Você está na página 1de 12

Bioresource Technology 111 (2012) 282293

Contents lists available at SciVerse ScienceDirect

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

Design and analysis of bioreneries based on raw glycerol: Addressing


the glycerol problem
John A. Posada, Luis E. Rincn, Carlos A. Cardona
Instituto de Biotecnologa y Agroindustria, Departamento de Ingeniera Qumica, Universidad Nacional de Colombia sede Manizales, Cra. 27 No. 64-60, Manizales, Colombia

a r t i c l e

i n f o

Article history:
Received 7 October 2011
Received in revised form 20 January 2012
Accepted 25 January 2012
Available online 6 February 2012
Keywords:
Glycerol conversion
Process design
Process simulation
Process assessment
Glycerol-based bioreneries

a b s t r a c t
Glycerol as a low-cost by-product of the biodiesel industry can be considered a renewable building block
for bioreneries. In this work, the conversion of raw glycerol to nine added-value products obtained by
chemical (syn-gas, acrolein, and 1,2-propanediol) or bio-chemical (ethanol, 1,3-propanediol, D-lactic acid,
succinic acid, propionic acid, and poly-3-hydroxybutyrate) routes were considered. The technological
schemes for these synthesis routes were designed, simulated, and economically assessed using Aspen
Plus and Aspen Icarus Process Evaluator, respectively. The techno-economic potential of a glycerol-based
biorenery system for the production of fuels, chemicals, and plastics was analyzed using the commercial
Commercial Sale Price/Production Cost ratio criteria, under different production scenarios. More income
can be earned from 1,3-propanediol and 1,2-propanediol production, while less income would be
obtained from hydrogen and succinic acid. This analysis may be useful mainly for biodiesel producers
since several protable alternatives are presented and discussed.
2012 Elsevier Ltd. All rights reserved.

1. Introduction
A key aspect of the manufacture of biodiesel is the co-production of glycerol, which is obtained at a weight ratio of 1/10 (glycerol/biodiesel). Currently, glycerol does no longer represent a
signicant benet for the biodiesel industry due to its low price
as consequence of the growing market of biodiesel. For example,
glycerol production increased 400% between 2004 and 2006, and
its price fell nearly 10-fold (Posada and Cardona, 2010a). Therefore,
economic exploitation of glycerol as raw material for its transformation to added-value products seems economically necessary.
In the present study, the production of nine added-value products
from glycerol by chemical or biochemical conversion routes were
analyzed based on techno-economic criteria. In addition, different
reaction conditions, strains and downstream processes were considered. The results obtained may be especially useful for biodiesel
producers since several protable transformations of raw glycerol
into added-value products are presented and discussed.
2. Chemical conversion of glycerol
Glycerol can be transformed to added-value products by
oxidation, reduction, decomposition, gasication and pyrolysis.
For example, the main oxygenated products from glycerol are glyceric acid, dihydroxyacetone, hydroxypyruvic acid, tartaric acid,
mesoxalic acid and oxalic acid, besides some intermediates (e.g.,
Corresponding author. Tel.: +57 6 8879300x50199; fax: +57 6 8879300x50452.
E-mail address: ccardonaal@unal.edu.co (C.A. Cardona).
0960-8524/$ - see front matter 2012 Elsevier Ltd. All rights reserved.
doi:10.1016/j.biortech.2012.01.151

glyceraldehyde, glycolic acid, and glyoxylic acid) (Posada, 2011).


Oxidation reactions have mostly been studied using palladium,
platinum and gold as catalysts, but palladium and platinum are
deactivated with the increment of the reaction time (Demirel-Glen
et al., 2005). Better catalytic performances have been achieved with
platinum and gold catalysts incorporating promoters (e.g., lead and
bismuth) to improve the oxidation of secondary alcohols and to
avoid the products degradation (Garcia et al., 1995). Glycerol reduction mainly generates 1,2-propylene glycol, 1,3-propylene glycol,
ethylene glycol, and other by-products (e.g., lactic acid, acetol, acroleine) besides the degradation products (e.g., propanol, methanol,
methane, and carbon dioxide) (Dasari et al., 2005a). This reaction
uses different catalysts such as copper, zinc, ruthenium, cobalt,
magnesium, molybdenum, nickel, palladium and platinum (Lahr
and Shanks, 2005), and a wide range of pressures (20005000 psi)
and temperatures (200350 C) have been reported (Casale and Gomez, 1994). However, the highest selectivities to propylene glycol
(the most important reduction product) have been reported for cupper-based catalysts, which also exhibit low selectivities for ethylene
glycol and other degradation by-products (Dasari et al., 2005b).
Gasication of glycerol consists of a thermochemical degradation process in the presence of gasication agents (i.e., air, pure
oxygen or steam) that produces syngas (main product) (Ahmed
et al., 2011). Syngas is a mixture of hydrogen and carbon monoxide
as main products, but methane, carbon dioxide and other residues
(e.g., tar, char and ash) are also generated. A wide range of conversions and selectivities have been reported depending not only on
the operational conditions such as temperature, pressure and glycerol concentration, but also on the pollutants presence into the raw

J.A. Posada et al. / Bioresource Technology 111 (2012) 282293

glycerol stream (e.g., methanol and KOH mainly) (Matsumura


et al., 2005). For example, an improvement in syngas production
was observed when high reaction temperatures were used but carbon dioxide production also increased (Valliyappan et al., 2008).
3. Biochemical conversion of glycerol
Ethanol, 1,3-propanediol, D-lactic acid, succinic acid, propionic
acid, and poly-3-hydroxybutyrate are some of the useful chemical
products that can be obtained by fermentation of glycerol (Hjek
and Skopal, 2010). 1,3-Propanediol production can be obtained

