Sunteți pe pagina 1din 9

Bioresource Technology 102 (2011) 490498

Contents lists available at ScienceDirect

Bioresource Technology
journal homepage: www.elsevier.com/locate/biortech

Integrated reactive absorption process for synthesis of fatty esters


Anton Alexandru Kiss a,, Costin Sorin Bildea b
a
b

AkzoNobel Research, Development & Innovation, Process Technology ECG, Velperweg 76, 6824 BM, Arnhem, The Netherlands
University Politehnica of Bucharest, Centre for Technology Transfer in Process Industries, Polizu 1-7, 011061 Bucharest, Romania

a r t i c l e

i n f o

Article history:
Received 6 May 2010
Received in revised form 13 August 2010
Accepted 22 August 2010
Available online 25 August 2010
Keywords:
Reactive absorption
Plantwide control
Energy integration
Green catalyst
Biofuel

a b s t r a c t
Reactive separations using green catalysts offer great opportunities for manufacturing fatty esters,
involved in specialty chemicals and biodiesel production. Integrating reaction and separation into one
unit provides key benets such as: simplied operation, no waste, reduced capital investment and low
operating costs.
This work presents a novel heat-integrated reactive absorption process that eliminates all conventional
catalyst related operations, efciently uses the raw materials and equipment, and considerably reduces
the energy requirements for biodiesel production 85% lower as compared to the base case. Rigorous
simulations based on experimental results were carried out using Aspen Plus and Dynamics. Despite
the high degree of integration, the process is well controllable using an efcient control structure proposed in this work. The main results are provided for a plant producing 10 ktpy fatty acid methyl esters
from methanol and waste vegetable oil with high free fatty acids content, using sulfated zirconia as solid
acid catalyst.
2010 Elsevier Ltd. All rights reserved.

1. Introduction
1.1. Background
Fatty esters are important chemicals used mainly in cosmetics,
pharmaceuticals, cleaning products and in the food industry. However, in the last decade the main interest has shifted to the large
scale production of biodiesel hence the stronger market drive
for more innovative and efcient processes. Biodiesel is a biodegradable and renewable alternative fuel with properties similar
to petroleum diesel (Bowman et al., 2006; Balat and Balat, 2008).
Typically, it is produced from green sources such as vegetable oils,
animal fat or even waste cooking-oils from the food industry (Encinar et al., 2005; Kulkarni and Dalai, 2006).
Nowadays, employing waste and non-edible raw materials is
mandatory to comply with the economical, ecological and also ethical requirements for biofuels. Basically, the food versus fuel competition can be easily avoided when the raw materials are waste
vegetable oils, non-food crops such as Jatropha (Kumar and Sharma, 2005) and Mahua (Puhan et al., 2005) or castor oil (Canoira
et al., 2010). However, waste raw materials can contain a substantial amount of free fatty acids (FFA), up to 100%. Consequently, an
efcient continuous biodiesel process is required that uses a solid
catalyst in order to suppress the costly downstream processing
steps and waste treatment.
Corresponding author. Tel.: +31 26 366 1714; fax: +31 26 366 5871.
E-mail addresses: Tony.Kiss@akzonobel.com, tonykiss@gmail.com (A.A. Kiss).
0960-8524/$ - see front matter 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.biortech.2010.08.066

Unlike petroleum diesel that is mixture of hydrocarbons, biodiesel consists of fatty acid methyl esters (FAME). Fatty esters are
currently produced by acid/base catalyzed trans-esterication, followed by several neutralization and product purication steps
(Hanna et al., 2005; Meher et al., 2006). Nevertheless, all conventional methods suffer from problems related to the use of liquid
catalysts, leading to severe economical and environmental penalties. Obviously, this has to change considering the impact of the
large scale production.
At present, the most common manufacturing technologies employ homogeneous catalysts, in batch or continuous processes
where both reaction and separation steps can create bottlenecks.
The literature overview reveals several processes currently in use
at pilot or industrial scale:
Batch processes allow better exibility with respect to composition of the feedstock. The trans-esterication is performed using an
acid or base catalyst, but the equipment productivity is low and
operating costs are high (Hanna et al., 2005; Lotero et al., 2005).
Continuous processes combine the esterication and trans-esterication steps, allowing higher productivity. Nonetheless, most of
these processes are still plagued by the disadvantages of using
homogeneous catalysts although solid catalysts emerged in the last
decade (Dale, 2003; Kiss et al., 2006a). Several reactive-distillation
processes were reported (He et al., 2006; Kiss et al., 2006b), along
with the ESTERFIP-H process developed by the French Institute of
Petroleum. The latter uses a solid catalyst based on Zn and Al
oxides, requiring high temperature (210250 C) and pressure
(3050 bar).

