ArticlePDF Available

Abstract and Figures

Design, economics, and plantwide control of a glycerol- tert -butyl alcohol (TBA) etherification plant are presented. The reaction takes place in liquid phase, in a plug flow reactor, using Amberlyst 15 as a catalyst. The products' separation is achieved by two distillation columns where high-purity ethers are obtained and a section involving extractive distillation with 1,4-butanediol as solvent, which separates TBA from the TBA/water azeotrope. Details of design performed in AspenPlus and an economic evaluation of the process are given. Three plantwide control structures are examined using a mass balance model of the plant. The preferred control structure fixes the fresh glycerol flow rate and the ratio glycerol + monoether : TBA at reactor-inlet. The stability and robustness in the operation are checked by rigorous dynamic simulation in AspenDynamics.
Content may be subject to copyright.
The Scientific World Journal
Volume 2012, Article ID 180617, 11 pages
doi:10.1100/2012/180617
The cientificWorldJOURNA
L
Research Article
Design and Control of Glycerol-
tert
-Butyl Alcohol
Etherification Process
Elena Vlad, Costin Sorin Bildea, and Grigore Bozga
Department of Chemical Engineering, University Politehnica of Bucharest, Street Gh. Polizu 1-7, 011061 Bucharest, Romania
Correspondence should be addressed to Costin Sorin Bildea, sbildea@upb.ro
Received 7 September 2012; Accepted 5 November 2012
Academic Editors: G. Morales and O. A. Scelza
Copyright © 2012 Elena Vlad et al. This is an open access article distributed under the Creative Commons Attribution License,
which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.
Design, economics, and plantwide control of a glycerol-tert-butyl alcohol (TBA) etherification plant are presented. The reaction
takes place in liquid phase, in a plug flow reactor, using Amberlyst 15 as a catalyst. The products’ separation is achieved by two
distillation columns where high-purity ethers are obtained and a section involving extractive distillation with 1,4-butanediol
as solvent, which separates TBA from the TBA/water azeotrope. Details of design performed in AspenPlus and an economic
evaluation of the process are given. Three plantwide control structures are examined using a mass balance model of the plant.
The preferred control structure fixes the fresh glycerol flow rate and the ratio glycerol + monoether :TBA at reactor-inlet. The
stability and robustness in the operation are checked by rigorous dynamic simulation in AspenDynamics.
1. Introduction
As byproduct of biodiesel production, one mole of glycerol
(G) is produced for every three moles of methyl esters, which
is equivalent to approximately 10% wt. of the total product.
As a result of the increased availability, the market price of
glycerol has dropped rapidly. Therefore, new uses for glycerol
need to be found. Although glycerol could be burnt as a fuel,
it could also be processed into more valuable components
[1,2].
Di- and triethers of glycerol are compounds soluble in
diesel and biodiesel, improving the quality of the fuel [3].
They diminish the emissions of particulate matter, carbon
oxide, and carbonyl compounds. Moreover, they provide
a 5 K reduction in cloud point and an 8% reduction in
viscosity when combined with biodiesel [4]. Therefore,
ethers of glycerol are interesting alternatives to commercial
oxygenate additives such as MTBE, ETBE, or TAME. Ethers
of glycerol can be obtained by etherification with olefins such
as i-butene (IB), alcohols such as tert-butyl alcohol (TBA) or
ethanol [5]orbytrans-esterification with another ester such
as methyl-t-butyl ether.
Reaction of i-butene with glycerol in presence of homo-
geneous [6] or heterogeneous [7,8] acid catalysts yields
a mixture of mono-, di-, and tri-tert-butyl glycerol ethers
(ME, DE, and TE, resp.). Several processes were proposed to
perform this transformation [6,912]. In all these processes,
the reaction takes place at 14 bar, necessary for keeping
the i-butene in liquid phase. Moreover, from the reactor
outlet, i-butene is separated as a vapour stream and must
be recompressed before being recycled, which is another
drawback of the process.
The etherification reaction could be performed at
lower pressure using tert-butyl alcohol as reactant and
ion exchange resins as catalyst [13,14], according to the
following reactions:
G+TBAME + H2O
ME + TBA DE + H2O
DE + TBA TE + H2O
(1)
Small amounts of i-butene are formed due to TBA dehydra-
tion:
TBA IB + H2O(2)
Yusof et al. [15] report glycerol conversions exceeding
64% and mono- to diether ratio in the range 4 : 1–6 : 1,
obtained using Amberlyst 15, Amberlite IR-120, Montmo-
rillonite K10, p-toluenesulfonic acid, and sulfuric acid as
2The Scientific World Journal
Reactor Separation
TBA
Glycerol
TBA recycle
Glycerol and
monoether recycle
0
0
1b
1a
2
3a
4a
Diether
Triether
MIX-TBA
MIX-GE
4b
Water
3b
Figure 1: Reactor-separation-recycle structure of the glycerol-TBA etherification plant.
catalysts. Frusteri et al. [16] studied the etherification of
glycerol with tert-butyl alcohol in presence of lab-made silica
supported acid catalysts. Experiments were carried out in
batch mode at temperature ranging from 303 to 363 K. Ozbay
et al. [17] compared various solid acid catalysts, such as
Amberlyst-15 (A-15), Amberlyst-16 (A-16) and Amberlyst-
35 (A-35), Nafion-SAC-13, and gamma-alumina. Amberlyst
15 showed the highest activity at about 110 degrees C,
while A-16 gave higher diether selectivity values. Chang and
Chen [18] present a systematic optimization of the glycerol
etherification with alcohol-tert-butylic based on the small-
scale experimental data. The reaction conditions (reacting
temperature, catalyst loading, solvent loading) leading to
maximum glycerol ethers concentration were investigated.
Kiatkittipong et al. [19] present several kinetic models for
glycerol etherification with TBA, with parameters obtained
by regressing measured data from an autoclave reactor.
Experiments were performed in a reactive distillation col-
umn, the results being compared with AspenPlus simulation
predictions. It should be remarked that the column worked
as a series of CSTRs (the trays) where the reactants were fed
in countercurrent. Therefore, the distillate stream contained
a mixture of water and TBA, while the bottom stream
contained TBA, glycerol, and mono-, di-, and triether. As
a result, separation of reactants and products from the
column-outlet streams was still necessary. In a recent paper
[20], the liquid phase etherification of glycerol with tert-
butyl alcohol was investigated in a continuous-flow reactor
using Amberlyst-15 as catalyst.
In a dierent reaction pathway, Al-Lal et al. [21]suggest
dehydration to epichlorohydrin followed be etherification
with TBA.
Although the feasibility of glycerol etherification with
TBA was proved by experimental results, no design, eco-
nomic evaluation, and controllability analysis of the entire
processhavebeenreported.Thegoalofthisworkistofill
this gap. The paper is organized as follows. Next section
presents the conceptual design of the plant, using a steady
state mass-balance model which includes a rigorous model
of the chemical reactor and ideal models for the separation
section. Following the degree of freedom analysis, plantwide
control structures are suggested. For each control structure,
the model is solved and the influence of operating parameters
on the reactants conversion is analyzed. Thus, a control
structure is selected, allowing processing of variable amounts
of glycerol and ensuring a unique, steady stable state. Then,
detailed design of the separation section is performed and
an economic evaluation of the process is accomplished. The
performance of the plantwide control (ability to process
impure glycerol and to change the production rate) is proved
by means of rigorous dynamic simulation performed in
AspenDynamics.
