane
flow rate and the parameter-parameter interactions effects on glycerol sieves (pores) of low MWCO (5 kDa) at temperatures from 40 to 50 °C.
purity. Membrane filtration was studied using a 5 kDa UF ceramic The result is more glycerol, water and solvent (methanol) sieving
tubular membrane. The membrane flux was not studied as the flux was through the membrane and more of the colloids and organic macro-
high and stayed almost the same at all the experimental conditions. molecules being excluded [15,16,25]. Methanol and water are later
Furthermore, effect of solution pH was also not studied as the pH op- vaporized hence increased glycerol purity since most of the larger or-
eration range of ceramic membrane was 1–14. Temperature, pressure, ganic materials and colloidal particles have already been rejected by
and flow rate were varied in the range of 25–50 °C, 345–1380 kPa, and the membrane at higher temperatures. Therefore, the highest glycerol
50–200 mL/min, respectively. 50 °C was chosen as the highest tem- purity results (92.2–93.7 wt%) were obtained at the high temperatures
perature to avoid methanol evaporation (boiling point 64.7 °C). of 50 °C.
Furthermore, to avoid membrane pore rupture, pressure > 1380 kPa Pressure is used as the driving force in ultrafiltration and it creates
and flow rate > 200 mL/min were not employed. The underlying as- the required force to enhance the sieving through membrane pores of
sumptions were that impurities with MWCO > 5 kDa like oil droplets molecules smaller than 5 kDa molecular weight [15,16,25]. In this
and colloidal particles would be retained by the membrane while gly- work, increasing pressure increased glycerol purity up to pressure va-
cerol (92.1 g/mol) would filter through membrane pores as filtration is lues above 700 kPa. It is to be noted that, not all molecules smaller than
based on the MWCO of a material. the MWCO are able to sieve through as some get blocked by adsorbed
From the best results obtained from physicochemical treatment and particles on membrane surfaces and do not come into contact with the
membrane filtration at different optimization conditions (Table 5), it membrane surface [15,16,25]. Looking at the results in terms of flow

83
C.G. Chol et al. Fuel Processing Technology 178 (2018) 78–87

Fig. 6. Response surface plots showing the effects of parameters on glycerol purity: (a) effects of temperature (°C) and pressure (kPa) on glycerol purity at a flow rate
of 125 mL/min; (b) effects of temperature (°C) and flow rate (ml/min) on glycerol purity at a pressure of 862.5 kPa; and (c) effects of flow rate (ml/min) and pressure
(kPa) on glycerol purity at a temperature of 40 °C.

rate, increasing the flow rate generally decreases the purity of the re- pressure than at lower temperature and lower pressure. The contour for
sultant glycerol. higher temperature and higher pressure is especially steeper at higher
pressure areas even when the temperature is not particularly as high as
3.2.1. Response surface methodology (RSM) 35 °C. This can be interpreted to mean higher effect of pressure up to
With regards to the combined effects of the three parameters of 1173 kPa on purity at even temperatures under 35 °C.
temperature, pressure and flow rate and the possible effects of their Temperature and flow rate response surface plots (Fig. 6(b)) give
interaction with each other on glycerol purity, response surface plots the best results at high temperature (up to 50 °C) and low flow rate
were constructed from the response (glycerol purity or yield) of every (50–100 mL/min). Looking at the contours at the bottom of the re-
two parameters and the results are presented in Fig. 6(a)–(c). The in- sponse surface plot, which are initially parallel and running diagonally
teraction between pressure and temperature (Fig. 6(a)) seems to give from right to the left, it can be seen that glycerol purity is increasing
the best results at higher temperatures from 40 to 50 °C and lower diagonally towards the left side area of low flow rate (< 80 mL/min)
pressure (from 345 to 700 kPa) as this is the area of highest resultant and higher temperature (> 45 to 50 °C). It can therefore be interpreted
glycerol purity. The increase in purity is thus in the direction of red to mean glycerol purity is increasing with increasing temperature and
colour from the barely blue edge through the largely green middle area. decreasing flow rate. The diagonal lines are no longer parallel towards
Looking at the contours at the bottom of the response surface plot in the left end when they become steeper, meaning higher rate of increase
Fig. 6(a) in the yellow area and their gradients at a point and over small in purity with decreasing flow rate (80 to 50 mL/min) and increasing
parts of a line, it can be seen that contour for lower temperature and temperature (45 to 50 °C).
pressure (30 °C and 759 kPa) were more gently sloping than those at The combined effect of pressure and flow rate (Fig. 6(c)) has its best
higher temperature and pressure (35 °C and 1173 kPa), which are glycerol purity results to the left side in the region of high pressure
steeper at starting points. This can be interpreted to mean a higher rate (1173 to 1380 kPa) and low flow rate (50 mL/min). Glycerol purity is
of increase in response (purity) at higher temperature and higher increasing from the green region to the red region. With regards to