A.A. Kiss, C.S. Bildea / Bioresource Technology 102 (2011) 490498

Supercritical processes were developed to solve the problem of


oil-alcohol miscibility that hinders the kinetics of trans-esterication, as well as to take advantage of not using a catalyst at all. Still,
the operating conditions are quite severe (T > 240 C, p > 80 bar)
and therefore require special equipment (He et al., 2007).
Enzymatic processes have low energy requirements, as the reaction is carried out at mild conditions temperatures of 5055 C
and ambient pressure. However, due to the lower yields and long
reaction times the enzymatic processes cannot compete yet with
other industrial processes (van Gerpen, 2005; Demirbas and Balat,
2006; Demirbas, 2008).
Two-step processes are much simpler as the glycerides are
hydrolyzed rst to fatty acids that are then esteried in a second
step to fatty esters (Kusdiana and Saka, 2004; Minami and Saka,
2006). These processes are now very attractive and gain market
share due to obvious advantages: high purity glycerol is obtained
as by-product of the hydrolysis step, and the esterication step
can be performed using solid catalysts in an integrated reactiveseparation setup (Kiss et al., 2008; Kiss, 2009). The use of solid catalysts avoids the neutralization and washing steps, leading to a
simpler and more efcient process. Besides, when the raw materials consist mainly of FFA, only the esterication step is required.
The recent literature is quite abundant in studies on integrated
processes, such as reactive distillation (Kiss et al., 2006b; Kiss et al.,
2008) and reactive absorption (Noeres et al., 2003; Ghaemi et al.,
2009). However, that vast majority of the reported studies on reactive separation processes for biodiesel production are solely based
on conventional reactive distillation (He et al., 2006; Kiss et al.,
2006b; Kiss et al., 2008), or alternatives such as entrainer-based
reactive distillation (RD) and dual RD (Dimian et al., 2009).
This work proposes a novel energy-efcient reactive absorption
(RA) process for biodiesel production, which is very well controllable in spite of the high degree of integration. The integration of
reaction and separation into one unit corroborated with the use
of a heterogeneous catalyst offers major advantages such as: reduced capital investment, low operating costs, simplied downstream processing steps as well as no catalyst-related waste
streams and no soap formation. The main results are given for a
plant producing 10 ktpy biodiesel by esterication of methanol
with free fatty acids, using sulfated zirconia as solid acid catalyst.
1.2. Problem statement
Biodiesel is a very appealing but still expensive alternative fuel.
The common problem of all conventional processes is the use of liquid catalysts that require neutralization and costly separation
steps that generates salt waste streams (van Gerpen, 2005; Meher
et al., 2006; Narasimharao et al., 2007). To solve these problems, in
this work we make use of solid acids applied in an esterication
process based on catalytic reactive separation. Such an integrated
process is able to shift the chemical equilibrium to completion
and preserve the catalytic activity of the solid acid by continuously
removing the products (Kiss et al., 2006b). Moreover, compared to
conventional processes the investment and operating costs are
much reduced. However, the market pressure demands further decrease of the operating costs by replacing the raw materials with
inexpensive waste oils (high FFA content), and reducing the energy
requirements per ton of product (Janulis, 2004; Kiss, 2009). Several
reactive separation processes based on fatty acids esterication
were reported in the recent years, aiming at high performance
and productivity, along with low energy requirements:
 Reactive distillation: 191.2 kW h/ton biodiesel (Kiss et al.,
2008).
 Dual reactive distillation: 166.8 kW h/ton biodiesel (Dimian
et al., 2009).

491

 Reactive absorption: 138.4 kW h/ton biodiesel (Kiss, 2009).


 Heat-integrated reactive distillation: 108.8 kW h/ton biodiesel
(Kiss, 2008).
This work adds heat-integration to the reactive absorption process, thus allowing a major reduction of the energy requirements:
from 138.4 down to 21.6 kW h/ton biodiesel, which is equivalent
to only 15% of the reference RA process previously reported. Considering the high degree of integration, the controllability of the
process is also carefully investigated.
2. Methods
In this work, we use rigorous computer simulations to evaluate
the integrated process for synthesis of fatty esters by reactive
absorption. Starting from a base case design described in previous
work, we apply pinch analysis in order to identify the possibilities
of heat recovery by internal process/process exchange, as well as
the optimal selection and placement of utilities. Then, steady state
process simulation (Dimian, 2003) is employed to obtain a description of material and energy streams in the process with the scope
of evaluating the plant and understanding its steady state behaviour. Finally, dynamic simulations are used to evaluate controllability of the plant and to demonstrate the effectiveness of a
proposed plantwide control strategy.
2.1. Process integration
The conceptual design of the process is based on a reactive
absorption column (RAC) that integrates the reaction and separation steps into one operating unit. The chemical equilibrium is
shifted towards products formation by continuous removal of the
reaction products, instead of using an excess of a reactant typically the alcohol. An additional ash vessel and a decanter are used
to guarantee the high purity of both products. Since methanol and
water are much more volatile than the fatty ester or fatty acid,
these will separate without problems in the top of the column.
Fig. 1a presents the owsheet of this process based on conventional reactive absorption, as reported by Kiss (2009).
The column has 15 theoretical stages with the reactive zone located between stages 312, the solid catalyst loading being 6.5 kg
per stage. The fatty acid is pre-heated then fed as hot liquid in the
top of the reactive column while a stoichiometric amount of alcohol is injected as vapor into the bottom of the column, thus creating a counter-current ow regime over the reactive zone. Water
by-product is removed as top vapor, then condensed and separated
in a decanter from which the fatty acids are recycled back to the
column while water by-product is recovered at high purity. The
fatty esters are delivered as high purity bottom product of the reactive column. The hot product is ashed rst to remove the remaining methanol, and then it is cooled down and stored.
The reference owsheet presented in Fig. 1a is relatively simple,
with just a few operating units, two cold streams that need to be
pre-heated (fatty acid and alcohol) and two hot streams that have
to be cooled down (top water and bottom fatty esters). Therefore
the heat-integration was performed by applying previously reported heuristic rules (Hamed et al., 1996; Chen and Yu, 2003).
Consequently, two feed-efuent heat exchangers (FEHE) should replace partially or totally each of the two heat exchangers HEX1 and
HEX2. Fig. 1b illustrates the improved process including heat-integration around the reactive absorption column. The top vapor
stream is used to pre-heat the fatty acid feed stream (FEHE1). An
additional heat exchanger (HEX1) is required to nally heat the
fatty acids to the desired reaction temperature. When changes of
the production rate are desired, the duty of this heat exchanger