283

biologically by several bacterial strains, but Klebsiella pneumoniae


and Clostridium butyricum are the most commercially promising
strains because of their high yield, productivity, and resistance to
both substrate and product inhibition (Ya-Nan et al., 2006). 1,2Propanediol has also been recently synthesized from glycerol using
genetically engineered Escherichia coli (Clomburg and Gonzalez,
2011) at a product concentration of 5.6 g/L and a mass yield of
21.3. Ethanol production can also be carried out by E. coli and high
ethanol yields at a concentration of 10 g/L have been achieved
(Durnin et al., 2009; Murarka et al., 2008; Shams Yazdani and
Gonzalez, 2008). Polyhydroxybutyrate (PHB) can be produced from

Fig. 1. Process ow diagrams for raw glycerol purication ((a) purication up to 88 and 98 wt.%, and (b) purication up to 99.7 wt.%) and study cases IIII for chemical
conversion of glycerol ((c) dehydration of glycerol, (d) steam gasication of glycerol, (e) hydrogenolysis of glycerol). E, Evaporation column, R: Reactor, Cen: Centrifuge, D:
Decanter, DC: Distillation Column, IER: Ionic Exchange Resin, HE: Heat Exchanger, H: Heater, Con: Condenser, D: Divisor, M: Mixer, Comp: Compressor, Sep: Separator.

284

J.A. Posada et al. / Bioresource Technology 111 (2012) 282293

glycerol by a number of bacteria, Bacillus megaterium and


Cupriavidus necator being the most important ones (Cavalheiro
et al., 2009; Pablo et al., 2008; Posada et al., 2011b). D-Lactic acid
production from glycerol using engineered E. coli strains was recently reported (Mazumdar et al., 2010) and analyzed (Posada
et al., 2011a). Fermentative production of succinic acid from glycerol has been achieved using Actinobacillus succinogenes and recombinant E. coli (Blankschien et al., 2010), although the
production of by-products (e.g., acetic acid, formic acid, lactic acid,
and ethanol) limits the possibility of scaling up this process to an
industrial level since a more complex downstream processing
would be required (Kim et al., 2004). Finally, propionic acid can
be produced by propionibacteria via the dicarboxylic acid pathway
with acetic acid and succinic acid as by-products (Coral et al., 2008;
Zhang and Yang, 2009).

4. Methodology
Initially, 13 chemical routes and nine fermentative products were
identied as possible transformation ways for glycerol conversion to
added-value products. Based on technical, economic and environmental criteria (i.e., temperature and pressure of reaction, levels of
conversion/selectivity/productivity, requirements of energy, price
and market of the main product and wastes production), three
chemical and six fermentative products were chosen to be analyzed.
The processes design followed a previously described strategy (Cardona and Snchez, 2007; Quintero et al., 2008) which combines the
hierarchical decomposition methodology with the process design
method known as breadth-rst. This strategy allows systematic generation of alternatives that consider specic characteristics of each
process and simultaneous comparison of different process alternatives. In this way, it is possible to screen different process alternatives and to evaluate them at the next level of the hierarchical
decomposition using process simulators. Therefore, the process simulator Aspen Plus (Aspen Technology, Inc., USA) was used for dening, structuring, specifying, and simulating the technological
schemes for either chemical or biochemical conversion of glycerol
to added-value components. The most complex and detailed technological schemes were obtained by means of rigorous simulations,
which involved sensitivity analyses and search of optimal operation
conditions. Since not only pure, but also raw glycerol has been used
as feedstock, a typical composition of a glycerol-rich stream obtained during the biodiesel production from IdaGold mustard
(Thompson and He, 2006) was considered as the feedstock. Since this
raw glycerol stream contains low concentration of glycerol, a purication process was analyzed in order to obtain the three most important qualities of commercially available glycerol (crude glycerol at
88 wt.%, technical glycerol at 98 wt.% and USP glycerol at
99.7 wt.%). In addition, specic compounds not available on the Aspen Plus Database (i.e., free fatty acids, alkyl esters, proteins, salts,
cell mass strains, enzymes and other complex molecules produced
by reactive-extractive processes) were incorporated to the database
for each simulation. For components and properties not included in
the Aspen Plus database, subroutines and special software were designed to predict the component properties needed for simulation.
For conventional molecules, the group contribution method was
used, while biomass and enzyme were considered as solids and nonconventional compounds, respectively. All results from software calculations are presented using an accuracy value of 0.01 (0.0001 is
used recurrently for comparison purposes in some cases).
The reactive systems were analyzed as follows. For the highly
exothermic dehydration of glycerol to acrolein, an acid catalyzed
process was considered and the reaction scheme reported by
Tsukuda et al., (2007) was used. The gasication process of glycerol
was modeled according to the molar distribution of the reaction

products reported by (Mozaffarian et al., 2004). A two-steps process for the selective production of 1,2-propanediol from glycerol
was developed according to the model proposed by Akiyama et
al. (2009). Subsequently, for biochemical conversion of glycerol
(i.e., glycerol fermentation), six main products were considered.
Also, it is important to note that, although several possibilities
for glycerol fermentation to added-value products have been reported, just a few publications describe accurate kinetic models.
Consequently, for the production of ethanol, poly-3-hydroxybutyrate, lactic acid, succinic acid and propionic acid, a yield approach
was used. Moreover, a stoichiometric approach was considered as
a valid and relevant approach for analyzing the reaction stage of
different technological schemes.
With respect to the economic assessment, capital and operating
costs were calculated using the software, Aspen Icarus Process
Evaluator (Aspen Technologies, Inc., USA). The economic parameters considered were those from Colombia, in US dollars for a 10year period at an annual interest rate of 16%, considering the
straight line depreciation method with a 33% income tax (Posada
et al., 2010). The Colombian labor cost used for operatives and
supervisors was US$2.14/h and US$4.29/h, respectively. Also, the
prices used for electricity, water and low pressure vapor were
US$0.03044/kWh, US$1.252/m3 and US$8.18/ton, respectively.