84
C.G. Chol et al. Fuel Processing Technology 178 (2018) 78–87

contours at the bottom of the pressure and flow rate response surface Purity (%) = 71.5 + 0.3 × T ( °C ) + 0.016 × P (kPa)
plot, especially the inner-most contour, it can be seen that the contour is
+ 0.036 × FR (mL/ min) − 00209 × T × P
steeper at the starting point on the right side in region of high flow rate
− 0.00016 × T × FR − 0.0000612 × P × FR (2)
(200 to 140 mL/min) and low pressure (345 to 552 psi) before it be-
comes more gentle at a latter stage. This signifies that the rate of in- This equation can be used to make predictions about the response
crease in glycerol purity at high but decreasing flow rate (200 to (% purity) in terms of given levels of each factor. The levels should be
140 mL/min) and low pressure (345 to 552 psi) is initially very high specified in the original units for each factor. This equation can how-
before the rate of increase in purity steadies at a slower rate as can be ever not be used to determine the relative impact of each factor as the
seen from the counter becoming a near straight line. coefficients are scaled to accommodate the units for each factor and the
Overall in terms of the effects of the interactions of the three intercept is not at the center of the design space.
parameters on glycerol purity, the best glycerol purity results were
obtained at high values for temperature (50 °C), low flow rate (50-mL/
3.3. Activated charcoal adsorption
min), and pressure in the moderate to higher range (~700 kPa).

Activated carbon adsorption was employed as a finishing step to


3.2.2. Statistical analysis further improve glycerol purity and to remove the colour [6]. Organic
To describe the effects of these three parameters (temperature, compounds smaller than the pore sizes of activated carbon get absorbed
pressure, and feed flow rate) on glycerol purity, a two-factor interaction between its pores [6]. Treatment with activated carbon generally re-
(2FI) model was found to be the most appropriate to describe the re- moves residual fatty acids and colour [6]. However, as in the present
lationship. On the basis of sequential model sum of squares (Type I) case, all the FFA was removed in physicochemical treatment, with ac-
from statistical analysis, a two-factor interaction (2FI) vs mean model tivated charcoal treatment, no significant improvement in glycerol
was found to be most appropriate because its additional terms were purity was obtained. The only improvement was significant reduction
significant, and the model was not aliased. A 2FI was also found to be in colour, as the enriched glycerol was colourless and resembled com-
the most appropriate by the Lack of Fit Tests method as it had insig- mercial grade glycerol in appearance [6].
nificant lack of fit (Prob > f = 0.0002). On the basis of the model Glycerol purity above 94% was not obtained in the present work,
summary of statistics, 2FI model was also found to be the most ap- possibly due to comparatively lesser contact time between the solution
propriate to describe this optimization process as it maximized Adjusted and the membrane surface as compared to batch mode. Very small
R-Squared and predicted R-Squared. colloidal particles (in the size range of 1–2 nm) can form during the
biodiesel production process, some of these particles may pass onto the
3.2.2.1. Analysis of variable (ANOVA) for 2FI model. A value of glycerol phase and may not be filtered by microfiltration membrane.
Prob > F < 0.05 indicates that the model terms are significant. BC is Nanofiltration membrane can possibly produce a better yield and
thus a significant model term in this case. “Adeq Precision” measures quality of purified glycerol by removing smaller colloidal particles not
signal to noise ratio with a ratio > 4 being desirable. The current removed by microfiltration, which can be a topic of future studies.
ANOVA gives “Adeq Precision” of 5.007 indicating an adequate signal.
This model can thus be used to navigate the design space.
3.4. Techno-economic feasibility study and cash flow analysis
Based on the table of coefficients, standard error and 95% confident
intervals, the final equation in terms of coded factors is given as Eq. (1):
Three different scenarios were considered from glycerol purifica-
Glycerol purity (%) = 86.49 + 1.25 × A + 0.025 × B − 1.73 × C tion. In scenario 1, purified glycerol is sold as it is with no value-added
products made from it. Scenario 2 involved selling 50% of the purified
− 1.35 × AB − 0.15 × AC − 2.37 × BC (1)
glycerol as it is, while converting 50% of purified glycerol into value-
Eq. (1) can be used to predict the response (glycerol purity) for added products (solketal and glycerol carbonate; 1:1 ratio). Scenario 3
given levels of each factor. By default, the higher levels of the factors involved converting all the purified glycerol produced into value-added
are coded as +1 and the low levels of the factors are coded as −1. The products – solketal and glycerol carbonate (50% each). The economic
coded equation is useful for identifying the relativity impact of the analysis in this work is based on these three scenarios, which were
factors by comparing the factor coefficients. compared on the basis of per unit cost of purifying glycerol and on per
Eq. (2) describes the response (glycerol purity) in terms of the three unit revenue generation ($US/kg of crude glycerol). The details of cost
parameters of temperature, pressure, and flow rate that are being op- determinants and comparative analysis of all three scenarios is pre-
timized. sented as supplementary section. Table 6 presents the cost and revenue