2. Conceptual Design and Plantwide Control
2.1. Reactor-Separation-Recycle Model. In this section, a
simplified model of the plant will be used to assess the
feasibility of the process and to perform a preliminary mass
balance of the plant. Based on this model, several plantwide
controlstructureswillbesuggestedandevaluated.Figure 1
presents the Reactor-Separation-Recycle structure of the
plant [22]. After the reaction takes place, the reactor euent
enters the separation section. Here, the di-, triethers and
water are removed from the plant, while the monoether and
unconverted reactants (glycerol and TBA) are recycled to the
reactor, after being mixed with fresh reactants. The notations
used here will follow Figure 1.Fk,jwill denote the mole flow
rate of species kin stream j.SubscriptsG,TBA,ME,DE,TE,
and W will be used for glycerol, tert-butyl alcohol, mono-,
The Scientific World Journal 3
di-,triethers,andwater,respectively.Subscripts0,1,2,3and
4 will denote the feed, reactor-inlet, reactor-outlet, recycle,
and product streams, respectively. For example, FG,1 stands
for the flow rate of glycerol at reactor inlet.
The model of the plant assumes a plug-flow reactor
(PFR) operated at a fixed temperature (70C) and perfect
separation. The etherification reactions leading to mono-, di-
, and triethers were considered, while TBA dehydration was
neglected (this assumption will be relaxed when designing
the whole plant). It was assumed that TBA, glycerol, and
monoether are recovered from the reactor outlet stream and
recycled. The model consists of (3) describing the reactor
and from (4)to(9) which describe the separation and
recycle:
dFj
dm =vR,j,j=G, TBA, ME, DE, TE, W
at m=0Fj=Fj,1 ,m(0, mcat),
(3)
FG,0 +FG,2FG,1 =0, (4)
FTBA,0 +F
TBA,2FTBA,1 =0, (5)
FME,1 FME,2 =0, (6)
FDE,1 =0, FDE,4 =FDE,2,(7)
FTE,1 =0, FTE,4 =FTE,2,(8)
FW,1 =0, FW,4 =FW,2,(9)
The reaction rates are calculated by (10)to(13), where νis
the matrix of stoichiometric coecients. The reaction rate
constants follow Arrhenius temperature dependence with
parameters given in Tabl e 1 [19]:
vR,j=
k
νk,jrk,(10)
r1=k1xGxTBA 1
Keq,1
xMExW, (11)
r2=k2xMExTBA 1
Keq,2
xDExW, (12)
r3=k3xDExTBA 1
Keq,3
xTExW, (13)
r4=k4xTBA.(14)
2.2. Plantwide Control Structures. The design degrees of
freedom represent the variables that must be specified
to completely define the process. Their number can be
calculated by subtracting the number of equations from the
number of variables. The control degrees of freedom are the
variables that can be manipulated, namely, the control valves
in the process. For most processes, the number of design
degrees of freedom is equal to the number of control degrees
of freedom [23]. The model of the glycerol etherification
plant consists of 15 equations and contains 17 variables
Tab le 1: Equilibrium constants and rate constants [19].
Equilibrium constant Rate constant/(mol s1kg1)
Keq1 =exp(2.581 754.8/T)k1=exp(17342 6835/T)
Keq2 =exp(1.228 942.1/T)k2=exp(26953 10382/T)
Keq3 =exp(1.779 2212/T)k3=exp(26953 10382/T)
k4=exp(23.358 12480/T)
(6 reactor-inlet, 6 reactor-outlet, 2 fresh reactants, and 3
product flow rates). Therefore, 2 degrees of freedom must be
fulfilled. This is in agreement with Figure 1, where two valves
must be used for level control, leaving 2 valves available for
manipulating flow rates.
2.3. Steady State Behavior. Theaimofthissectionisto
investigate the steady state behavior of an etherification plant
when dierent plantwide control structures are applied. We
will assume that the nominal flow rate of fresh glycerol
is 2.15 kmol/h, which is the typical output of a 15,000
tones/year biodiesel plant.
It should be remarked that the main task of plantwide
control system is controlling the inventory of reactants,
products, and impurities. Controlling the inventory of
reactants within the plant can be performed in two ways
[24]: (a) by evaluating, directly or indirectly, the inventory
of each reactant and controlling it by feedback using the
corresponding fresh feed as manipulated variable; (b) by
fixing the fresh feed rate and using the self-regulation
property of the mass balance [25]. The latter assumes that
the entire amount of reactant brought into the process is
converted into products, which are subsequently separated
and removed from the plant. Consequently, three control
structures will be further considered. Control structure CS1
attempts controlling the inventory of reactants by the use of
feedback [24]. Thus, the flow rates of TBA (F1b =FTBA,1)and
glycerol + ethers (F1a =FG,1 +FME,1) at reactor-inlet are fixed.
The amount of reactants in the buer vessels are used as
indirect indications of inventories. Therefore, accumulation
or depletion of reactants is avoided by adjusting the fresh
reactant feed rates.
Control structures CS2 and CS3 make use of the self-
regulating property of the mass balance [25]. In both control
structures the flow rate of fresh glycerol is set to the value
FG,0. Control structures CS2 and CS3 dier by the second
flow specification: TBA at reactor inlet (F1b =FTBA,1)and
ratio r1=F1b/F1a, respectively. It should be remarked that,
for all control structures, the amount of product obtained
equals the amount of glycerol fed in the process, F4a =FG,0.
2.3.1. Control Structure CS1: Glycerol and TBA Inventories
Controlled by Feedback. Figure 2 presents the principle of
control structure CS1 and results concerning the behavior
of the plant when this control structure is applied. The
reactor uses 400 kg of catalyst. The top diagram shows
the amount of glycerol that is processed (FG,0) versus the
glycerol-ethers reactor-inlet flow rate (F1a), at dierent values
of the reactor-inlet TBA flow rate (FTBA,1). It can be observed
4The Scientific World Journal
TBA
FC
LC
Glycerol
0
0
1b
1a
3b
3a
TBA recycle
Glycerol and
ethers recycle
FC
LC
To reactor
To reactor
FTBA, 1
F1a
(a)
0
1
2
3
4
0246810
20 15
mcat =400 kg
FTBA, 1/(kmol/h) =25
FG,0/(kmol/h)
F1a/(kmol/h)
(b)
0.2
0.4
0.6
0.8
1
0246810
20
15
FTBA, 1/(kmol/h) =15
XG
F1a/(kmol/h)
mcat =400 kg
(c)
0
0.1
0.2
0.3
0.4
0246810
20 15
mcat =400 kg
FTBA, 1/(kmol/h) =25
F1a/(kmol/h)
XTBA
(d)
Figure 2: Control structure CS1. Processed glycerol and reactants conversion versus reactor-inlet glycerol + ethers flow rate.
that the nominal capacity of 2.15 kmol/h could be increased
by changing the flow rate F1a or by modifying the TBA flow
at reactor inlet FTBA,1.
The lower diagrams show the dependence of the glycerol
and TBA conversions, XG=1FG,2/FG,1,andXTBA,1 =
1FTBA,2/FTBA,1 versus the flow F1a.ForeachF1a value,
a single steady state exists, which is an advantage of this
control structure. However, because the fresh glycerol is
used to control the buer-vessel level, this control structure
cannot be applied when the flow rate of glycerol is set by the
upstream biodiesel plant.
2.3.2. Control Structures CS2-CS3: Self-Regulating Glycerol
Inventory. In control structures CS2 and CS3 the flowrate
of fresh glycerol FG,0 is set, which is a very convenient,
direct way to change the production rate. In addition, control
structure CS2 fixes the reactor-inlet TBA flow rate, while the
ratio TBA: glycerol + ME is fixed in CS3.
Figure 3 presents results obtained when CS2 is applied.