492

A.A. Kiss, C.S. Bildea / Bioresource Technology 102 (2011) 490498

Fig. 2 shows the composite curves and the yearly cost as function of the minimum delta T. This conrms the heuristics ndings,
as there is just a minimal need for an external hot utility to preheat the fatty acid feed stream.
2.2. Process simulation

Process simulations embedding experimental results were performed using Aspen Plus (AspenTech, 2009). The RA column was
simulated using the rigorous RADFRAC unit with RateSep (ratebased) model enabled, and explicitly considering three phase balances. The physical properties required for the simulation and
the binary interaction parameters for the methanolwater and
acid-ester pairs were available in the Aspen Plus database of pure
components, while the other interaction parameters were estimated using the UNIFAC Dortmund modied group contribution
method (AspenTech, 2009). Using other state-of-the art estimation
methods, such as UNIFAC and UNIFAC Lyngby modied, leads to
similar results.
As phase splitting may occur (Kiss, 2009), the RA column is
modeled using VLLE data. Phase splitting must be accounted for
as the free water phase can deactivate the solid acid catalyst. Nevertheless, as revealed latter by the column composition proles,
the liquid molar fraction of water does not exceed 0.1 on any given
stage. Therefore, the catalyst deactivation does not occur here, under the designed process conditions.
The fatty components were conveniently lumped into one
fatty acid and its fatty ester according to the reaction:
R  COOH CH3 OH $ R  COO  CH3 H2 O. Dodecanoic (lauric)
acid/ester was selected as lumped component due to the availability of experimental results, kinetics and VLLE parameters for this

Fig. 1. Synthesis of fatty esters by reactive absorption base case owsheet (a) and
heat-integrated process (b).

can be changed to ensure the required temperature of the columninlet acid stream (F-ACID). The hot liquid product of the FLASH, a
mixture of fatty esters, is used to pre-heat and vaporize the alcohol
feed stream. If changes of the production are expected, the nominal
design should include a bypass of the hot stream (dashed line in
Fig. 1b) which can be used for control purposes.

Fig. 2. Composite curve (a) and yearly cost versus minimum delta T (b).

493

A.A. Kiss, C.S. Bildea / Bioresource Technology 102 (2011) 490498

system (Kiss et al., 2006a; Dimian et al., 2009). Moreover, as lauric


acid is among the lightest fatty acids, this represents the worst case
scenario in terms of separation from alcohol. The assumption of
lumping components is very reasonable since fatty acids and their
corresponding fatty esters have similar properties. This approach
was already reported to be successfully used to simulate other
fatty esters production processes (Kiss et al., 2006b; Kiss et al.,
2008; Dimian et al., 2009; Kiss, 2009).
In this work, sulfated zirconia is considered as a solid acid catalyst, as kinetic data for the esterication with methanol is available from previous work (Kiss et al., 2008; Dimian et al., 2009;
Kiss, 2009). The esterication reaction is reversible, hence the reaction rate accounts for both direct and reverse reactions:

r k1 W cat C Acid C Alcohol  k2 W cat C Ester C Water

Table 2
Design parameters for simulating the reactive absorption column.
Parameter

Value

Units

Total number of theoretical stages


Number of reactive stages
Column diameter
HETP
Valid phases
Volume liquid holdup per stage
Mass catalyst per stage
Catalyst bulk density
Fatty acid conversion
Fatty acid feed (liquid, at 160 C)
Methanol feed (vapor, at 65 C)
Production of biodiesel (FAME)
RA column productivity

15
10 (from 3 to 12)
0.4
0.6
VLL
18
6.5
1050
>99.99
1167
188
1250
19.2

m
m

L
kg
kg/m3
%
kg/h
kg/h
kg/h
kg FAME/kg cat/h

where k1 and k2 are the kinetic constants for the direct (esterication) and reverse (hydrolysis) reactions, Wcat is the weight amount
of catalyst, and CComponent is the molar concentration of the components present in the system. Note that by changing the weight
amount of catalyst used (Wcat) the reaction rate can be similar for
different catalysts. As water is continuously removed from the system, the reverse hydrolysis reaction is extremely slow hence the
second term of the reaction rate can be neglected. Therefore, a simple kinetic expression can be used:

r k1 W cat C Acid C Alcohol k C Acid C Alcohol

k A exp Ea =RT

where CAcid and CAlcohol are mass concentrations, A is the Arrhenius


factor (A = 250 kmol m3 kg2 s1) and Ea is the activation energy
(Ea = 55 kJ/mol).
3. Results and discussion
3.1. Heat integrated design
The integrated reactive absorption process was designed
according to previously reported process synthesis methods for
reactive separations (Noeres et al., 2003). The next simulation results are given for a plant producing 10 ktpy (1250 kg/h) fatty acid