Table 6
Comparative manufacturing costs and revenues tables for the three glycerol production scenarios.
Scenario 1 Scenario 2 Scenario 3

Item Total annual cost Annual cost and Total annual cost Annual cost and Total annual cost Annual cost and
($US) revenue ($US /kg of ($US) revenue ($US /kg of ($US) revenue ($US /kg of
crude glycerol) crude glycerol) crude glycerol)

Direct manufacturing costs 664,727.90 4.75 3,720,230.13 27.70 6,133,283.95 45.66


Utilities 51,161.88 0.38 162,435.96 1.21 272,174.33 2.03
Indirect manufacturing costs 41,191.16 0.31 63,584.18 0.47 63,584.18 0.47
Maintenance & Repair (6% of 70,613.42 0.53 109,001.44 0.81 109,001.44 0.81
fixed capital cost)
Operating supplies (15% of 10,592.01 0.08 16,350.22 0.12 16,350.22 0.12
maintenance & repair)
Depreciation (10% of capital 117,689.03 0.88 181,669.07 1.35 181,669.07 1.35
cost)
Total manufacturing expenses 1,073,664.42 7.12 4,253,270.99 31.67 6,776,063.19 50.45
Total annual revenues 347,150 2.58 5,353,259.13 39.85 10,793,485.08 80.36

85
C.G. Chol et al. Fuel Processing Technology 178 (2018) 78–87

20 The net present value (NPV) of the project is approximately $6 million


Cumulative discounted Cash flow (in million dollors)