5. Results and discussions


Since the feedstock for all the processes is a raw glycerol stream
obtained from a typical biodiesel plant, the rst step of consideration was purication. Fig. 1a. shows the simplied owsheet for
raw glycerol purication to obtain 88 wt.% (crude glycerol) and
98 wt.% (technical glycerol). Production of glycerol at 99.7 wt.%
(glycerol USP grade) required a further rening process using an
ion exchange resin which removes the triglycerides still contained
in the mixture, as shown in Fig. 1b (Posada and Cardona, 2010b). In
addition to the purication analysis, two different scenarios were
considered due to the high concentration of methanol (32.6 wt.%)
in the glycerol stream. In the rst scenario, the methanol obtained
during the evaporation stage was considered as a waste, while in
the second scenario this stream was recycled and reused for
transesterication to produce biodiesel (methanol at 99 wt.%). In
other words, methanol was considered as a by-product with an
economical value in the last scenario. Therefore, the purication
costs of raw glycerol up to 88 wt.% were 0.1574 US$/L (scenario
I) and 0.0984 US$/L (scenario II), and up to 98 wt.% were 0.1782

Table 1
Composition of fed raw glycerol, mass ow of puried glycerol and purication costs.
Materials (kg/h)
Raw glycerol (triglycerides 2%, methanol 32.59%,
glycerol 60.05%, NaOCH3 2.62%,
proteins 0.13%, fats 1.94%) (wt.%)
Products
Methanol
Glycerol at 88% (option 1)
Glycerol at 98% (option 2)
Glycerol at 99.7% (option 3)
Purication costs-Scenario I*
Glycerol at 88%
Glycerol at 98%
Glycerol at 99.7%
Purication costs-Scenario II**
Glycerol at 88%
Glycerol at 98%
Glycerol at 99.7%
*

Mass ow (kg/h)
973.30

Mass ow (kg/h)
301.98
665.25
596.60
586.18
(US$/L)
0.22
0.26
0.35
(US$/L)
0.16
0.20
0.28

Scenario I: Methanol is considered as a waste.


Scenario II: Methanol is considered as a purication by-product with economic
value.

**

285

J.A. Posada et al. / Bioresource Technology 111 (2012) 282293

US$/L (scenario I) and 0.1124 US$/L (scenario II). It is important to


note that these purication costs were always lower than their
commercial Commercial Sale Prices. In addition, when anhydrous
methanol at 99 wt.% is recovered, a reduction of 1926% in the
purication costs was obtained. The economic results showed that
the purication processes were protable for all glycerol grades
(Table 1).

commercial Commercial Sale Prices. In addition, when anhydrous


methanol at 99 wt.% is recovered, a reduction of 1926% in the
purication costs was obtained. The economic results showed that
the purication processes were protable for all glycerol grades
(Table 1).

5.1. Glycerol purication

For the chemical conversion of glycerol, dehydration, steam


gasication, and hydrogenolysis were considered and their respective products were acrolein, hydrogen, and 1,2-propanediol. These
three options were simulated using the processes conguration
shown in Fig. 1ce. A glycerol conversion of 100% was reached
for all processes, and the respective molar yields for acrolein,
hydrogen, and 1,2-propanediol were 85.2%, 78.2% and 79.97%
(Table 2). In addition, the rst two processes were heat-integrated,
recovering 175 and 67 W/(feeding kg) for dehydration and gasication processes, respectively.
The results obtained during the economic assessment of the
chemical conversion of glycerol are summarized in Fig. 2a and b.
The ratio, Commercial Sale Price/Total Production Cost, is shown in
Fig. 2a, while the main contributors to the production costs are
shown in Fig. 2b. Since the production of acrolein at 98.5 wt.% requires a powerful coolant system, two nal concentrations were
considered and the industrial grade (92 wt.%) was also included
in this analysis. During the acrolein production at 92 wt.%, most
of the production costs were represented by raw materials and services (close to 72% of the total production costs). For acrolein production at 98.5 wt.%, the service contributions were about 67% of
total production costs. Thus, the production process of acrolein at

Since the feedstock for all the processes is a raw glycerol stream
obtained from a typical biodiesel plant, the rst step of consideration was purication. Fig. 1a shows the simplied owsheet for
raw glycerol purication to obtain 88 wt.% (crude glycerol) and
98 wt.% (technical glycerol). Production of glycerol at 99.7 wt.%
(glycerol USP grade) required a further rening process using an
ion exchange resin which removes the triglycerides still contained
in the mixture, as shown in Fig. 1b (Posada and Cardona, 2010b). In
addition to the purication analysis, two different scenarios were
considered due to the high concentration of methanol (32.6 wt.%)
in the glycerol stream. In the rst scenario, the methanol obtained
during the evaporation stage was considered as a waste, while in
the second scenario this stream was recycled and reused for
transesterication to produce biodiesel (methanol at 99 wt.%). In
other words, methanol was considered as a by-product with an
economical value in the last scenario. Therefore, the purication
costs of raw glycerol up to 88 wt.% were 0.1574 US$/L (scenario
I) and 0.0984 US$/L (scenario II), and 98 wt.% were 0.1782 US$/L
(scenario I) and 0.1124 US$/L (scenario II). It is important to note
that these purication costs were always lower than their

5.2. Chemical conversion of glycerol, case study IIII

Table 2
Main processes streams and molar yields for cases of study IIV. Production of acrolein, hydrogen, 1,2-propanediol and D-lactic acid from glycerol with different qualities.
Case study
Scenario

I:
II:
III: 1,2IV: D-Lactic acid
Acrolein Hydrogen Propanediol
(1) Pure glycerol
diluted at 20 g/l

Main reactives
(kg/h)
Raw glycerol at
88 wt.%
Pure glycerol at
98 wt.%
100.0
Glycerol
directly
diluted at
10 wt.%
Water
Hydrogen

(3) Pure glycerol


diluted at 40 g/l

579

579

579

27,996

27,996

13,780

(4) Raw glycerol


diluted at 40 g/l

(5) Raw glycerol


diluted at 60 g/l

581

581

13,821

9041

100.0

30

23,333
374,765

Strain
Cell mass

LA01(pZSKLMgldA) LA02(pZSglpKglpD) LA02Ddld(pZSglpKglpD) LA02Ddld(pZSglpKglpD) LA02Ddld(pZSglpKglpD)