methyl esters (FAME) from methanol and waste vegetable oil with
highest free fatty acids (FFA) content, using solid acids as green
catalysts. Table 1 presents the complete mass balance of the process, while Table 2 lists the main design parameters, such as column size, catalyst loading, and feed condition.
High conversion of the reactants is achieved, with the productivity of the RA unit exceeding 19 kg fatty ester/kg catalyst/h. The
purity specication is higher than 99.9%wt for the nal biodiesel
product (FAME stream). Note that the total amount of the optional
recycle streams (REC-BTM) is not signicant representing less than
0.9% of the biodiesel production rate.
The liquid and vapor composition proles along the reactive
absorption column are plotted in Fig. 3a and b, respectively. The
concentration of fatty ester and methanol increases from the top
to bottom, while the concentration of the fatty acid and water increases from the bottom to the top of the column. Therefore, the
top product of the reactive separation column is water vapors with
small amounts of fatty acids, while the bottom liquid stream is the
fatty esters product (biodiesel) with a limited amount of methanol.
In particular cases, the concentration of methanol in the liquid
phase could be further increased by working at higher pressure.
Fig. 3c plots the temperature and reaction rate proles in the
reactive column. The RA column is operated in the temperature
range of 135165 C, at ambient pressure. The reaction rate exhibits a maximum in the middle, where the reactive zone is placed.

Table 1
Mass balance of a 10 ktpy FAME process based on integrated reactive-absorption.

Temperature/(C)
Pressure/(bar)
Vapor fraction
Mole Flow/(kmol/h)
Mass ow/(kg/h)
Mass ow/(kg/h)
Methanol
Acid
Water
Fame
Mass fraction
Methanol
Acid
Water
Fame
Mole fraction
Methanol
Acid
Water
FAME

F-ACID

F-ALCO

BTM

REC-BTM

REC-TOP

TOP

WATER

FAME

160
1.05
0
5.824
1166.7

65.4
1.05
1
5.876
188.3

136.2
1.03
0
6.125
1261.3

146.2
1.216
1
0.252
11.3

51.8
1
0
0.059
9.369

162.1
1
1
5.886
114.4

51.8
1
0
5.828
105.06

30
0.203
0
5.873
1250

0
116.74
0
0

188.3
0
0
0

9.125
Trace
Trace
1252.2

7.544
Trace
Trace
3.764

0.002
9.218
0.24
0.846

0.103
9.233
105.2
0.846

0.101
0.016
104.93
Trace

1.581
Trace
Trace
1248.4

0
1
0
0

1
0
0
0

0.007
Trace
Trace
0.993

0.667
Trace
10 ppb
0.333

172 ppm
0.894
0.023
0.082

894 ppm
0.08
0.912
0.007

965 ppm
148 ppm
0.999
513 ppb

0.001
Trace
Trace
0.999

0
1
0
0

1
0
0
0

0.046
Trace
Trace
0.954

0.931
Trace
26 ppb
0.069

873 ppm
0.726
0.211
0.062

546 ppm
0.008
0.992
670 ppm

0.001
13 ppm
0.999
43 ppb

0.008
Trace
Trace
0.992

A.A. Kiss, C.S. Bildea / Bioresource Technology 102 (2011) 490498

Molar fraction (liq)


0

0.2

0.6

0.8

Water

3
Stage

0.4

100%

Purity / [wt%]

494

90%

Acid

Water

FAME

80%
70%

9
12

Methanol

60%
0.8

Ester

0.9

1.1

1.2

Molar ratio alcohol:acid / [-]


0.2

0.4

0.6

0.8

Molar fraction (liq)


Molar fraction (vap)
0

0.4

0.6

0.8

Acid

3
Stage

0.2

Water

80

FEHE2

0.6

Tacid = 180 C

40

0.4
160 C

20

0.2

140 C

HEX1

0
0.8

0
0.9

1.1

1.2

Molar ratio alcohol:acid / [-]

c
Ester

0.2

Methanol

0.4

0.6

0.8

Molar fraction (vap)


Temperature / C
130
0

140

150

160

170

Product rate / kmol/h]

12

1
0.8

140 C

60

15

Split
160 C

FEHE1
180 C

0.5

Recycle Bottom

140 C

0.4

Water
160 C

0.3

180 C

0.2

6
FAME

180 C

0.1

160 C

Recycle Top

140 C

4
0.8

0
0.9

1.1

Recycle rate / [kmol/h]

100

Split fraction

Heat duty / [kW]

15

1.2

Stage

Molar ratio alcohol:acid / [-]


6
9

Fig. 4. Purity of top (water) and bottom (FAME) products (a), duty of heat
exchangers (b), product and recycle ow rates (c) versus molar reactants ratio
(alcohol:acid).

Reaction rate

12
Temperature

15

0.5
1 1.5 2 2.5
Ester formation / kmol/hr

Fig. 3. Composition, temperature and reaction rate proles along the column.