and $3.65 million respectively at 10% and 15% discounting interest


rates since that is the final discounted cumulative cash flow amount at
15 project conclusion.
Since this is an investment of over $2 million in total capital input
0% 10% 15% and expected discounted cumulative cash flow is $6 million at 10%
10
discounting rate and $3.65 million at 15% discounting rate in net
present value (NPV), this is a project worth undertaking as it promises
to be profitable within 10 years after start-up. Per unit, annually, crude
5
glycerol purification in this scenario costs $50.45 while it returns
$80.36.
These are decent returns for a biodiesel plant as they will help offset
0
production cost and increase profit margins. These returns indicate that
it would be worth it investing in crude glycerol purification and va-
lorization technology in addition to an existing biodiesel production
-5
0 2 4 6 8 10 12 14 plant.
Year A similar study for incorporation of glycerol purification plant in
Fig. 7. Cash flow analysis at different discounting rates. biodiesel production facility also suggested that it is beneficial to carry
out glycerol purification after biodiesel production which will lead to
revenue generation and reduction in wastewater treatment cost [31]. A
generation comparison of these three scenarios. As it can be seen,
study carried out by Hunsom et al. for the economic aspect of glycerol
scenario 3 generated highest revenue compared to the other scenarios
purification estimated the cost of glycerol purification using adsorption
and also has higher profit margin as compared to scenario 2. Scenario 1
and solvent extraction which was about $33/kg of crude glycerol,
is outright unprofitable. It is also beneficial to opt for scenario 3 in
which is much higher compared to obtained in the present work (about
regards to possible overproduction and environmental hazards of crude
$7/kg in case of scenario 1) [28].
glycerol as it reduces glycerol availability in the market by converting it
into fuel additives. This reduces overall crude and purified glycerol
4. Conclusions
supply, as all of the purified glycerol is being utilized for value added
chemical production.
Crude glycerol was purified by combined physicochemical and
semi-continuous tubular membrane filtration in this work.
3.4.1. Cash flow and profitability analysis Physicochemical treatment comprising saponification, acidification,
In the cash flow analysis part of economic feasibility analysis, only phase separation, and solvent extraction, removed most of the im-
the best case scenario (scenario 3) was evaluated for profitability using purities such as FFA, glycerides, esters, and inorganics (as ash). The
a plot of cumulative cash flow against years of project life (Fig. 6). Two physicochemical treatment enhanced the glycerol purity to > 86 wt%.
discounting interest rates (10% and 15%) were used in cash flow ana- Further, due to the membrane filtration of treated feed, the best gly-
lysis. cerol purity (just under technical grade at 93.7 wt% glycerol) were
These are the assumptions that went into generating the cumulative obtained at 50 °C temperature, 50 mL/min flow rate, and 700 kPa
cash flow chart in Fig. 7, which is for scenario 3 as mentioned earlier. A pressure. Ultrafiltration membrane rejected large organics and colloidal
fixed capital of $1.83 million and working capital of $0.27 million (15% particles from crude glycerol by sieving mechanism as per Darcy law,
of fixed capital) were invested in the project over a three-year period the operating law for ultrafiltration membrane filtration. Pressure being
during which there was no production. That is a total of $2 million in the driving force in ultrafiltration, contributes to improving the overall
total invested capital. Fixed capital was invested at a rate of $0.61 glycerol purity by providing the force necessary to push the solute
million per year for the first three years with working capital also being particles of lower MWCO through the membrane. The combined effects
invested in Year 3, making a total investment of $0.88 million invest- of factors on the highest achievable glycerol purity are best described
ment in Year 3 ($0.61 million + $0.27 million). Working capital ($0.27 by a two-factor interaction (2FI) model. Adsorption using activated
million) was to be recovered at the end of the project (Year 13 of total carbon as a finishing step improves glycerol purity further by adsorbing
project life or year 10 of operation). Fixed capital was depreciated some organic macromolecules and remove colour. The techno-eco-
linearly at 10% over 10 years of operation (Years 4 to 13 inclusive) with nomic analysis showed that the process is economically feasible when
zero salvage value at the end of project life. No allowances were con- at least 50% of the purified glycerol is converted to value-added fuel
sidered and federal income tax was assumed to be 35%. During the first additives such as solketal and glycerol carbonate but more profitable
two years of project start-up (Years 4 and 5), revenues were assumed to when all glycerol is valorized and sold as solketal and glycerol carbo-
be 80% of the projected amount ($8.63 million) while expenses were nate. A capital investment of $2.1 million over three years on a project
assumed to be 120% ($8.14 million) of the projected value. The normal that will be operational for 10 years after start-up results in $6.0 million
projected revenues and expenses for later years (Years 6 through 13) or $3.65 million in net present value (NPV) at a 10% or 15% dis-
were $10.73 million and $6.78 million, respectively. counting rate in accordance with scenario 3, where all the purified
From Fig. 7, it can be observed that the project payback period is glycerol was converted to fuel additives – solketal and glycerol carbo-
just before year 6, which is how long it would take to recover the un- nate. The unit annual general cost of scenario 3 per kg of crude glycerol
discounted fixed capital invested in the project. The project payback purified is $50.45 while the unit revenue is projected to be $80.36.
period is therefore just under three years after project start-up after year
3. The Discounted Break-even Period (DBEP) for the project would be Notes
6.3 and 6.6 years (between years 6 and 7) for 10% and 15% discounting
interest rates respectively as that is how long it would take for the The authors declare no conflict of interests.
discounted cumulative project capital to be recovered. So the net
payout time (NPT) is 3.3 and 3.6 years for 10% and 15% discounting Acknowledgments
rates respectively. NPT is the point from project start-up (Year 3) to
when cumulative cash flow becomes positive (between Years 6 and 7). Authors sincerely acknowledge Rlee Prokopishyn and Heli Eunike