72.2
40.8
21.4
28.3
31.5

Residues (kg/h)
94
General
aqueous
residues
Main products (kg/h)
Acrolein at
5.2
98.5 wt.%
Acrolein at
5.6
92 wt.%
Hydrogen at
91 wt.%
Syngas

1,2Propanediol
at 99 wt.%

10000.0

1.0

20.7

7997

D-Lactic

acid at
99 wt.%

Molar yield

(2) Pure glycerol


diluted at 20 g/l

85.20%

78.20%

85.00%

27,531

27,536

13,308

13,333

8577

420

416

422

442

433

82.00%

81.20%

83.30%

85.90%

93.40%

286

J.A. Posada et al. / Bioresource Technology 111 (2012) 282293

Fig. 2. Commercial Sale Price/Production Cost ratio for study cases IVI ((a) study cases IIII, and (b) study cases IVVI) and share of product purication costs ((c) study cases
IIII, and (d) study cases IVVI).

high purity was not economically viable. The highest Commercial


Sale Price/Total Production Cost ratio was obtained for 1,2-propanediol (i.e., 1.57) indicating that this compound was able to generate
the highest economic return for the chemical conversion of
glycerol.
Although raw material cost represents usually around 50% of
the total production costs for chemical processes, in the case of
hydrogen production from glycerol the sum of raw material and
services costs were almost 36% (Fig. 2b). In addition, the main
investment cost was for process units, where the equipment
depreciation value was 45.14%. This amount is explained by the
strict and complex equipment requirements due to the extreme
operational conditions (i.e., high temperatures and pressures).
5.3. Biochemical conversion of glycerol, case study IVIX
5.3.1. Case study IVVI: carboxylic acids (D-lactic acid, succinic acid,
and propionic acid)
Biochemical production of carboxylic acids from glycerol (i.e.,
D-lactic acid, succinic acid, and propionic acid) was performed in
a whole technological scheme, composed of glycerol purication,
glycerol fermentation, and carboxilic acid recovery and purication. In each case, ve fermentation scenarios were simulated considering different strains, substrate concentrations, fermentation
times, and fermentation stages. The general considerations related
to the used strains and the molar yields to each carboxylic acid are
presented in Tables 2 and 3.
For D-lactic acid production ve scenarios using pure (98 wt.%)
or crude (88 wt.%) glycerol were considered. The glycerol consumptions were higher than 90%, with fermentation times between 36 and 72 h. The downstream process for D-lactic acid
recovery and purication from the fermentation broth was designed based on a reactive-extraction process using tri-n-octylamine and dichloromethane as extractant and active diluent,

respectively. The complete owsheet for D-lactic acid production


is shown in Fig. 3a. The nal production of D-lactic acid was directly
related not only to fermentation yield but also to substrate consumption. In other words, while the order with respect to yield
was: Scenario 5 > Scenario 4 > Scenario 3 > Scenario 1 > Scenario
2 (Table 3), the order with respect to the D-Lactic acid production
was: Scenario 4 > Scenario 5 > Scenario 3 > Scenario 1 > Scenario
2 (Table 2). The change between Scenarios 5 and 4 occurred due
to incomplete consumption of glycerol during the fermentation
in the Scenario 5.
For succinic acid production, glycerol consumption was between 58.5% and 96% using pure glycerol (98 wt.%). The downstream process used a reactive-extraction stage where complex
molecules were produced from the fermentation products (i.e.,
succinic acid, formic acid and acetic acid). Tri-n-octylamine (TOA)
and 1-octanol were used as extraction and diluent agents, respectively. The owsheet for succinic acid production is shown in
Fig. 3b. The main simulation results for each scenario are shown
in Table 2 and signicant differences for the nal production of
succinic acid can be observed among the ve analyzed scenarios.
The nal production of succinic acid depended mainly on the
succinic acid yield. For instance, while the order for succinic acid
yield was: Scenario 2 < Scenario 1 < Scenario 5 < Scenario 4 < Scenario 3 (Table 3), the order for succinic acid production was: Scenario 2 < Scenario 1 < Scenario 5 < Scenario 3 < Scenario 4 (Table
3). The switch of Scenarios 3 and 4 occurred because the glycerol
consumption difference was higher than the succinic acid yield.
For propionic acid production, ve fermentation scenarios were
analyzed (Table 3) and fed conditions were as follows: Scenarios 1,
2, 3 and 5 used pure glycerol at 20, 50, and 46, and 41 g/L, respectively; Scenarios 4 used crude glycerol at 17 g/L. For Scenarios 4
and 5, a brous-bed bioreactor packed with immobilized cells
was considered. The fermentation times in each case were: 120,
150, 280, 160, and 104 h for Scenarios 1, 2, 3, 4, 5, respectively.

Table 3
Main processes streams and molar yields for cases of study V and VI. Production of succinic acid and propionic acid from glycerol with different qualities.
Case study

V: Succinic acid

Scenario

(1) Raw glycerol


diluted at 20 g/l

VI: Propionic acid


(2) Raw glycerol
diluted at 9.6 g/l

(3) Raw glycerol


diluted at 8.3 g/l

(4) Raw glycerol


diluted at 9.1 g/l

(5) Raw glycerol


diluted at 6.5 g/l

(2) Raw glycerol


diluted at 50 g/l

(3) Raw glycerol


diluted at 46 g/l

(4) Raw glycerol


diluted at 17 g/l

(5) Raw glycerol


diluted at 41 g/l

579

579

579

579

579

27,996

11,008
358

11,988
838

12,037
904

13,493
533

Propionibacterium
acidipropionici ACKTet
29.6

Immobilized P,
acidipropionici
ACK-Tet
53.4

Main
reactives
(kg/h)
Raw
glycerol
at
88 wt.%
Pure
glycerol
at
98 wt.%
Water
Tri-n795
1-Octanol