3.2. Sensitivity analysis


In spite of the recent progress in understanding the feasibility,
design and control of reactive separations, the conceptual design
of RA may lead to several different process congurations and
operating parameters. In this work, sensitivity analysis was used
to evaluate the range of the operating parameters: reactants ratio,
feed streams temperature, decanting temperature, ashing pressure, and recycle rates. The most signicant results of the sensitivity analysis are presented in Fig. 4, for a constant acid ow rate of

1166.7 kg/h (5.824 kmol/h) and variable alcohol ow rate. The


optimal molar ratio of the reactants (alcohol:acid) is very close to
the stoichiometric value of one, as illustrated by Fig. 4a.
If this ratio is higher than one (i.e. excess of alcohol) then the
fatty acid is completely converted to fatty ester, but the excess of
alcohol becomes a signicant impurity in the top stream and thereafter in the water by-product. On the contrary, when the ratio is
less than one (i.e. excess of fatty acids) then the purity of water
by-product remains high, but the conversion of fatty acid is incomplete hence the bottom product contains unreacted fatty acid
that cannot be easily removed from the nal product by simple
ashing. Since the separation of fatty acids from fatty esters is
more difcult than the separation of fatty acids from water, this
situation should be avoided. In practice, using a very small excess
of methanol (up to 1%) or an efcient control structure that can ensure the stoichiometric ratio of reactants is sufcient for the complete conversion of the fatty acids and prevention of the difcult
separation previously mentioned. Note that integrated reactive

A.A. Kiss, C.S. Bildea / Bioresource Technology 102 (2011) 490498


Table 3
Comparison between integrated reactive-absorption vs reactive-distillation processes
(at a production rate of 1250 kg/h or 10 ktpy fatty esters).
Equipment/parameter/units

RD

HIRD

RA

HIRA

Reactive column reboiler duty (heater), kW


HEX-1 heat duty (fatty acid heater), kW
HEX-2 heat duty (methanol heater), kW
Reactive column condenser duty (cooler),
kW
HEX-3 water cooler/decanter, kW
COOLER heat duty (biodiesel cooler), kW
FLASH heat duty (methanol recovery), kW
Compressor power (electricity), kW
Reactive column, number of reactive stages
Feed stage number, for acid/alcohol streams
Reactive column diameter, m
Reux ratio (mass ratio R/D), kg/kg
Boil-up ratio (mass ratio V/B), kg/kg
Productivity, kg ester/kg catalyst/h
Energy requirements per ton biodiesel,
kWh/ton FAME
Steam consumption, kg steam/ton FAME

136
95
8
72

136
0
0
72

n/a
108
65
n/a

n/a
27
0
n/a

6
141
0
0.6
10
3/10
0.4
0.10
0.12
20.4
191.2

6
38
0
0.6
10
3/10
0.4
0.10
0.12
20.4
108.8

77
78
0
0.6
10
1/15
0.4
n/a
n/a
19.2
138.4

0
14
0
0.6
10
1/15
0.4
n/a
n/a
19.2
21.6

295

168

214

34

separations can be easily controlled by PID controllers in a multiloop framework (Dimian et al., 2009; van Diggelen et al., 2010)
or by using more advanced techniques such as model predictive
control (Nagy et al., 2007).
Fig. 4b shows the duty of heat exchangers versus the molar
reactants ratio (alcohol:acid), for different temperatures of the acid
feed. Vaporization of an increasing ow rate of alcohol requires
more energy to be exchanged in FEHE2, which is realized by
increasing the amount of hot liquid that passes through the heat
exchanger. This can be accomplished if in the base case design
only a certain fraction of the hot stream (80% in our base case design, variable Split dashed line in Fig. 4b) is used for alcohol
vaporization. The fraction of hot liquid that is necessary for alcohol
vaporization increases to almost 100% when the alcohol ow rate

495

is increased to 120% of the base case value. The dependence of


FEHE2 heat duty versus the alcohol:acid ratio is insensitive with
respect to the temperature of the acid feed. Increasing the alcohol
ow rate leads to a larger ow rate at the top outlet of the column.
More energy is exchanged in FEHE1 with the result of lower duty
for the heater HEX1. In general, higher acid temperatures require
larger heat duties in both FEHE1 and HEX1.
Fig. 4c illustrates the ow rates of products and recycle streams
versus the molar reactants ratio (alcohol:acid), at different temperatures of the fatty acid feed stream. The ow rate of the FAME
stream is unchanged when the reactants ratio changes. However,
below the stoichiometric ratio of one, this stream contains large
quantities of unreacted acid. The ow rate of the WATER product
stream increases with the alcohol ow rate. Above the stoichiometric ratio of one, this stream contains water and the excess of
methanol. Higher temperature of the acid stream leads to increasing the top recycle because more acid is vaporized and reduces
the bottom recycle.
Table 3 shows a head-to-head comparison of the novel heatintegrated RA process proposed in this study against the previously
reported reference RD and RA processes (Kiss et al., 2008; Kiss,
2008, 2009). The heating and cooling requirements are gures that
ultimately translate into equipment size and cost. Remarkably, the
energy demand is less than 22 kW h/ton biodiesel or 34 kg steam
per ton of biodiesel produced. Also, compared to the reference
reactive absorption base case (Kiss, 2009), the heating and cooling
requirements are signicantly reduced by 85% and 90%,
respectively.
It should be noted that other heat-integration alternatives are
possible. For example, the top outlet stream could be used to
vaporize the methanol feed in FEHE1, while the liquid product
from the FLASH could be used to pre-heat the acid feed FEHE2. This
alternative has slightly higher energy requirements compared to
the owsheet presented in Fig. 1b (29 kW instead of 27 kW), while
the duty of the FAME cooler decreases to 3.3 kW. However, an
additional heat exchanger (13 kW) is necessary to cool the top

Fig. 5. Plantwide control structure: concentration of acid in bottom stream is controlled by manipulating the alcohol ow rate. The temperature of the column-inlet acid
stream is manipulated by a methanol concentration controller.