86
C.G. Chol et al. Fuel Processing Technology 178 (2018) 78–87

for their technical guidance. For crude glycerol samples, authors thank 201400229.
Milligan biofuels Ltd., SK, Canada. [13] K.Y. Wang, T.-S. Chung, The characterization of flat composite nanofiltration
membranes and their applications in the separation of Cephalexin, J. Membr. Sci.
247 (2005) 37–50, http://dx.doi.org/10.1016/j.memsci.2004.09.007.
Funding [14] S. Kongjao, S. Damronglerd, M. Hunsom, Purification of crude glycerol derived
from waste used-oil methyl ester plant, Korean J. Chem. Eng. 27 (2010) 944–949,
http://dx.doi.org/10.1007/s11814-010-0148-0.
This work was supported by the Agriculture Development fund [15] N.K. Wang, L.K. Chen, J.P. Hung, Y.-T. Shammas, Membrane and desalination
(ADF), Saskatchewan Ministry of Agriculture (Canada) [award no. technologies, New York Handb. Environ. Eng, 2011, pp. 1–27.
20130289] and Mitacs Canada. [16] R. Singh, Membrane Technology and Engineering for Water Purification:
Application, Systems Design and Operation, Colorado Springs, 2015.
[17] X. Luo, X. Ge, S. Cui, Y. Li, Value-added processing of crude glycerol into chemicals
Appendix A. Supplementary data and polymers, Bioresour. Technol. 215 (2016) 144–154, http://dx.doi.org/10.
1016/j.biortech.2016.03.042.
[18] P. Saifuddin, N. Refal, H. Kumaran, Rapid purification of glycerol by-product from
Supplementary data to this article can be found online at https://
biodiesel production through combined process of microwave assisted acidification
doi.org/10.1016/j.fuproc.2018.05.023. and adsorption via chitosan immobilized with yeast, Res. J. Appl. Sci. Eng. Technol.
7 (2014) 593–602.
References [19] S. Sadhukhan, U. Sarkar, Production of purified glycerol using sequential desali-
nation and extraction of crude glycerol obtained during trans-esterification of
Crotalaria juncea oil, Energy Convers. Manag. 118 (2016) 450–458, http://dx.doi.
[1] A. Vlysidis, M. Binns, C. Webb, C. Theodoropoulos, A techno-economic analysis of org/10.1016/j.enconman.2016.03.088.
biodiesel biorefineries: assessment of integrated designs for the co-production of [20] W.N.R.W. Isahak, J.M. Jahim, M. Ismail, N.F. Nasir, M.M. Ba-Abbad, M.A. Yarmo,
fuels and chemicals, Energy 36 (2011) 4671–4683, http://dx.doi.org/10.1016/j. Purification of crude glycerol from industrial waste: experimental and simulation
energy.2011.04.046. studies, J. Eng. Sci. Technol. 11 (2016) 1056–1072.
[2] A. Almena, M. Martín, Technoeconomic analysis of the production of epi- [21] R. Manosak, S. Limpattayanate, M. Hunsom, Sequential-refining of crude glycerol
chlorohydrin from glycerol, Ind. Eng. Chem. Res. 55 (2016) 3226–3238, http://dx. derived from waste used-oil methyl ester plant via a combined process of chemical
doi.org/10.1021/acs.iecr.5b02555. and adsorption, Fuel Process. Technol. 92 (2011) 92–99, http://dx.doi.org/10.
[3] J.A. Posada, L.E. Rincón, C.A. Cardona, Design and analysis of biorefineries based 1016/j.fuproc.2010.09.002.
on raw glycerol: addressing the glycerol problem, Bioresour. Technol. 111 (2012) [22] M. Carmona, A. Lech, A. de Lucas, A. Pérez, J.F. Rodriguez, Purification of glycerol/
282–293, http://dx.doi.org/10.1016/j.biortech.2012.01.151. water solutions from biodiesel synthesis by ion exchange: sodium and chloride
[4] A.A. Apostolakou, I.K. Kookos, C. Marazioti, K.C. Angelopoulos, Techno-economic removal. Part II, J. Chem. Technol. Biotechnol. 84 (2009) 1130–1135, http://dx.