579

579

579

27,996
octylamine

58,803

67,945

1369

2738

2555

3103

456

Strain

Escherichia coli

Mannheimia sp,
Pasteurellaceae

Mannheimia sp,
Pasteurellaceae

Mannheimia sp,
Pasteurellaceae

Anaerobiospirillum
succiniciproducens

Propionibacterium
acidipropionici

Propionibacterium
acidipropionici

Cell mass

10.3

6.8

6.8

6.8

5.0

54.1

43.8

Propionibacterium
acidipropionici ACKTet
53.4

Residues
(kg/h)
General
aqueous
residues

27,415

58,146

58,093

59,806

86,920

27,763

10,725

11,709

11,757

13,218

380

453

464

435

373

285

266

348

288

98.21%

70.86%

67.13%

88.26%

73.34%

579

Main products (kg/h)


Succinic
398
acid at
99 wt.%
Propionic
acid at
98 wt.%
Molar yield

54.4%

52.0%

71.4%

579

60,675

63.4%

87,083

58.8%

J.A. Posada et al. / Bioresource Technology 111 (2012) 282293

(1) Raw glycerol


diluted at 20 g/l

287

288

J.A. Posada et al. / Bioresource Technology 111 (2012) 282293

Fig. 3. Process ow diagrams for biochemical conversion of glycerol to: (a) D-lactic acid (case study IV), (b) succinic acid (case study V), (c) propionic acid (case study VI), (d)
PHB (case study VII), and (e) 1,3-propanediol (case study IX). H: Heater, HE: Heat Exchanger, R: Reactor, F: Fermentor, C: Centrifuge, Con: Condenser, DC: Distillation Column,
D: Divisor, Dec: Decanter, M: Mixer, Comp: Compressor, IER: Ionic Exchange Resin, RDC: Reactive Distillation Column, RE: Reactor Extractor, Sep: Separator.

The nal owsheet is presented in Fig. 3c. The downstream process


used a reactive-extraction unit with TOA as extractant of undissociated carboxylic acids (i.e., propionic acid, succinic acid, and acetic
acid). A clear relationship between the nal production of propionic acid was found to be related to fermentation yield. Thus, the
order for propionic acid production was: Scenario 1 > Scenario
4 > Scenario 5 > Scenario 2 > Scenario 3 (Table 3).
The 15 scenarios related to the carboxylic acids production
were economically assessed and the total production costs were
divided into (i) glycerol purication plus glycerol fermentation
and (ii) acid recovery and purication. The rst section (glycerol
purication plus fermentation) in all cases was between 9% and
26% of total productions costs, while the second section (acid
recovery and purication) was between 74% and 91% (Fig. 2d).
Also, it was noticed that utility costs increased with increasing global fermentation yield and decreasing glycerol concentration in the
fermentation media. This behavior can be explained by inefcient
usage of glycerol during the fermentation process and higher
requirement of services. A big difference in total production costs

was found for the production of D-lactic acid. While the lowest production cost was 1.015 US$/kg (Scenario 4), the highest one was
1.2 US$/kg (Scenario 2) and the increasing order for the total production costs was: Scenario 4 < Scenario 5 < Scenario 3 < Scenario
1 < Scenario 2. In addition, the ratio of Commercial Sale Price/Production Cost was higher than unity for all scenarios (Fig. 2c), and
was 1.53 for the Scenario 4. These results indicate that D-lactic acid
production from raw glycerol by mean of engineered E. coli could
be a protable alternative for glycerol usage. The lowest total production cost of D-lactic acid from raw glycerol was obtained for the
Scenario 4 in which three fermentative advantages were present:
(i) use of crude glycerol (85 wt.%), (ii) high glycerol concentration
in the fermentation media (40 g/l), and (iii) total consumption of
glycerol (Fig. 2c, D-lactic acid). Succinic acid production costs varied from 2.008 to 2.949 US$/kg. Although these differences are
mainly due to purication costs, they can be explained by complex
relationships between fermentation variables (e.g., strain, glycerol
concentration, glycerol purity, glycerol consumption, etc.) and succinic acid yields affecting the global performance of the process.

Table 4
Main processes streams and molar yields for cases of study VII and IX. Production of PHB, ethanol and 1,3-propanediol from glycerol with different qualities.
VII: PHB

Scenario

Raw glycerol
Down
stream
process I

Raw glycerol
Down
stream
process II

Raw glycerol
Down stream
process III

1000

1000

1000

Main reactives (kg/


h)
Raw glycerol at
88 wt.%
Pure Glycerol at
98 wt.%
Glycerol directly
diluted at
10 wt.%
Water
Strain
Cell mass

C. necator
JMP 134
45.6

VIII: Ethanol

C. necator
JMP 135
45.6

C. necator
JMP 136
45.6

Pure
glycerol
Down
stream
process I

Pure glycerol
Down
stream
process II

Pure glycerol
Down stream
process III

1000

1000

1000

C. necator
JMP 137
37.9

C. necator
JMP 138
37.9

C. necator
JMP 139
37.9

Residues (kg/h)
General aqueous
residues
Main products (kg/h)
25.0
Poly hydroxy
butirate at
99.9 wt.%
Ethanol at
99.5 wt.%
1,3-Propanediol at
99 wt.%
Molar yield

98%

25.4

98%

25.1

98%

48.3

98%

48.7

98%

IX: 1,3-Propanediol

Raw
glycerol
Diluted at
10 g/l

Raw
glycerol
Diluted at
20 g/l

576

578

Raw
glycerol
Diluted at
10 g/l

(1) Raw glycerol


diluted at 68 g/l

(2) Raw glycerol


diluted at 94 g/l

(3) Raw glycerol


diluted at 73 g/l

579

579

579

575

57.057

28.318

56.955

8536

6187

7952

Escherichia
coli
23.0

Escherichia
coli
17.3

Escherichia
coli
23.0

Klebsiella
pneumoniae
27.2

Klebsiella
pneumoniae
20.3

Klebsiella
pneumoniae
23.7

57.348

28.584

57.234

9270

6848

8688

273.07

290.51

293.38
252.70

238.40

248.40

52.38%

40.69%

48.34%

48.5

98%

92%

92%

92%

J.A. Posada et al. / Bioresource Technology 111 (2012) 282293

Case study

289

290

J.A. Posada et al. / Bioresource Technology 111 (2012) 282293

Fig. 4. Commercial Sale Price/Production Cost ratio for study cases VIIIX. (a) PHB production downstream process IIII for raw (88 wt.%) and pure (98 wt.%) glycerol, (b)
ethanol production (from: sugar cane, corn and crude glycerol), (c) ethanol production from: crude glycerol (at 10 and 20 g/l) and pure glycerol (at 10 g/l), and (d) 1,3propanediol production (Scenarios 13).