The top diagram shows glycerol conversion (XG) plotted
versus the flow rate of fresh glycerol (FG,0), for dierent
amounts of catalyst used in the reactor. The system exhibits
two steady states—at small FG,0, or no steady state at
all—at large FG,0. This behavior is a major disadvantage
of this control structure. It can be observed that an
amount of 200 kg of Amberlyst is sucient to process
2.15 kmol/h of glycerol, but does not allow a large increase
of this value. However, 400 kg of catalyst ensures enough
flexibility.
The bottom diagrams present the conversion of the
glycerol and TBA versus the fresh glycerol flow rate, for
dierent values of reactor-inlet TBA flow rate and 400kg of
catalyst. The extent of the feasibility region increases with
the reactor-inlet TBA flow rate. It can be observed that TBA
conversion has (almost) the same value on both branches.
Moreover, TBA conversion is independent of the amount of
catalyst used:
XTBA 2FG,0
FTBA,1
.(15)
Figure 4 presents results obtained when the control struc-
ture CS3 is used. Glycerol conversion versus fresh glycerol
flow rate is plotted for dierent amounts of catalyst. Inde-
pendently on the catalyst mass and ratio between reactor-
inlet flow rates, a unique steady states exists. Compared
to CS1, dierent amounts of glycerol can be processed.
However, there is a maximum flow rate of glycerol that can
be processed, which increases with the catalyst amount. It can
be seen that 400 kg of catalyst allows doubling the production
rate. For 400 kg of catalyst, the lower diagrams show that the
glycerol and TBA conversions have small sensitivity to the
production rate or the ratio between reactor-inlet flowrates.
The Scientific World Journal 5
TBA
FC
LC
Glycerol
LC
FC
0
0
1b
1a
3b
3a
TBA recycle
Glycerol and ethers
recycle
To reactor
To reactor
FTBA,1
FG,0
(a)
0
0.2
0.4
0.6
0.8
1
012345
300 400
XG
FG,0/(kmol/h)
mcat/(kg) =200
FTBA, 1 =20 (kmol/h)
(b)
0
0.2
0.4
0.6
0.8
1
012345
20
25
XG
FG,0/(kmol/h)
mcat =400 kg
FTBA, 1/(kmol/h) =15
(c)
0
0.2
0.4
0.6
0.8
1
012345
20 25
FG,0/(kmol/h)
mcat =400 kg
FTBA, 1/(kmol/h) =15
XTBA
(d)
Figure 3: Control structure CS2. Reactants conversions versus fresh glycerol flow rate, for dierent values of catalyst mass and reactor-inlet
TBA flow rate.
In conclusion, control structure CS3 oers the advantage
of a unique steady state together with easily setting the fresh
glycerol flow rate.
2.4. Separation Section. Liquid-liquid and vapor-liquid equi-
libria were analyzed using AspenPlus. Glycerol and TBA are
present in AspenPlus database, from where their physical
properties were taken. After defining the molecular structure
of the ethers, their properties were estimated using group
contribution methods. The behavior of the liquid phase
was described by the NRTL activity model. The interac-
tion parameters were taken from Aspen Plus database or
were estimated using UNIFAC Dortmund modified method
[26].
Tab l e 2 presents the boiling points of the main compo-
nents and their azeotropes. Small amounts of i-butene that
are formed by TBA dehydration can be easily removed due
to lower boiling point. The separation of TBA and water
from glycerol-ethers mixtures appear to be easy and will be
handled by distillation (column C1).
Also, glycerol and monoether which are recycled can
be obtained as a bottom product of a distillation column
(column C2). Obtaining high purity DE product seems
dicult due to the low-boiling G-DE azeotrope. However,
the residue curve map (Figure 5) of the DE-ME-G mixture
shows only one distillation region where ME acts as a solvent
for glycerol, allowing therefore high-purity diether to be
obtained in one distillation unit.
TBA and water form a low-boiling homogeneous
azeotrope. This can be broken by using a suitable solvent, for
example, 1,4-butanediol.
Figure 6 shows the residue curve map of the TBA-Water-
1,4-butanediol mixture.
The water-TBA mixture is firstly separated to TBA and
azeotrope (column C3). The azeotrope enters in the lower
part of the extractive distillation column (EX), while the
solvent is fed at the top. The distillate contains water, while
the bottom stream consists of solvent and TBA, which is
further separated in column C4.
3. Plant Flowsheet
Figure 7 presents the flowsheet, while Figure 8 details the
azeotrope separation section. The control loops are also
depicted. A detailed stream report of each section is pre-
sented in Tables 3and 4. The etherification of glycerol with
TBA takes place in a plug flow reactor in the presence of
400 kg Amberlyst. The reaction temperature and pressure are
set to 70C and 5 bar, respectively, when the reaction mixture
is liquid. The reactor-outlet stream is routed to Column C1.
TBA and water are separated as top product, while a mixture
of glycerol and ethers leaves the column as bottom product.
6The Scientific World Journal
TBA
LC
Glycerol
LC
FC
0
0
1b
1a
3b
3a
TBA recycle
Glycerol and ethers
recycle
To reactor
To reactor
x
FC
FC
(a)
0
0.2
0.4
0.6
0.8
1
012345
300
400
XG
FG,0/(kmol/h)
mcat/(kg) =200
r1=4
(b)
0
0.2
0.4
0.6
0.8
1
012345
3
4
XG
FG,0/(kmol/h)
mcat/(kg) =400
r1=2
(c)
0
0.1
0.2
0.3
0.4
012345
3
4
FG,0/(kmol/h)
mcat/(kg) =400
r1=2
XTBA
(d)
Figure 4: Control structure CS3. Reactants conversions versus fresh glycerol flow rate, for dierent values of catalyst mass and ratio between
reactor-inlet flow rates.
The column is operated under vacuum (0.1 bar) to avoid
high temperature in the bottom of the column. Column
C2 separates the mixture of di- and triethers. The bottom
product, containing glycerol and monoether, is recycled. The
column is operated under vacuum (0.1 bar) to avoid product
degradation.
Column C3 separates the TBA/water azeotrope from
TBA which is mixed with fresh TBA and recycled. The
extractive distillation column (EX) is fed on bottom with
TBA/water azeotrope and on top with 1,4-butanediol, which
is the solvent. The solvent extracts TBA and is eliminated
on the bottom of the column, while the water is removed
as liquid on the top. The column has partial condenser in
order to eliminate isobutene traces. Column C4 recovers the
solvent. TBA is removed on the top of the column, is mixed
with TBA stream from column C1 and with fresh TBA and is
recycled.
An economic evaluation of the process was performed.
A payback period of 10 years was considered and the
total annual cost of the plant (TAC) was calculated as the
following:
TAC =capital cost
payback period +energycost.(16)
The capital cost, including the costs of reactor, distillation
columns, and extractive distillation column, was calcu-
lated using well-known relationships [27]. The energy cost
Tab le 2: Boiling point for pure components and azeotropes at P=
1bar.
Component/azeotrope T/(C) Destination
IB 6.25 Byproduct
TBA (0.6209)/W (0.3791) 79.97
TBA 82.42 Recycle
W 100 Byproduct
G (0.1951)/DE (0.8049) 233.5
DE 240.4 Product
ME 256.61 Recycle
G 287.85 Recycle
includes the costs of cooling water (0.08 US$/m3)and
electricity (8·106US$/kJ). Tab l e 5 summarizes the results.
4. Dynamics and Control
The dynamics of the plant must be considered in order to
prove the stability of the operating point and the resiliency
with respect to disturbances.