496

A.A. Kiss, C.S. Bildea / Bioresource Technology 102 (2011) 490498

stream to the decanting temperature (50 C), which is a drawback


of this alternative.
3.3. Plantwide control
Heat-integrated reactive absorption offers indeed signicant
advantages such as minimal capital investment and operating
costs, as well as no catalyst-related waste streams and no soap formation. However, the controllability of the process is just as important as the capital and operating costs savings. In processes based
on reactive distillation or absorption, feeding the reactants according to their stoichiometric ratio is essential to achieve high products purity (Dimian et al., 2009; Bildea and Kiss, in press). Thus,
the fatty acid is completely converted to fatty esters when there
is an excess of methanol, but the excess of methanol becomes an
impurity in the top stream and thereafter in the water by-product.
On the contrary, when there is an excess of fatty acids, the purity of
water by-product remains high, but the conversion of fatty acids is
incomplete. In the later case the bottom product contains
unreacted fatty acids that can not be easily removed from the nal
product by simple ashing. Since the separation of fatty acids from
fatty esters is more difcult than the separation of fatty acids from
water, this situation should be avoided. It is important to remark
that this constraint must be fullled not only during the
normal operation, but also during the transitory regimes arising
due to planned production rate changes or unexpected
disturbances.
However, the integrated biodiesel processes based on reactive
absorption have fewer degrees of freedom compared to reactive
distillation. This makes it challenging to correctly set the reactants
feed ratio and consequently avoiding impurities in the products. In
the following, we will prove that, in spite of the high degree of integration this reactive absorption process is very well controllable. A
key result of this study is an efcient control structure that can ensure the stoichiometric ratio of reactants and fullls the excess of
methanol operating constraint, which is sufcient for the total conversion of the fatty acids and for prevention of the difcult separations (e.g. fatty acid fatty ester).
Fig. 5 presents the proposed plantwide control structure. The
production rate is set by the ow rate of the acid stream (controller
FC1). After mixing with the recycle and passing the feed-efuent
heat exchanger FEHE1, the temperature of the acid fed to the column is controlled (by TC1) by manipulating the duty of the heat
exchanger HEX1. Compared to Fig. 1b, more details concerning
the practical implementation of the FEHE2 evaporator are given.
The ow rate of alcohol evaporated and fed to the column is set
by changing the split of the hot stream from the ash-outlet
stream (controller FC2). The temperature of the alcohol stream

entering the column is therefore the boiling temperature. Because


the hot FAME stream is not available during the startup of the
plant, the evaporator should include the option of using another
heat source. This is shown by the dashed drawing in Fig. 5. It
should be noted that setting the ow rates such that the stoichiometric ratio is fullled is not possible using only the ow controllers
FC1 and FC2 because of unavoidable measurement or control
implementation errors. For this reason, an additional concentration
controller is necessary. An excess of alcohol will have as result the
complete consumption of acid and a drop of the acid concentration
at the bottom of the column. On the other hand, large quantities of
acid will be present when this reactant is in excess. We conclude
that the imbalance in the alcohol:acid ratio can be detected by
measuring the concentration of acid in the bottom outlet of the
column. Therefore, the concentration controller CC2 is used to give,
in a cascade manner, the correct setpoint to the alcohol ow rate
controller FC2. In this way, when production rate changes are
implemented by changing the acid ow rate, the correct ratio between reactants is achieved. However, for large production rate
changes the overall reaction rate is not sufciently large to guarantee the complete conversion of both reactants and high purity cannot be achieved. In particular, large amounts of methanol will be
present in the top product of the absorption column. The concentration controller CC1 avoids this by increasing the setpoint of
the temperature controller TC1. The result is higher temperatures
inside the column, and therefore increased reaction rates. In
addition, the control of the material inventory is achieved by
conventional level and pressure control loops. Details of the controller action and controller tuning parameters are presented in
Table 4.
Fig. 6a and b depict the dynamic simulation results for the owsheet presented in Fig. 5, when the recycle of methanol vapors is
considered. The simulation starts from the steady state. At time
t = 1 h, the acid ow rate is increased by 10%, from 1166.7 to
1282 kg/h (5.824 to 6.4 kmol/h). Then, at time t = 5 h, the acid ow
rate is decreased to 1041.7 kg/h (5.2 kmol/h), representing 10%
decrease with respect to the nominal value. The new production
rate is achieved in about 1 h. The dynamic response of the water
product stream is slower. The purity of FAME remains high
throughout the dynamic regime, with the acid concentration below
the 2000 ppm requirement of the ASTM D675108 standard (i.e.
acid number less than 0.50 mg KOH/g biodiesel).
Similarly, Fig. 6c and d show the dynamic simulation results for
the same owsheet presented in Fig. 5, but without the bottom recycle of methanol vapors. The same scenario is assumed for the dynamic simulation. Again, the production rate responds quickly to
changes of acid ow rate (Fig. 6c). However, a degradation of dynamic performance is observed when the purities of the product

