You are on page 1of 24

Accepted Manuscript

Title: Comparison of process technologies for chitosan


production from shrimp shell waste: A Techno-economic
approach using Aspen Plus®

Authors: David Gómez-Rı́os, Rolando Barrera-Zapata,


Rigoberto Rı́os-Estepa

PII: S0960-3085(17)30032-9
DOI: http://dx.doi.org/doi:10.1016/j.fbp.2017.02.010
Reference: FBP 845

To appear in: Food and Bioproducts Processing

Received date: 17-11-2016


Revised date: 22-2-2017
Accepted date: 24-2-2017

Please cite this article as: Gómez-Rı́os, David, Barrera-Zapata, Rolando, Rı́os-Estepa,
Rigoberto, Comparison of process technologies for chitosan production from shrimp
shell waste: A Techno-economic approach using Aspen Plus®.Food and Bioproducts
Processing http://dx.doi.org/10.1016/j.fbp.2017.02.010

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

Experimental characterization of shrimp waste and optimal conditions of reaction have been

established for chitosan synthesis.

Technical feasibility of chitosan two production processes has been studied through CAPE

simulation.

Experimental kinetics for reactions involved in chitosan synthesis are proposed for simulation

purposes.

Chitosan production using physical-chemical method of chitin isolation is more profitable than

if fermentative stage is included.

1
Comparison of process technologies for chitosan production from shrimp shell
waste: A Techno-economic approach using Aspen Plus®

David Gómez-Ríos a, Rolando Barrera-Zapata b, Rigoberto Ríos-Estepa c.


a
Facultad de Ingeniería Mecánica, Universidad Antonio Nariño UAN, Calle 52 No. 40 - 88,
Medellín, Colombia. b Grupo CERES, Departamento de Ingeniería Química, Facultad de
Ingeniería, Universidad de Antioquia UdeA, Calle 70 No. 52-21, Medellín, Colombia. c Grupo
de Bioprocesos, Departamento de Ingeniería Química, Facultad de Ingeniería, Universidad de
Antioquia UdeA, Calle 70 No. 52-21, Medellín, Colombia.

Keywords. Chitin from shrimp shell waste, Aspen Plus plant simulation, techno-economic
comparison, industrial waste valorization.

Abstract. Techno-economic analyses for two different processing methods for chitosan
production from shrimp shell waste were performed. Experimental lab-scale data that included
the shrimp shell waste characterization as well as plant process simulated data were used for the
techno-economic comparison of the processes. All the simulations were carried out in the Aspen
Plus® software. The economic assessment was made in terms of profitability, production costs,
net present value, internal rate of return and payback. Both studied processes were found to be
profitable and cost-competitive in the geographic regional context, with a mean gross margin of
70% and mean internal rate of return of 25.5%.

1. Introduction.

Chitosan is a polymeric material composed of β-(1-4) D-glucosamine units; it is obtained by


chemical or enzymatic deacetylation of the natural amino polysaccharide chitin, which is
commonly found as a structural compound of arthropods (e.g., lobster and shrimp) and fungi,
and is formed of β-(1-4) N-acetyl D-glucosamine units (Pillai et al., 2009). Chitosan has
chelating properties, biocompatibility and biodegradability, is non-toxic and shows good
adsorption properties when compared with different synthetic plastic materials (Peniche et al.,
2008). It has many beneficial applications in food industry, agriculture, water treatment,
manufacturing of medical devices and drug-delivery, manufacturing of artificial membranes and
tissues, and as a constituent in nutraceuticals, cosmetics, antifungal and antibacterial products,
among others (Instituto Colombiano Agropecuario, 2012; Peniche et al., 2008).

Production of chitosan is directly related to the fishing industry. Latin-American countries,


with coasts over the Pacific Ocean, consolidate around 41% of global exportations of shrimp;
Ecuador, Argentina and Mexico are the larger producers in the region. In Colombia and Ecuador,
the Penaeus Vannamei shrimp farming is nowadays one of the most important and prospective

2
activities of aquaculture business, because its large potential of expansion and the tendencies to
growth in local and international markets (Global Industry Analysts Inc., 2014; Instituto
Colombiano Agropecuario, 2012).