For control structure CS3, a dynamic model of the
plant was built in AspenDynamics [28]. The controllers
were tuned by a simple version of the direct synthesis
method. According to this method, the desired closed-
loop response for a given input is specified. Then, with
the model of the process known, the required form and
The Scientific World Journal 7
(DE)
(G)
(ME)
0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 0.55 0.6 0.65 0.7 0.75 0.8 0.85 0.9 0.95
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
0.95 0.85 0.75 0.65 0.55 0.45 0.35 0.25 0.15 0.05
Figure 5: Residue curve map of the glycerol-monoether-diether mixture.
TBA
Solvent
Water
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1
C3
C4
EX
Figure 6: Residue curve map of the TBA-water-1,4-butanediol mixture.
8The Scientific World Journal
1
TBA1
GM1
2
3a
3b
TBA0
G0
W
Azeotrope
separation
C1
C2
PFR
x
r1
LC FC
FC
PC
FC
LC
PC
LC
XCTC
TC
TC
LC
LC
D
+
T
Figure 7: Flowsheet and control loops of the reaction section.
X
3a
S0
S
TBA1b
TBA1a
AZ
W
IB
PC
PC
LC
LC
TC
TC
TC
LC
PC
LC
TC
TC
TC
XC
XC
TC
LC
C3
EX
C4
TBA
+
S
Figure 8: Flowsheet and control loops of the azeotrope separation section.
The Scientific World Journal 9
Tab le 3: Stream results for the reaction section (Figure 7).
Stream name G0 TBA1 TBA0 GM1 1 2 3a 3b W D + T
Flow/(kmol/hr) 2.15 26.5 4.37 6.23 32.75 32.8 26.47 6.32 4.35 2.24
Flow/(kg/hr) 198 1883 324 759 2642 2642 1635 1006.6 80.8 445.6
Tem p / (C) 20 79.8 25 140.9 70 70 29.3 177 25 139.5
Pressure/(bar) 1 5 1 5 5 4.95 1.2 0.15 1 0.1
Flow/(kmol/hr)
G 2.15 Trace 0 2.94 2.94 0.8 Trace 0.81 0 0.01
TBA 0 24.9 4.37 Trace 24.9 20.6 20.6 0.002 0.04 0.003
ME 0 Trace 0 3.27 3.27 3.27 Trace 3.27 0 0.007
DE 0 Trace 0 0.01 0.01 2.14 Trace 2.14 0 2.13
TE 0 Trace 0 Trace Trace 0.002 Trace 0.002 0 0.0019
Water 0 1.5 0 Trace 1.5 5.81 5.81 0.005 4.31 0.005
S 0 0.1 0 0 0.1 0.1 Trace 0.1 Trace 0.1
IB 0 0 0 0 Trace 0.04 0.04 Trace 0 Trace
Tab le 4: Stream results for the azeotrope separation section (Figure 8).
Stream name 3a TBA1a AZ S0 W TBA + S IB TBA1b S
Flow/(kmol/hr) 26.46 11.7 14.7 90.6 4.35 100.9 0.04 10.38 90.5
Flow/(kg/hr) 1635 858.5 776.6 8164.5 80.8 8858 2.22 700.1 8158
Tem p . / (C) 29.3 97.8 76 30 25 156 25 80 231
Pressure/(bar) 1.2 1.8 1.0 1.2 1 1.5 1 1.0 1.13
Flow/(kmol/hr)
TBA 20.61 11.5 9.1 0 0.04 9.05 Trace 9.03 Trace
IB 0.04 Trace 0.04 0 0.003 Trace 0.039 Trace Trace
S 0 0.0 Trace 90.6 Trace 90.6 0 0.1 90.5
Water 5.8 0.2 5.57 0 4.31 1.2 0.001 1.2 Trace
Tab le 5: Economic evaluation.
Reactor Column C1 Column C2 Column C3 Column EX Column C4
Diameter (m) 0.35
Reflux ratio 0.32 1.83 4 1.5 2
Diameter (m) 0.85 0.55 0.7 0.5 0.55
No of trays 6 15 41 30 7
Height (m) 2 Reboiler duty (kW) 343 170 716.2 525 563.3
Condenser duty (kW) 348.7 168.6 654 170 324.4
Cost ($) 9296 Cost ($) 297 966 221 775 589 424 759 704 338 157
Energy Cost =291 663$/year
Equipment Cost =2 440 746$
TAC =535 738$/year (10 years payback)
the tuning of the feedback controller are back-calculated.
For all controllers, the acceptable control error, Δεmax ,
and the maximum available control action, Δumax,were
specified. Then the controller gain, expressed in engineering
units, was calculated as Kc=Δumax/Δεmax and translated
into percentage units. First-order open-loop models were
assumed in order to calculate the integral time of the pressure
and temperature control loops. As rough evaluations of the
process time constants τ, 12min and 20min were used,
respectively. It can be shown that the direct synthesis method
requires that the reset time of a PI controller is equal to the
time constant of the process, τi=τ. For the level controllers,
a large reset time τi=60 min was chosen as no tight control
is required.
Figure 9 presents results of dynamic simulation. Molar
and mass flow rates together with mass fractions are
shown. Starting from the steady state (fresh glycerol:
198 kg/h), two disturbances were introduced. At time of
2 h, a 10% wt. water impurity in the fresh glycerol was
introduced. Later (time =40 h), the flow rate of fresh
glycerol (90% wt. purity) was increased to 220 kg/h. It can
be seen that the nominal operating point is stable, and
the plant achieves stable operation when disturbances are
introduced.
10 The Scientific World Journal
0
200
400
600
800
1000
0 20406080100
500
1000
1500
2000
2500
F/(kg/h)
F/(kg/h)
FTBA
FGM1
FTBA0
t/(h)
FG0
(a)
50
200
350
500
0 20 40 60 80 100
0.95
0.96
0.97
0.98
0.99
1
Mass fraction
F/(kg/h)
FW
FD+T
XD+T
XD
Xw
t/(h)
(b)
0
5
10
15
20
0 20406080100
0
10
20
30
40
F/(kmol/h)
F/(kmol/h)
FG0
t/(h)
FAZ
FGM1
FTBA0
FTBA1
FW
FD+T
(c)
0
500
1000
1500
0 20406080100
0.95
0.96
0.97
0.98
0.99
1
Mass fraction
F/(kg/h)
t/(h)
FAZ
FTBA1b
XAZ
XTBA1b
(d)
Figure 9: Dynamic simulation results. Refer to Figures 7and 8for nomenclature of various streams. xTBA1b and xAZ are mass fractions of
TBA in streams TBA1b and AZ. xDis the mass fraction of diether in the product stream D + T.
5. Conclusions
Production of glycerol ethers by etherification of glycerol
with tert-butyl alcohol catalyzed by heterogeneous acid
catalysts, such as Amberlyst 15, is feasible. For a typical
glycerol flow rate of 2.15kmol/h, the reaction can be carried
on in a PFR using 400 kg of catalyst. The glycerol conversion
is high. However, recycle of the monoether byproduct is
necessary. The separation products-unconsumed reactants,
are dicult due to formation of the water-TBA azeotrope,
which can be broken using a suitable solvent. The TAC of the
plant is rather high, 536 000 USD/year. The recommended
control structure sets the fresh glycerol feed rate and the ratio
G + ME :TBA at reactor inlet.
Acknowledgments
The work has been funded by the Sectoral Operational Pro-
gramme Human Resources Development 2007–2013 of the
Romanian Ministry of Labour, Family and Social Protection
through the Financial Agreement POSDRU/88/1.5/S/61178
and by CNCSIS-UEFISCSU, Project IDEI 1545/2008-
Advanced modeling and simulation of catalytic distillation
for biodiesel synthesis and glycerol transformation.