Table 4
Controller tuning parameters.
Controller

Control
action

Range of manipulated
variable

Range of controlled
variable

Setpoint

Bias

Gain
[%/%]

Integral time
[min]

Level LC1 (Buffer vessel)


Level LC2 (Dec., org. lq.)
Level LC3 (Dec., aq. lq.)
Level LC4 (Alc. evap.)
Level LC5 (Col. sump)
Level LC6 (Flash)
Pressure PC1 (Column)
Pressure PC2 (Flash)
Temp. TC1 (Acid inlet)
Mole fraction CC1 (Methanol, column top)
Flow rate FC2 (Methanol inlet)
Mole fraction CC2 (Acid, col. btm.)
Temperature TC2 (Flash)

Direct
Direct
Direct
Reverse
Direct
Direct
Direct
Reverse
Reverse
Direct
Reverse
Direct
Reverse

02350 kg/h
040 kg/h
0210 kg/h
0400 kg/h
02500 kg/h
02490 kg/h
012 kmol/h
01.22 kmol/h
00.2 GJ/h
140180 C
0.51
28 kmol/h
00.1 GJ/h

01 m
01 m
01 m
02
01.75 m
01 m
0.951.05 bar
0.180.22 bar
140180 C
15  103
48 kmol/h
12  103
130140 C

0.5 m
0.5 m
0.45 m
1
0.875 m
0.5 m
1 bar
0.2 bar
Output of CC1
3  103
Output of CC2
1  103
137 C

1180 kg/h
14.2 kg/h
105 kg/h
197 kg/h
1259 kg/h
1245
5.91 kmol/h
0.29 kmol/h
0.12 GJ/h
167.6 C
0.9
6.15 kmol/h
0.01 GJ/h

1
1
1
1
10
1
1
1
1
0.1
1
0.2
1

60

12
12
20
40
20
40
20

A.A. Kiss, C.S. Bildea / Bioresource Technology 102 (2011) 490498

Flow rates / [kg/h]

1500

FAME

1250

Acid

1000

300

4. Conclusions

250

The novel heat-integrated reactive absorption process proposed


in this work eliminates all conventional catalyst related operations,
improves efciency and considerably reduces the energy requirements for biodiesel production 85% lower as compared to the
base case. Another important result is an efcient control structure
that ensures the required reactants ratio and fullls the excess of
methanol operating constraint that is sufcient for the total conversion of the fatty acids and for prevention of difcult separations.
Remarkably, in spite of the high degree of integration this reactive
absorption process is very well controllable as illustrated by the results of rigorous dynamic simulations.

200
Alcohol

750
500

150
100

Water

250

50

0
0

497

0
10

Mass fraction / [-]

Time / [h]

0.98

2000

0.96

1500

0.94

1000
500

0.92
Acid in FAME

Mass fraction / [ppm]

FAME

Water

0.9

Acknowledgements

2500

0
10

Time / [h]

Flow rates / [kg/h]

1500

FAME

300

1250

Acid

250

1000

200
Alcohol

750

150

500

100
Water

250

50

0
0

10

1
0.96

3500

FAME

3000
2500

0.92

2000

Water

1500

0.88

1000
0.84
Acid in FAME

0.8
0

500

Mass fraction / [ppm]

d
Mass fraction / [-]

Time / [h]

0
10

Time / [h]
Fig. 6. Dynamic simulation results for the owsheet with (a, b) and without (c, d)
methanol recycle: Acid ow rate disturbance of +10% at 1 h, and 10% at 5 h.
Production rate changes are easily achieved and the products purity is maintained
at high values.

streams are taken into account (Fig. 6d). Therefore, the alternative
with methanol recycle is preferable.