Fig. 5. Processing stages required for ethanol production from sugar cane, corn, and crude glycerol.

J.A. Posada et al. / Bioresource Technology 111 (2012) 282293

The order for succinic acid recovery and purication costs, and total production costs of succinic acid from raw glycerol were identical: Scenario 1 < Scenario 3 < Scenario 4 < Scenario 2 < Scenario 5.
The ratio of Commercial Sale Price/Production Cost was only higher than unity for the Scenario 1, but this value was still too close to
the unity. Thus, succinic acid production from glycerol still requires efforts to make this process more protable (Fig. 2c). The total production costs of propionic acid from raw glycerol ranged
from 1.683 to 1.942 US$/kg of propionic acid. The order for total
production costs was: Scenario 2 < Scenario 3 < Scenario 5 < Scenario 1 < Scenario 4. The ratio between Commercial Sale Price/Production Cost was calculated for the ve scenarios, and this ratio
was higher than the unity for: Scenarios 2, 3, and 5. Thus, this process could be protable only when high concentrations of glycerol
were used in the fermentation media and high yields of propionic
acid were obtained (Fig. 2c). Overall, the biotechnological route depends highly on fermentation performance, which depends on the
use of crude feedstock, high concentrations of feedstock in the fermentation media, and total consumption of feedstock. Consequently, one of the main goals of metabolic engineering should
be to develop specic strains able to completely consume the glycerol in raw substrates.
5.3.2. Case study VII: PHB
Polyhydroxy butyrate production from glycerol by C. necator
JMP 134 was simulated considering two fermentation stages, and
three extraction methods were analyzed for downstream processes
(named DSP-I, DSP-II and DSP-III, Fig. 3d). The rst downstream
process used a thermal treatment at 85 C, and then an enzymatic
digestion with pancreatin from Burkholdeira sp. PTU9 to cause cell
disruption and PHB is release. The second method used a highpressure homogenizer followed by a solvent extraction process
which modies the cell membrane permeability and causes the
PHB dissolution. The third process used an alkaline pretreatment
with a NaOH solution; then a digestion process using NaOCl and
sodium dodecyl sulfate (SDS). In all downstream processes, PHB
at 99.9 wt.% was obtained by spray drying. Additionally, two substrate qualities were used for these simulations, i.e., crude glycerol
(88 wt.%) and pure glycerol (98 wt.%) (Table 4). In all cases, the cost
of raw material was represented by the glycerol purication cost
described in Section 5.1. The glycerol purication process represents only between 4.8% and 5.6% of the total production cost of
PHB when glycerol at 88 wt.% was used, but this value increased
to between 6.3% and 7.7% when glycerol at 98 wt.% was used. In
general, lower production costs were obtained when glycerol at
98 wt.% was used in the fermentation process, due to the higher
PHB yield in the fermentation stage and the lower energy requirements in the spray drying process. The total PHB production costs
obtained were between 2.11 and 2.44 US$/kg when glycerol at
88 wt.% was used, and between 1.94 and 2.38 US$/kg for glycerol
at 98 wt.%. The comparison between the PHB Production Costs
and the Commercial Sale Price of PHB (i.e., Commercial Sale Price/
Production Cost ratio) is presented in Fig. 4a. In both cases (i.e.,
88 and 98 wt.%), a lower benet was obtained for Downstream Process II, which used a solvent extraction stage. This extraction required heating the expensive chemical solvent to 110 C, which
increased utility costs. The highest benet was achieved when
the three following conditions were met: (i) purication of raw
glycerol to a 98 wt.% level, (ii) two continuous fermentation stages,
and (iii) recovery of PHB with Downstream Process I.
5.3.3. Case study VIII: ethanol
Three different scenarios for ethanol production were considered and utilization of raw glycerol and traditional feedstocks
(i.e., sugar cane and corn) was compared from an economic point
of view. The main production stages for these three raw materials

291

Fig. 6. Global economic sale prices/total production cost ratio.

are compared in Fig. 5. All possibilities used a standard conguration consisting of raw material conditioning, fermentation, separation and dehydration, and waste treatment. Three fermentative
scenarios were analyzed: crude glycerol diluted to 10 g/L, crude
glycerol at 20 g/L, and pure glycerol at 10 g/L. The main simulation
results are summarized in Table 4. Similarly to the PHB production
process, the cost of raw material was represented by the glycerol
purication cost (Section 5.1). In all cases, utilities and capital costs
represented the highest cost of the purication process, i.e., around
20% and 30%, respectively. As a result, the lowest bioconversion
cost was obtained when crude glycerol (88 wt.%) diluted to 20 g/
L was used as feedstock, since it used a lower quantity of water
than the other two processes. Higher amounts of water were necessary when pure glycerol (98 wt.%) was used, increasing the
equipments size and the utility requirements. The global production costs for raw glycerol bioconversion to fuel ethanol were obtained by adding purication and bioconversion costs. In all
cases, the purication costs represented almost 35% of the total
production costs, while the bioconversion costs were near 65%.
These global production costs from raw glycerol were lower than
those reported by Quintero et al. (2008) for fuel ethanol production
from corn (0.3381 US$/L). Thus, using raw glycerol at 10 g/L and
20 g/L could represent a saving of 15% and 20%, respectively. The
global production costs were higher than those reported by Quintero et al. (2008) for fuel ethanol production from sugar cane
(0.2153 US$/L), but lower than the international prices for fuel ethanol ranging from (0.4552 US$/L (ICIS pricing, 2011) to 0.6057 US$/
L (Johnson and Taconi, 2007). The comparison of Commercial Sale
Price/Production Cost ratio for ethanol production from both sugar
cane and corn (Fig. 4b), and from glycerol (Fig. 4c), indicates that
the production process of fuel ethanol from raw glycerol by
E. coli could also be a protable alternative.
5.3.4. Case study IX: 1,3-propanediol production
Three production scenarios were considered based on previous
optimization work on glycerol fermentation in two continuous
stages. Three different operational congurations were available
depending on the desired process objective, namely global yield,
1,3-propanediol outlet concentration, or high simultaneous productivity (Posada, 2011). Under this framework, the rst scenario considered optimal conditions of volumetric productivity in the rst
fermentation stage and optimal nal concentration of 1,3-propanediol. The second scenario considered both conditions, the highest nal concentration of 1,3-propanediol and the highest productivity in
the second fermentation stage. Finally, the third scenario consid-