References
[1] M. Pagliaro, R. Ciriminna, H. Kimura, M. Rossi, and C. D.
Pina, “Recent advances in the conversion of bioglycerol into
value-added products,European Journal of Lipid Science and
Technology, vol. 111, no. 8, pp. 788–799, 2009.
[2] N. Rahmat, A. Z. Abdullah, and A. R. Mohamed, “Recent
progress on innovative and potential technologies for glycerol
transformation into fuel additives: a critical review,Renew-
able and Sustainable Energy Reviews, vol. 14, no. 3, pp. 987–
1000, 2010.
[3] R. S. Karinen and A. O. I. Krause, “New biocomponents from
glycerol,Applied Catalysis A, vol. 306, pp. 128–133, 2006.
[4] A. Demirbas, “Glycerol-based fuel oxygenates for biodiesel
and diesel fuel blends,Energy Sources A, vol. 31, no. 19, pp.
1770–1776, 2009.
[5] J. A. Melero, G. Vicente, G. Morales, M. Panigua, G. Morates,
and P. Munoz, “Etherification of biodiesel-derived glycerol
with ethanol for fuel formulation over sulfonic modified
catalysts,Bioresource Technology, vol. 103, pp. 142–151, 2012.
[6] A. Behr and L. Obendorf, “Development of a process for
the acid-catalyzed etherification of glycerine and isobutene
forming glycerine tertiary butyl ethers,” Chemical Engineering
and Technology, vol. 25, no. 7, pp. 185–189, 2003.
[7] K. Klep ´
aˇ
cov´
a, D. Mravec, A. Kaszonyi, and M. Bajus, “Etherifi-
cation of glycerol and ethylene glycol by isobutylene,Applied
Catalysis A, vol. 328, no. 1, pp. 1–13, 2007.
[8] J. A. Melero, G. Vicente, G. Morales et al., “Acid-catalyzed
etherification of bio-glycerol and isobutylene over sulfonic
mesostructured silicas,Applied Catalysis A, vol. 346, no. 1-2,
pp. 44–51, 2008.
[9] W. N. Versteeg, O. Ijben, W. N. Wernink, K. Klepacova, and S.
van Loo, “Method of preparing GTBE,” WO 2009/147541 A1,
2009.
[10] V. P. Gupta, “Glycerine ditertiary butyl ether preparation,” US
5476971, 1995.
The Scientific World Journal 11
[11] H. Noureddini, “Process for producing biodiesel fuel with
reduced viscosity and a cloud point below 32 degrees Fahren-
heit,” US 6015440, 2000.
[12] J. K. Cheng, C. L. Lee, Y. T. Jhuang, J. D. Ward, and I. L.
Chien, “Design and control of the glycerol tertiary butyl ethers
process for the utilization of a renewable resource,Industrial
& Engineering Chemistry Research, vol. 50, pp. 12706–12716,
2011.
[13] K. Klep ´
aˇ
cov´
a, D. Mravec, and M. Bajus, “tert-Butylation of
glycerol catalysed by ion-exchange resins,Applied Catalysis A,
vol. 294, no. 2, pp. 141–147, 2005.
[14] K. Klep´
aˇ
cov´
a, D. Mravec, and M. Bajus, “Etherification of
glycerol with tert-butyl alcohol catalysed by ion-exchange
resins,Chemical Papers, vol. 60, no. 3, pp. 224–230, 2006.
[15] Y. A. B. Yusof, R. B. Abu Bakar, O. T. Lye, and S. B. Ahmad,
“Process for producing etherified compounds from alcohol,
US 0167503, 2008.
[16] F. Frusteri, F. Arena, G. Bonura, C. Cannilla, L. Spadaro,
and O. di Blasi, “Catalytic etherification of glycerol by tert-
butyl alcohol to produce oxygenated additives for diesel fuel,
Applied Catalysis A, vol. 367, no. 1-2, pp. 77–83, 2009.
[17] N. Ozbay, N. Oktar, G. Dogu, and T. Dogu, “Conversion
of biodiesel by-product glycerol to fuel ethers over dierent
solid acid catalysts,International Journal of Chemical Reactor
Engineering, vol. 8, article A18, 2010.
[18] J. S. Chang and D. H. Chen, “Optimization on the etherifica-
tion of glycerol with tert-butyl alcohol,Journal of the Taiwan
Institute of Chemical Engineers, vol. 42, no. 5, pp. 760–767,
2011.
[19] W. Kiatkittipong, P. Intaracharoen, N. Laosiripojana, C.
Chaisuk, P. Praserthdam, and S. Assabumrungrat, “Glycerol
ethers synthesis from glycerol etherification with tert-butyl
alcohol in reactive distillation,Computers and Chemical
Engineering, vol. 35, no. 10, pp. 2034–2043, 2011.
[20] N. Ozbay, N. Oktar, G. Dogu, and T. Dogu, “Eects of sorption
enhancement and isobutene formation on etherification of
glycerol with tert-butyl alcohol in a flow reactor,Industrial &
Engineering Chemistry Research, vol. 51, pp. 8788–8795, 2012.
[21] A.-M. Al-Lal, J.-E. Garc´
ıa-Gonz´
alez, A. Llamas, A. Monjas,
andL.Canoira,“Anewroutetosynthesizetert -butyl ethers
of bioglycerol,Fuel, vol. 93, no. 1, pp. 632–637, 2012.
[22] A. C. Dimian and C. S. Bildea, Chemical Process Design:
Computer-Aided Case Studies, Wiley-VCH, Manheim, Pa,
USA, 2008.
[23] W. L. Luyben, “Design and control degrees of freedom,
Industrial & Engineering Chemistry Research, vol. 35, no. 7, pp.
2204–2214, 1996.
[24] C. S. Bildea and A. C. Dimian, “Fixing flow rates in recycle
systems: Luyben’s rule revisited,Industrial & Engineering
Chemistry Research, vol. 42, no. 20, pp. 4578–4585, 2003.
[25] J. Downs, “Distillation control in a plantwide control environ-
ment,” in Practical Distillation Control, W. Luyben, Ed., D. Van
Nostrand, New York, NY, USA, 1992.
[26] U. Weidlich and J. Gmehling, “A modified UNIFAC model
1. Prediction of VLE, hE, and γ,” Industrial & Engineering
Chemistry Research, vol. 26, no. 7, pp. 1372–1381, 1987.
[27] A. C. Dimian, Integrated Design and Simulation of Chemical
Processes, Elsevier, Amsterdam, The Netherlands, 2003.
[28] Aspen Technology, Aspen Plus User Guide, Aspen Technology,
Burlington, Mass, USA, 2009.
... The thermal desorption of adsorbed ammonia and of ammonia released by the thermal decomposition of ammonium ions of the zeolite was performed with a TPDRO 1100 series device (Thermo Scientific, Schwerte, Germany). The 29 Si MAS NMR measurements were recorded on an NMR spectrometer (DMX-400, Bruker), whereas the 27 Al MAS NMR and 1 H MAS NMR measurements were performed on the NMR spectrometer (AVANCE-600 spectrometer, Bruker). Before measurement, all samples were activated at 400 • C for 1 h, at a heating rate of 5 • C/min. ...
... The natural zeolite clinoptilolite precursor, ammonium exchanged clinoptilolite, and different calcined samples were characterized regarding their structure, crystallinity, morphology, stability, porosity, and acidic properties by XRD, SEM/TEM, FTIR lattice vibration spectra, as well as by nitrogen adsorption and desorption, TG-DSC and TPDA measurements. Additionally, 29 Si MAS NMR, 27 Al MAS NMR, 1 H MAS NMR measurements as well as selected 1 H- 29 Si NMR Redor and 1 H- 27 Al NMR Trapdor experiments were performed. ...