analysis of a biodiesel production process from vegetable oils, Fuel Process. doi.org/10.1002/jctb.2144.
Technol. 90 (2009) 1023–1031, http://dx.doi.org/10.1016/j.fuproc.2009.04.017. [23] A.H. Ooi, T.L. Yong, K.C. Dzulkefly, K. Wan Yunus, W.M.Z. Hazimah, Crude gly-
[5] Y. Xiao, G. Xiao, A. Varma, A universal procedure for crude glycerol purification cerine recovery from glycerol residue waste from a palm kernel oil methyl ester
from different feedstocks in biodiesel production: experimental and simulation plant, J. Oil Palm Res. 13 (2001) 16–22.
study, Ind. Eng. Chem. Res. 52 (2013) 14291–14296, http://dx.doi.org/10.1021/ [24] Y. Zhang, M.A. Dubé, D.D. McLean, M. Kates, Biodiesel production from waste
ie402003u. cooking oil: 2. Economic assessment and sensitivity analysis, Bioresour. Technol. 90
[6] M.S. Ardi, M.K. Aroua, N.A. Hashim, Progress, prospect and challenges in glycerol (2003) 229–240, http://dx.doi.org/10.1016/S0960-8524(03)00150-0.
purification process: a review, Renew. Sust. Energ. Rev. 42 (2015) 1164–1173, [25] T. Peters, Membrane technology for water treatment, Chem. Eng. Technol. 33
http://dx.doi.org/10.1016/j.rser.2014.10.091. (2010) 1233–1240, http://dx.doi.org/10.1002/ceat.201000139.
[7] R. Dhabhai, E. Ahmadifeijani, A.K. Dalai, M. Reaney, Purification of crude glycerol [26] P.T. Ulrich, G. Vasude, Chemical Engineering Process Design and Economics: A
using a sequential physico-chemical treatment, membrane filtration, and activated Practical Guide, Process Publishing, 2004.
charcoal adsorption, Sep. Purif. Technol. 168 (2016), http://dx.doi.org/10.1016/j. [27] R. Turton, R.C. Bailie, W.B. Whiting, J.A. Shaeiwitz, Analysis, Synthesis, and Design
seppur.2016.05.030. of Chemical Processes, Pearson Education, Inc., 2009.
[8] J. Van Gerpen, Biodiesel processing and production, Fuel Process. Technol. 86 [28] W. Hunsom, M. Saila, P. Chaiyakam, P. Kositnan, Comparison and combination of
(2005) 1097–1107, http://dx.doi.org/10.1016/j.fuproc.2004.11.005. solvent extraction and adsorption for crude glycerol enrichment, Int. J. Renew.
[9] H.W. Tan, A.R.A. Aziz, M.K. Aroua, Glycerol production and its applications as a Energy Res. 3 (2013) 364–371.
raw material: a review, Renew. Sust. Energ. Rev. 27 (2013) 118–127, http://dx.doi. [29] Cai Tianfeng, Li Huipeng, Zhao Hua, Liao Kejian, Purification of crude glycerol from
org/10.1016/j.rser.2013.06.035. waste cooking oil based biodiesel production by orthogonal test method, China Pet.
[10] F. Yang, M.A. Hanna, R. Sun, Value-added uses for crude glycerol—a byproduct of Process. Petrochem. Technol. 15 (2013) 48–53.
biodiesel production, Biotechnol. Biofuels 5 (2012) 13, http://dx.doi.org/10.1186/ [30] Dow Chemical Company, http://www.dow.com/optim/optimadvantage/physical-
1754-6834-5-13. properties/viscosity.htm, (2016).
[11] Z.J. He, I. Pinnau, A. Morisato, Nanostructured poly(4-methyl-2-pentyne)/silica [31] A. Singhabhandhu, T. Tezuka, A perspective on incorporation of glycerin pur-
hybrid membranes for gas separation, Desalination 146 (2003), http://dx.doi.org/ ification process in biodiesel plants using waste cooking oil as feedstock, Energy 35
10.1016/S0011-9164(02)00463-0. (2010) 2493–2504, http://dx.doi.org/10.1016/j.energy.2010.02.047.
[12] C. Rosaria, P.C. Della, R. Michele, P. Mario, Understanding the glycerol market, Eur.
J. Lipid Sci. Technol. 116 (2014) 1432–1439, http://dx.doi.org/10.1002/ejlt.

87

You might also like