292

J.A. Posada et al. / Bioresource Technology 111 (2012) 282293

ered the optimal global productivity and taking into account both
fermentation stages. Additionally, an integrated reactor-extraction
scheme was proposed to recover and purify the 1,3-propanediol
from the fermentation broth to avoid the problems caused by the
high hydrophilicity of 1,3-propanediol and its high boiling point
(Fig. 3e). The nal production of 1,3-propanediol was mainly related
to the fermentation yield of both fermentation stages. Thus, the order for the nal concentration of 1,3-propanediol after the second
fermentation stage was: Scenario 2 (511.4 mmol/l) > Scenario 1
(419.7 mmol/l) > Scenario 3 (412.2 mmol/l). And the order for the
fermentative yield of glycerol to 1,3-propanediol was: Scenario
1 > Scenario 3 > Scenario 2; the order for the actual production of
1,3-propanediol was: Scenario 1 > Scenario 3 > Scenario 2 (Table 4).
For the economic assessment of 1,3-propanediol production
from raw glycerol, the total production cost can be divided into
glycerol purication plus glycerol fermentation and 1,3-propanediol recovery and purication. The rst set of costs accounted for
29.432.4% of the total costs and the second set for 67.670.6%.
The 1,3-propanediol purication cost was estimated to be between
0.652 and 0.758 US$/kg. Among the three scenarios, the Commercial Sale Price/Production Cost ratio was higher than unity for all
scenarios, and for the Scenario 3 this ratio was 1.832. These results
indicate that 1,3-propanediol production from raw glycerol by K.
pneumoniae could be a protable alternative for glycerol usage
(Fig. 4d).
Finally, the results for the economic assessment (Commercial
Sale Price/Total Production ratio), as primary feasibility indicator
for the production of added-value products derived from glycerol,
are summarized in Fig. 6. The glycerol purication process (glycerol USP) was also included as an additional high valued derivative.
Initially, all products may generate income because all of them
have a ratio higher than one. More income can be earned from
1,3 propanediol and 1,2 propanediol production, while less income
would be obtained from hydrogen and succinic acid synthesis because of their low production rates and high purication costs. An
interesting fact revealed by Fig. 6 is that purifying glycerol to USP
grade can be more protable than its transformation into products
such as ethanol or acrolein. However, this situation may change in
the near future with higher technological maturity represented by
higher conversion levels, improved selectivities and better downstream technologies.
6. Conclusions
Results obtained for glycerol purication and nine added-value
products generated from glycerol showed that not only quality
requirements were successfully reached, but that all the processes
were protable. In this rst process design approach, glycerol has
been demonstrated to be a renewable and non-expensive feedstock for a system of bioreneries. Moreover, given the fact that
the quality of the glycerol analyzed in this work does not depend
directly on the country where it is produced these results can be
potentially applied to other glycerol conversion facilities in the
world.
Acknowledgement
To the Universidad Nacional de Colombia and Colombian Institute for Development of Science and Technology (COLCIENCIAS),
for the nancial support of this work.
References
Ahmed, I.I., Nipattummakul, N., Gupta, A.K., 2011. Characteristics of syngas from cogasication of polyethylene and woodchips. Applied Energy 88 (1), 165174.