... The natural zeolite clinoptilolite precursor, ammonium exchanged clinoptilolite, and different calcined samples were characterized regarding their structure, crystallinity, morphology, stability, porosity, and acidic properties by XRD, SEM/TEM, FTIR lattice vibration spectra, as well as by nitrogen adsorption and desorption, TG-DSC and TPDA measurements. Additionally, 29 Si MAS NMR, 27 Al MAS NMR, 1 H MAS NMR measurements as well as selected 1 H- 29 Si NMR Redor and 1 H- 27 Al NMR Trapdor experiments were performed. ...
Article
Full-text available
Natural zeolite clinoptilolite CLIN with a framework ratio of Si/Al ≥ 4 containing mainly potassium and calcium ions in its internal channel system was used as a starting material. The acidic HCLIN catalysts were prepared under soft conditions avoiding the use of environmental less-benign mineral acids. The starting material was ion exchanged using a 0.2 M aqueous ammonium nitrate solution at a temperature 80 °C for 2 h. The obtained NH4CLIN was converted into the acid HCLIN catalyst by calcination at 300–600 °C. The obtained samples were characterized by XRD, FTIR, SEM/TEM, AAS, and EDX element mapping. The state of aluminium and silicon was studied by 27Al- and 29SiMAS NMR spectroscopy. The textural properties of the catalysts were investigated by nitrogen adsorption and desorption measurements. The Brønsted acidity of the HCLIN catalysts was studied by temperature-programmed decomposition of the exchanged ammonium ions releasing ammonia as well as 1H MAS NMR, {1H–27Al} Trapdor, and {1H–27Al} Redor experiments. The strongly agglomerated samples were crystalline and thermally stable up to >500 °C. Although a part of the clinoptilolite framework is maintained up to 600 °C, a loss of crystallinity is already observed starting from 450 °C. The specific surface areas of the starting CLIN and ammonium exchanged NH4CLIN are low with ca. 26 m2/g. The pores are nearly blocked by the exchangeable cations located in the zeolite pores. The thermal decomposition of the ammonium ions by calcination at 400 °C causes an opening of the pore entrances and a markable increase in the specific micropore area and micropore volume to ca. 163 m2/g and 0.07 cm3/g, respectively. It decreases with further rising calcination temperature indicating some structural loss. The catalysts show a broad distribution of Brønsted acid sites (BS) ranging from weak to strong sites as indicated the thermal decomposition of exchanged ammonium ions (TPDA). The ammonium ion decomposition leaving BS, i.e., H+ located at Al–O–Si framework bridges, starts at ≥250 °C. A part of the Brønsted sites is lost after calcination specifically at 500 °C. It is related to the formation of penta-coordinated aluminium at the expense of tetrahedral framework aluminium. The Brønsted sites are partially recreated after repeated ammonium ion exchange. The catalytic performance of the acidic HCLIN catalysts was tested in the etherification of glycerol as a green renewable resource with different C1-C4 alcohols. The catalysts are highly active in the etherification of glycerol, especially with alcohols containing the branched, tertiary alkyl groups. Highest activity is observed with the soft activated catalyst HCLIN300 (300 °C, temperature holding time: 1 min). A total of 78% conversion of glycerol to mono and di ether were achieved with tert-butanol at 140 °C after 4 h of reaction. The mono- and di-ether selectivity were 75% and 25%, respectively. The catalyst can be reused.
... The rate at which the etherification proceeds is significantly influenced by factors including the acidic strength and porousness of the catalytic material. Previous literature describes the use of both homogeneous and heterogeneous catalyst to produce glycerol ethers by glycerol etherification [12,13]. Although homogeneous catalyst yields higher conversion, solid acid catalysts are preferred to avoid neutralization steps required to glycerol ethers after reaction [14]. ...
Article
Full-text available
Commercial biodiesel production is becoming popular due to its sustainability and reduced greenhouse gas emissions in comparison to fossil fuels. Biodiesel production involves the co-generation of crude glycerol that is not directly viable. Valorization of glycerol is essential from techno-economic perspective to achieve goals of circular economy. In this study, the glycerol etherification was carried out using tert-butyl alcohol in presence of dodecatungstophosphoric/silica (DTP/SiO2) catalyst for production of tert-butyl glycerol ethers that can be used as fuel additives. A series of DTP/SiO2 catalysts were prepared with various compositions by impregnating heteropoly acids (HPA, H3PW12O40). The product analysis was performed to monitor mono-, di-, and tri-tert-butyl glycerol ethers. The kinetic studies were conducted based on Langmuir–Hinshelwood model. The rate constants were determined from experimental data by regression analysis. The catalysts were characterized by X-ray diffraction, SEM/EDX, and thermogravimetric analysis. Varying DTP loadings resulted in different catalytic activities. Activation energy (38.23 kJ/mol), activation enthalpy (25.94 kJ/mol), activation entropy (− 163.8 J/mol/K), and Gibbs free energy of activation (88.67 kJ/mol) were attained for DTP/SiO2 (20%) catalyst that exhibited the best selectivity for di-tert-butyl glycerol ethers. Graphical Abstract
... The reaction selectivity gives frequently mono-(i.e., MTBG), di-(i.e., DTBG), and tri-(i.e., TTBG) tert-butyl glycerol ethers as main products (Di Serio et al., 2010). While MTBG has been used as a surfactant, the most functionalized products h-GTBE (i.e., DTBG and TTBG) are potentially blendable bioadditives to the liquid fuels, replacing the traditional oxygenate additives such as MTBE (i.e., methyl tert-butyl ether) (Vlad et al., 2012). Particularly, h-GTBE bioadditives increase the cetane number, reduce the emission of particulate matter and the emission of greenhouse effect gases (Huang and Kim, 2015). ...
Article
In this work, phosphomolybdic acid salts were synthesized exchanging their protons by the Sn(II) cations (i.e., H3-xSnx/2PMo12O40; x = 0.0, 1.0, 2.0, 3.0), and evaluated in the etherification reactions of glycerol with tert-butyl alcohol (TBA). The catalytic activity of the phosphomolybdate salts containing a variable load of Sn(II) ions was further compared to phosphotungstic and silicotungstic acid salts (i.e., H3-xSnx/2PW12O40; x = 0.0, 1.0, 2.0, 3.0 and H4-xSnx/2PW12O40; x = 0.0, 1.0, 2.0, 4.0, respectively). The integrity of the Keggin anion after the synthesis was confirmed by infrared spectroscopy analysis. Physical properties were determined through thermal analyses, powder X-Rays diffraction patterns, N2 adsorption-desorption isotherms, and scanning electron microscopy. The acidity properties were analyzed by adsorbed pyridine FT-IR spectroscopy, NH3-programmed-temperature desorption, and potentiometric titration. Among all the Sn(II) heteropoly salts tested, the Sn3PMo12O40 was the most active and selective catalyst toward glyceryl ethers. The highest activity of Sn3PMo12O40 was assigned to the greater Lewis acidity strength. Modeling, simulation, and economic analysis allowed to evaluate the viability of the process. An annual capacity of 30,000 metric tons was developed, and the process is feasible with a return on investment (ROI) of 30.7%.
... Also, these ethers can improve fuel quality and reduce emissions [20,28,29]. Glycerol ethers could also be substitute MTBE (methyl tertbutyl ether) and ETBE (ethyl tert-butyl ether) that lead to contamination of underground water [30]. ...