C. S. Bildea acknowledges the support of CNCSIS UEFISCSU,


Project Number PNII IDEI 1545/2008 Advanced modelling
and simulation of catalytic distillation for biodiesel synthesis and
glycerol transformation. We also thank A. C. Dimian and G.
Rothenberg (University of Amsterdam), as well as A. B. de Haan
(Eindhoven University of Technology, The Netherlands) for the
helpful discussions.
References
Technology, Aspen., 2009. Aspen Plus: User guide volume 1 & 2. Burlington, MA.
Balat, M., Balat, H., 2008. A critical review of bio-diesel as a vehicular fuel. Energy
Convers. Manage. 49, 27272741.
Bildea, C.S., Kiss, A.A., 2010. Dynamics and control of a biodiesel process by reactive
absorption, Chem. Eng. Res. Des., doi: 10.1016/j.cherd.2010.05.007.
Bowman, M., Hilligoss, D., Rasmussen, S., Thomas, R., 2006. Biodiesel: a renewable
and biodegradable fuel. Hydrocarb. Process. 85, 103106.
Canoira, L., Galean, J.G., Alcantara, R., Lapuerta, M., Garcia-Contreras, R., 2010. Fatty
acid methyl esters (FAMEs) from castor oil: production process assessment and
synergistic effects in its properties. Renewable Energy 35, 208217.
Chen, Y.H., Yu, C.C., 2003. Design and control of heat-integrated reactors. Ind. Eng.
Chem. Res. 42, 27912808.
Dale, B., 2003. Greening the chemical industry: research and development priorities
for bio-based industrial products. J. Chem. Technol. Biotechnol. 78, 10931103.
Demirbas, M.F., Balat, M., 2006. Recent advances on the production and utilization
trends of bio-fuels. Energy. Convers. Manage. 47, 23712381.
Demirbas, A., 2008. Comparison of trans-esterication methods for production of
biodiesel from vegetable oils and fats. Energy Convers. Manage. 49, 125130.
Dimian, A.C., 2003. Integrated design and simulation of chemical processes.
Elsevier, Amsterdam.
Dimian, A.C., Bildea, C.S., Omota, F., Kiss, A.A., 2009. Innovative process for fatty acid
esters by dual reactive distillation. Comput. Chem. Eng. 33, 743750.
Encinar, J.M., Gonzalez, J.F., Rodriguez-Reinares, A., 2005. Biodiesel from used frying
oil. Variables affecting the yields and characteristics of the biodiesel. Ind. Eng.
Chem. Res. 44, 54915499.
Ghaemi, A., Shahhosseini, S., Maragheh, M.G., 2009. Nonequilibrium modelling of
reactive absorption processes. Chem. Eng. Commun. 196, 10761089.
Hamed, O.A., Aly, S., AbuKhousa, E., 1996. Heuristic approach for heat exchanger
networks. Int. J. Energy Res. 20, 797810.
Hanna, M.A., Isom, L., Campbell, J., 2005. Biodiesel: current perspectives and future.
J. Sci. Ind. Res. 64, 854857.
He, B.B., Singh, A.P., Thompson, J.C., 2006. A novel continuous-ow reactor using
reactive distillation for biodiesel production. Trans. ASAE 49, 107112.
He, H., Wang, T., Zhu, S., 2007. Continuous production of biodiesel fuel from
vegetable oil using supercritical methanol process. Fuel 86, 442447.
Janulis, P., 2004. Reduction of energy consumption in biodiesel fuel life cycle.
Renewable Energy 29, 861871.
Kiss, A.A., Dimian, A.C., Rothenberg, G., 2006a. Solid acid catalysts for biodiesel
production towards sustainable energy. Adv. Synth. Catal. 348, 7581.
Kiss, A.A., Rothenberg, G., Dimian, A.C., Omota, F., 2006b. The heterogeneous
advantage: biodiesel by catalytic reactive distillation. Top. Catal. 40, 141150.
Kiss, A.A., Dimian, A.C., Rothenberg, G., 2008. Biodiesel by reactive distillation
powered by metal oxides. Energy Fuels 22, 598604.
Kiss, A.A., 2009. Novel process for biodiesel by reactive absorption. Sep. Purif.
Technol. 69, 280287.
Kiss, A.A., 2008. Biodiesel production by heat-integrated reactive distillation. Comp.
Aided Chem. Eng. 25, 775780.
Kulkarni, M.G., Dalai, A.K., 2006. Waste cooking oil-an economical source for
biodiesel: a review. Ind. Eng. Chem. Res. 45, 29012913.
Kumar, N., Sharma, P.B., 2005. Jatropha Curcus a sustainable source for production
of biodiesel. J. Sci. Ind. Res. 64, 883889.

498

A.A. Kiss, C.S. Bildea / Bioresource Technology 102 (2011) 490498

Kusdiana, D., Saka, S., 2004. Two-step preparation for catalyst-free biodiesel fuel
production: hydrolysis and methyl esterication. Appl. Biochem. Biotechnol.
115, 781792.
Lotero, E., Liu, Y.J., Lopez, D.E., Suwannakarn, K., Bruce, D.A., Goodwin, J.G., 2005.
Synthesis of biodiesel via acid catalysis. Ind. Eng. Chem. Res. 44, 53535363.
Meher, L.C., Vidya Sagar, D., Naik, S., 2006. Technical aspects of biodiesel production by
trans-esterication: a review. Renewable Sustainable Energy Rev. 10, 248268.
Minami, E., Saka, S., 2006. Kinetics of hydrolysis and methyl esterication for
biodiesel production in two-step supercritical methanol process. Fuel 85, 2479
2483.
Nagy, Z.K., Klein, R., Kiss, A.A., Findeisen, R., 2007. Advanced control of a reactive
distillation column. Comp. Aided Chem. Eng. 24, 805810.

Narasimharao, K., Lee, A., Wilson, K., 2007. Catalysts in production of biodiesel: a
review. J. Biobased Mater. Bioenergy 1, 1930.
Noeres, C., Kenig, E.Y., Gorak, A., 2003. Modelling of reactive separation processes:
reactive absorption and reactive distillation. Chem. Eng. Process 42, 157178.
Puhan, S., Vedaraman, N., Rambrahamam, B.V., Nagarajan, G., 2005. Mahua
(Madhuca Indica) seed oil: a source of renewable energy in India. J. Sci. Ind.
Res. 64, 890896.
Van Diggelen, R.C., Kiss, A.A., Heemink, A.W., 2010. Comparison of control strategies
for dividing-wall columns. Ind. Eng. Chem. Res. 49, 288307.
Van Gerpen, J., 2005. Biodiesel processing and production. Fuel Process. Technol. 86,
10971107.

S-ar putea să vă placă și