Akiyama, M., Sato, S., Takahashi, R., Inui, K., Yokota, M., 2009. Dehydration
hydrogenation of glycerol into 1,2-propanediol at ambient hydrogen pressure.
Applied Catalysis A: General 371 (12), 6066.
Blankschien, M.D., Clomburg, J.M., Gonzalez, R., 2010. Metabolic engineering of
Escherichia coli for the production of succinate from glycerol. Metabolic
Engineering 12 (5), 409419.
Cardona, C.A., Snchez, .J., 2007. Fuel ethanol production: process design trends
and integration opportunities. Bioresource Technology 98 (12), 24152457.
Casale, B., Gomez, A.M., 1994. Catalytic method of hydrogenating glycerol. U.S.
Patent No. 5276,181.
Cavalheiro, J.M.B.T., de Almeida, M.C.M.D., Grandls, C., da Fonseca, M.M.R., 2009.
Poly(3-hydroxybutyrate) production by Cupriavidus necator using waste
glycerol. Process Biochemistry 44 (5), 509515.
Clomburg, J.M., Gonzalez, R., 2011. Metabolic engineering of Escherichia coli for the
production of 1,2-propanediol from glycerol. Biotechnology and Bioengineering
108 (4), 867879.
Coral, J., Karp, S., Porto de Souza Vandenberghe, L., Parada, J., Pandey, A., Soccol, C.,
2008. Batch fermentation model of propionic acid production by
propionibacterium acidipropionici in different carbon sources. Applied
Biochemistry and Biotechnology 151 (2), 333341.
Dasari, M.A., Kiatsimkul, P.-P., Sutterlin, W.R., Suppes, G.J., 2005a. Low-pressure
hydrogenolysis of glycerol to propylene glycol. Applied Catalysis A: General 281
(12), 225231.
Dasari, M.A., Kiatsimkul, P.-P., Sutterlin, W.R., Suppes, G.J., 2005b. Low-pressure
hydrogenolysis of glycerol to propylene glycol. Applied Catalysis A: General
281, 225231.
Demirel-Glen, S., Lucas, M., Claus, P., 2005. Liquid phase oxidation of glycerol over
carbon supported gold catalysts. Catalysis Today 102103, 166172.
Durnin, G., Clomburg, J., Yeates, Z., Alvarez, P.J.J., Zygourakis, K., Campbell, P., Gonzalez,
R., 2009. Understanding and harnessing the microaerobic metabolism of glycerol
in Escherichia coli. Biotechnology and Bioengineering 103 (1), 148161.
Garcia, R., Besson, M., Gallezot, P., 1995. Chemoselective catalytic oxidation of
glycerol with air on platinum metals. Applied Catalysis A: General 127 (12),
165176.
Hjek, M., Skopal, F., 2010. Treatment of glycerol phase formed by biodiesel
production. Bioresource Technology 101 (9), 32423245.
ICIS Pricing, 2011. Ethanol Prices and Pricing Information, vol. 2011, ICIS.com.
Johnson, D.T., Taconi, K.A., 2007. The glycerin glut: options for the value-added
conversion of crude glycerol resulting from biodiesel production. Environ
Progress 26 (4), 338348.
Kim, B., Hong, Y., Hong, W., 2004. Effect of salts on the extraction characteristics of
succinic acid by predispersed solvent extraction. Biotechnology and Bioprocess
Engineering 9 (3), 207211.
Lahr, D.G., Shanks, B.H., 2005. Effect of sulfur and temperature on rutheniumcatalyzed glycerol hydrogenolysis to glycols. Journal of Catalysis 232, 386
394.
Matsumura, Y., Minowa, T., Potic, B., Kersten, S., Prins, W., vanSwaaij, W.,
vanderBeld, B., Elliott, D.C., Neuenschwander, G.G., Kruse, A., Jr, M.J.A., 2005.
Biomass gasication in near- and super-critical water: status and prospects
(review). Biomass and Bioenergy 29, 269292.
Mazumdar, S., Clomburg, J.M., Gonzalez, R., 2010. Escherichia coli strains engineered
for homofermentative production of D-lactic acid from glycerol. Applied and
Environmental Microbiology 76, 43274336.
Mozaffarian, M., Deurwaarder, E.P., Kersten, S.R.A., 2004. Green Gas (SNG)
Production by Supercritical Gasication of Biomass. Twente University and
ECN Biomass.
Murarka, A., Dharmadi, Y., Yazdani, S.S., Gonzalez, R., 2008. Fermentative utilization
of glycerol by Escherichia coli and its implications for the production of fuels and
chemicals. Applied and Environmental Microbiology 74 (4), 11241135.
Pablo, I.N., Pettinari, M.J., Galvagno, M.A., Mndez, B.S., 2008. Poly(3hydroxybutyrate) synthesis from glycerol by a recombinant Escherichia coli
arcA mutant in fed-batch microaerobic cultures. Applied Microbiology and
Biotechnology 77, 13371343.
Posada, J.A., 2011. Design and analysis of technological schemes for glycerol
conversion to added value products. In: Department of Electric and Electronic
Engineering, PhD., Universidad Nacional de Colombia Sede Manizales,
Manizales, Colombia.
Posada, J.A., Cardona, C.A., 2010a. Anlisis de la renacin de glicerina obtenida
como co-producto en la produccin de biodiesel (Validation of glycerin rening
obtained as a by-product of biodiesel production). Ingeniera y Universidad 14,
227.
Posada, J.A., Cardona, C.A., 2010b. Design and analysis of fuel ethanol production
from raw glycerol. Energy 35 (12), 52865293.
Posada, J.A., Cardona, C.A., Rincn, L.E., 2010. Sustainable biodiesel production from
palm using in situ produced glycerol and biomass for raw bioethanol. In: 32nd
Symposium on Biotechnology for Fuels and Chemicals, April 1922, Clearwater
Beach, Florida.
Posada, J.A., Cardona, C.A., Gonzalez, R., 2011a. Analysis of the production process of
optically pure D-lactic acid from raw glycerol using engineered Escherichia coli
strains. Applied Biochemistry and Biotechnology, 120.
Posada, J.A., Naranjo, J.M., Lpez, J.A., Higuita, J.C., Cardona, C.A., 2011b. Design and
analysis of poly-3-hydroxybutyrate production processes from crude glycerol.
Process Biochemistry 46 (1), 310317.
Quintero, J.A., Montoya, M.I., Snchez, O.J., Giraldo, O.H., Cardona, C.A., 2008. Fuel
ethanol production from sugarcane and corn: comparative analysis for a
Colombian case. Energy 33 (3), 385399.

J.A. Posada et al. / Bioresource Technology 111 (2012) 282293


Shams Yazdani, S., Gonzalez, R., 2008. Engineering Escherichia coli for the efcient
conversion of glycerol to ethanol and co-products. Metabolic Engineering 10
(6), 340351.
Thompson, J.C., He, B.B., 2006. Characterization of crude glycerol from biodiesel
production from multiple feedstocks. Applied Engineering in Agriculture 22 (2),
261265.
Tsukuda, E., Sato, S., Takahashi, R., Sodesawa, T., 2007. Production of acrolein from
glycerol over silica-supported heteropoly acids. Catalysis Communications 8
(9), 13491353.

293

Valliyappan, T., Bakhshi, N.N., Dalai, A.K., 2008. Pyrolysis of glycerol for the
production of hydrogen or syn-gas. Bioresource Technology 99 (10), 4476
4483.
Ya-Nan, Z., Guo, C., Shan-Jing, Y., 2006. Microbial production of 1,3-propanediol
from glycerol by encapsulated Klebsiella pneumoniae. Biochemical Engineering
Journal 32, 9399.
Zhang, A., Yang, S.-T., 2009. Propionic acid production from glycerol by
metabolically
engineered
Propionibacterium
acidipropionici.
Process
Biochemistry 44 (12), 13461351.

Você também pode gostar