Article
Full-text available
In this study, the effects of glycerol ethers as a biobased additive on some important fuel properties of the diesel-bio-diesel blend are investigated. For this, firstly, glycerol ethers are synthesized by the etherification reaction of glycerol with tert-butyl alcohol with the presence of Amberlyst-15 as an acidic catalyst. The process of synthesis is followed by analyzing product composition through Fourier-transform infrared spectroscopy and gas chromatography-mass spectroscopy. Lastly, synthesized glycerol tert-butyl ethers are mixed with a diesel-biodiesel blend, and the effects of glycerol ethers on the fuel properties are determined. The analysis results show that glycerol tert-butyl ether mixture consists of mono-and di-ethers of glycerol. Due to the dehydration reaction and hindrance effect, the tri-ether of glycerol is not generated. The outcomes of determination of fuel properties reveal that glycerol ethers can be blended with a biodiesel-diesel mixture without any significant change in the fuel properties by fully meeting the ASTM D7467 standard, indicating glycerol ethers could serve as a biobased additive to diesel-biodiesel blend. Full-text share link: https://rdcu.be/b6RWI
... There is a big plenty of the recent papers devoted to the glycerol ethers synthesis (via the glycerol etherification with the olefins or the corresponding alcohols) and, as well ones devoted to the synthesis of the glycerol ketals, which had been summed up in detail in the recent reviews [7,8,15]. A strict number of technological concepts have been being lately developed for the syntheses of solketal (the cyclic ketal formed from glycerol and acetone) [16][17][18][19][20][21] and of glycerol tert-butyl ethers (GTBEs) [22][23][24] (the synthetic routes from glycerol to GTBEs are here presented at the Scheme 1). At the same time, there is not enough information about the characteristics of these products. ...
Article
The main physico-chemical features, as well as gasoline blending features of eight diol derivatives (1,3-diisopropyl and 1,3-di-tert-butyl glycerol ethers – di-GIPE and di-GTBE; O-isopropyl and O-tert-butyl solketal ethers – SIPE and STBE; isopropylidene ketals of ethylene and propylene glycol – DMD and TMD; cyclic ketals of glycerol with the methyl ethyl ketone and the methyl isobutyl ketone – GBK and GHK) have been estimated in order to reveal the relations between the molecular structures of the compounds and their efficiency as the gasoline oxygenated additives. For the oxygenates, the experimental values of the density, the boiling/melting point, the viscosity and the net heat of the combustion have been obtained. The characteristics of the estimated oxygenate-gasoline blends have involved their density, the calorific value, the fractional composition, the saturated vapor pressure, the research octane number and the motor octane numbers (ON), the oxidation stability, the gum content and the pour point. For di-GTBEs, as the most effective octane booster (among the compounds tested), the effect on the gasoline calorific value has been researched in the combination with the antiknock features improvement, which shows the evident advantages over ethanol and MTBE relatively to ΔAKI/ΔNHOC ratio.
Article
Biodiesel production continuously increases. This also increases the production of glycerol, the main by-product of transesterification, whose surplus is not fully absorbed by the market. An interesting destination for glycerol is the production of additives for the biofuel itself, as a green and economical alternative to conventional additives. The objective of this work is to simulate the production processes of three oxygenated fuel additives from the glycerol that is generated in the biodiesel production process: solketal, glycerol tert-butyl ethers (GTBE) and glycerol ethyl ethers (GEE). The processes were simulated in Aspen HYSYS® v8.8 and the capital and production costs were estimated. Among the evaluated alternatives, solketal production showed to be the most promising, since it is simpler than those of the other products and it works with acetone, which is cheaper than tert-butanol - required for GTBE production - and is needed in lower flowrates than ethanol - required for GEE production.
Article
A preliminary evaluation of the techno-economic feasibility of an industrial glycerol etherification process for bio-additives production has been carried out. The attention has been focused on the etherification reaction in presence of isobutylene (IB) or tert-butanol (TBA) in a batch reactor, catalyzed by an optimized heterogeneous system. Specifically, the Hyflon-based catalyst resulted more efficient and selective to desired products than commercial resin, limiting the competitive reactions of IB oligomerization and byproduct formation. In case of IB, a mixture of poly-ethers, suitable as oxygenated additives for diesel engines, containing very low amount of mono-tert-butylethers (<2 wt%) was obtained in a single step. This allowed a simplification of the process without the need of separation units and a lower production cost with respect to other reference processes. When TBA was used and water was formed as byproduct, a permeoselective membrane was exploited improving the glycerol conversion and enhancing the poly-ethers productivity by continuous water removal. In fact, very high poly-ethers yield (>85%) was achieved under optimized reaction conditions. The analysis of the whole process demonstrated the techno-economic sustainability of the proposed routes, leading to a production price of glycerol poly-ethers of 1,14 USD/kg and 2,38 USD/kg, by using IB or TBA respectively.
Article
Etherification of glycerol with alcohols/olefins is of great importance in the chemical processing industry for the valorization of biomass feedstocks. Herein, the functionalized mesoporous polymer (divinylbenzene-co-sodium p-stryene sulfonic acid/P-SO3H) catalysts are synthesized solvothermally and characterized by different techniques to understand the physico-chemical properties. Etherification of glycerol with tert-butyl alcohol/isobutylene was studied in the high pressure stirred autoclave using P-SO3H catalyst. The material exemplified the complete glycerol conversion with very high yield of desired product. The characterization data reveals that the mesoporosity, high surface area, tunable acidity and acid strength of the material are responsible for this good catalytic activity of the material. Minimal amount of P-SO3H catalyst is sufficient to achieve >88% yield of h-GTBE which is the best result ever reported for the etherification of glycerol with IB. The material is efficient and resistant towards leaching of active sites. The results achieved herein presented P-SO3H as an efficient catalyst which can be used for the large scale synthesis of potential additives for diesel.
Article
Full-text available
The aim of this research is the investigation of using glycerol tert-butyl ethers as a bio-renewable fuel additive to diesel-biodiesel blend in a diesel engine. In this context, an experimental study was carried out to synthesize glycerol tert-butyl ethers and investigate the effect of glycerol ethers on diesel engine performance and emissions. Glycerol tert-butyl ethers were synthesized by the etherification reaction of glycerol with tert-butyl alcohol (TBA) in a high-pressure stainless steel batch reactor using Amberlyst-15 as solid acidic catalyst. Produced glycerol ethers were mixed by 2 vol.% with a biodiesel-diesel blend to investigate the glycerol ethers usage as a fuel additive in a diesel engine. Three test fuels were considered for the experiments: (i) diesel fuel, (ii) B20 (80 vol.% diesel + 20 vol.% corn oil biodiesel); (iii) BD+GTBEs (80 vol.% diesel + 18 vol.% corn oil biodiesel + 2 vol.% glycerol tert-butyl ethers). Some basic fuel properties of test fuels were determined prior to engine tests. Engine performance and emissions tests were performed on a single-cylinder research diesel engine at four brake power output. The results revealed that the addition of glycerol tert-butyl ethers in the fuel showed no significant effect on determined fuel properties and caused a decrease in engine performance and exhaust emissions except for CO emissions when compared to base diesel fuel. However, BD+GTBEs presented similar performance and emissions characteristics with B20, as a consequence, glycerol tert-butyl ethers have the possibility of being utilized as a bio-renewable fuel additive to diesel-biodiesel blends. Published by Editorial Board Members of IJAET © This article is distributed by Turk Journal Park System under the CC 4.0 terms and conditions.
Article
Full-text available
In the presented work, liquid phase etherification reaction of glycerol (G) with tert-butyl alcohol (TBA) was investigated in a continuous flow reactor using Amberlyst-15 (A-15) as the catalyst. Some experiments were also repeated in a batch reactor at autogenous pressure. Effects of feed composition, reaction temperature (70–110 °C), and reaction pressure (1–5 bar) on the product distribution were investigated. When flow reactor pressure was reduced from 5 to 1 bar, enhanced glycerol conversion and isobutene formation were observed. Conversion of glycerol reached 66% at 110 °C, at a space time of 18 s·g·cm–3. However, batch reactor results indicated glycerol conversion values over 90% with diether selectivity of about 36% at much longer reaction times, over 400 min. Results indicated significance of equilibrium limitations, which initiated the idea of performing sorption enhanced reaction experiments in the flow reactor using Zeolite 4A and 5A as water adsorbents. Results of these sorption enhanced reactor experiments proved significant enhancement of glycerol conversion and diether yield by in situ removal of produced water during the etherification reaction. Among these two adsorbents, performance of Zeolite 4A was better. Diether selectivity increased from about 0.20 to values over 0.33 in the sorption enhanced reactor experiments performed with a zeolite/catalyst ratio of 4/1 at 90 °C.
Article
The development of distillation column control strategies involves not only the understanding of how to control columns as isolated unit operations but also the understanding of how its operation influences and is influenced by the other unit operations that comprise a chemical process. The role of distillation as a unit operation that separates components usually makes the operation of the distillation column central to the control of component inventories within a process. This can be contrasted with unit operations that do not change stream compositions such as heat exchangers and heat exchanger networks.
Article
This paper analyzes the nonlinear behavior of several recycle systems involving first- and second-order reactions. The results, presented in term of dimensionless numbers, explain some control difficulties reported by previous studies. It is shown that conventional control structures, fixing the flow rate of fresh reactants and relying on self-regulation, can lead to parametric sensitivity, unfeasibility, state multiplicity, or instability, particularly at low conversions. These problems can be solved by fixing the flow rate in the recycle loop, as stated by Luyben's rule. This paper demonstrates that a particular location for fixing the recycle flow rate is advantageous, namely, the reactor inlet. This decouples the reactor from the rest of the plant and avoids undesired phenomena due to mass recycles. For example, the unstable closed-loop behavior observed with nonisothermal PFRs disappears. The HDA plant case study illustrates the proposed strategy.
Book
Introduction ScopeEconomic IssuesTechnologyLarge-Scale Reactor Technology Efficient Heat TransferThe Mixing SystemsFast Initiation SystemsKinetics of Polymerization Simplified AnalysisMolecular-Weight Distribution Simplified AnalysisKinetic ConstantsReactor Design Mass BalanceMolecular-Weight DistributionHeat BalanceHeat-Transfer CoefficientsPhysical PropertiesGeometry of the ReactorThe Control SystemDesign of the Reactor Additional Cooling Capacity by Means of an External Heat ExchangerAdditional Cooling Capacity by Means of Higher Heat-Transfer CoefficientDesign of the JacketDynamic Simulation ResultsAdditional Cooling Capacity by Means of Water AdditionImproving the Controllability of the Reactor by Recipe ChangeConclusions References ScopeEconomic IssuesTechnology Efficient Heat TransferThe Mixing SystemsFast Initiation Systems Simplified Analysis Simplified Analysis Mass BalanceMolecular-Weight DistributionHeat BalanceHeat-Transfer CoefficientsPhysical PropertiesGeometry of the ReactorThe Control System Additional Cooling Capacity by Means of an External Heat ExchangerAdditional Cooling Capacity by Means of Higher Heat-Transfer CoefficientDesign of the JacketDynamic Simulation ResultsAdditional Cooling Capacity by Means of Water AdditionImproving the Controllability of the Reactor by Recipe Change
Article
Glycerol is a by-product of biodiesel production, for which new uses are being sought. Etherification of glycerol with isobutene in liquid phase with acidic ion exchange resin catalyst gave five product ethers and, as a side reaction, isobutene reacted to C8–C16 hydrocarbons. The effect of the reaction conditions on the system was studied and conditions for optimal selectivity toward ethers were discovered near with isobutene/glycerol molar ratio of 3 at 80 °C. The conditions controlling the distribution of the product ethers were studied and it was found that the extent of the etherification reaction and thus the main ether products can be changed by varying the reaction conditions.
Article
The excess of bioglycerol from the production of biodiesel finds no profitable use in the already saturated glycerol market. In this paper, the synthesis of 1,3-di-tert-butoxypropan-2-ol as a component for fossil diesel/biodiesel is reported. This new application could add value to the bioglycerol at the same time that it could enhance the properties and environmental performance of biodiesel. The synthesis consists in a three-step process that starts by the transformation of epichlorohydrin, obtained by the transformation of bioglycerol by the new Solvay process. The results of the overall process are satisfactory in terms of selectivity, conversion and yield. A spectroscopic analysis has been carried out, which shows that among the resulting products of the process, 1,3-di-tert-butoxypropan-2-ol is the most abundant and only little amounts of by products are formed, making the overall process technically viable. Also a study of the net carbon dioxide emissions produced by the combustion of this component has been developed.
Article
Triglycerides are reacted in a liquid phase reaction with methanol and a homogeneous basic catalyst. The reaction yields a spatially separated two phase result with an upper located non-polar phase consisting principally of non-polar methyl esters and a lower located phase consisting principally of glycerol and residual methyl esters. The glycerol phase is passed through a strong cationic ion exchanger to remove anions, resulting in a neutral product which is flashed to remove methanol and which is reacted with isobutylene in the presence of a strong acid catalyst to produce glycerol ethers. The glycerol ethers are then added back to the upper located methyl ethyl ester phase to provide an improved biodiesel fuel.
Article
The etherification of glycerol with tert-butyl alcohol (TBA) on solid acid catalysts provides a promising way to transform glycerol into a value added product to be used as an oxygenated additive to diesel fuels. Systematic optimization on this etherification reaction was carried out based on the small-scale experimental data (with a 35cm3 reacting mixture loaded in a 50cm3 stainless steel sampler shaken by a high rotational shaker). Effects of the possible processing variables were optimized via the proposed optimizing algorithm to reach the global optimum. The optimizing algorithm consists of the following two steps. Initially, the experiments designed by the global optimizer (rbfSolve routine in TOMLAB/CGO) are conducted before the region surrounding the global extreme is reached. When the region of global extreme is approaching, additional-optimizing experiments designed by the identified RBF model are then carried out to accelerate the rate to achieve the global optimum. The objective function was chosen as the total weight of the products glycerol mono-buthoxy ethers (GMBEs) and glycerol di-buthoxy ethers (GDBEs). Performing the designed experiments for the three control variables including the reacting temperature, the catalyst loading, and the solvent loading at the constant recipe and operating conditions, the global optimal operating condition was found using 24 experiments.
Article
Glycerine is a versatile educt mostly produced from renewable materials. It is formed as a byproduct in the production of biodiesel via transesterification. If the demand for biodiesel grows in the future, large quantities of glycerine could be available at low cost, thus giving sense to a use of glycerine for commodity-like products. A process for the technical production of "higher ethers" was designed and compared to an existing process by ARCO Chemical Technology. ARCO claims a process for producing glycerine ditertiary butyl ether that is restricted to a partial conversion to realize a phase separation after the reaction. The lower phase containing nonreacted glycerine is recycled to the reaction. The workup of the upper phase containing the ethers consists of washing all catalyst and monoether of the upper phase into the wastewater. Thus, 0.8 kg monoether/1 kg "higher ethers" are lost in 1.8 kg wastewater contaminated with organics. Compared to that, the new process allows to almost completely convert glycerine and isobutene into the desired "higher ethers". Thus, the fundamentals were developed to start up and optimize this process in a miniplant if the economical interest rises in the future.