In food industry approximately one-third of food produced for human consumption is lost or
wasted globally, generating waste at rate of around 1.3 billion Ton/year (Kwan et al., 2015). In
the industrial shrimp processing, approximately 20% of the gross weight of shrimp is discarded
as waste (Shirai, 1999); such waste can be used as raw material for the production of chitosan,
representing an important source of revenues for shrimp industry. The global market for chitosan
is projected to exceed 118000 tons by 2018 and it has been forecasted a relative lack of suppliers
for meeting the demand (Global Industry Analysts Inc., 2014).

Interestingly, Latin America has the potential to attend the growing global demand of chitosan
(Goycoolea et al., 2004). Despite this, market studies do not show an important participation of
the region in the global chitosan production (Global Industry Analysts Inc., 2014). The
increasing interest in the use of shrimp waste for industrial production of chitosan has promoted
the research in related topics, as well as few industrial initiatives. However, the uncertainty
regarding the costs/profitability ratio hampers the investment in this emerging industry. At the
present time, only five known producers are active in the region (Gómez-Ríos, 2016).

Techno-Economic Analysis (TEA) is in principle a cost-profit comparison of different


engineering alternatives (Lauer, 2008). TEA is normally based on process specifications,
material and energy requirements, equipment, services, pricing, production costs and investment
(Sinnott and Towler, 2009). The development of a TEA can be improved by means of Computer-
Aided Process Engineering (CAPE) tools (Alshekhli et al., 2011), which are often used as
reliable tools for specific simulation tasks, i.e., development of energy balances and/or mass
balances, yield prediction, parameter estimation in mathematical models, process behavior
simulation at different production scenarios, sizing and costing, among others (Alshekhli et al.,
2011; Piccolo and Bezzo, 2009).

In this contribution, Techno-Economic Analyses are performed for two different chitosan
production methods from shrimp shell waste. Experimental lab-scale results and CAPE tools are
used as data source for the TEA. The software Aspen Plus® v.8.0 is used for the process plant
simulations. These TEAs are intended to give insights about the possibility of establishing a
chitosan industrial production plant in Colombia or Latin American countries with similar
geographic conditions for shrimp culture.

2. EXPERIMENTAL (MATERIALS AND METHODS)

2.1 Raw material characterization. The Penaeus Vannamei shrimp waste were characterized
by means of a composition analysis of different samples. Samples were acquired in different
harvesting seasons and from different geographic locations, i.e., Colombia and Ecuador.

3
Analysis included moisture content, mineral content (Ghorbel-Bellaaj et al., 2011; Hernández et
al., 2009) and pigment content (Herrera et al., 2011). Proteic content was quantified by
determination of total nitrogen and soluble protein in crude shrimp shells and the fatty fraction
was determined by Soxhlet extraction (Rødde et al., 2008; Shirai, 1999). The average
composition for the shrimp waste samples is presented in Table 1.

Table 1. Average composition of Penaeus vannamei shrimp shells

Component Composition (%)

Astaxanthin 0.41

Calcium carbonate 5.13

Calcium phosphate 12.90

Chitin 29.97

Moisture 4.45

Magnesium carbonate 1.52

Total fat 6.40

Total protein 36.60

Sodium carbonate 2.62

2.2 Experimental set up for chitosan production. In this contribution, cost estimates for
chitosan production from shrimp cell waste were acquired from plant simulations in Aspen
plus®. For this, lab-scale experiments were performed in order to identify the appropriate system
operating conditions and acquire accurate information for feeding the related plant models, e.g.,
reaction kinetics and specific processing stages.

2.2.1 Isolation of chitin by physical-chemical (PC) methods. Clean and dried shrimp waste were
grinded to a particle size between 0.5 and 10mm; then, depigmentation was carried out by means
of material suspension into an ethanol-water solution 80% w/w using a solid-liquid ratio of 1:5
(g/mL) at 200 RPM (Escobar et al., 2011). Once the material was grinded and discolored,
proteins were extracted by means of alkaline hydrolysis using aqueous sodium hydroxide 3.7%
w/w (Ahlafi et al., 2013; Percot et al., 2003) at 71°C under constant stirring (Bajaj et al., 2011;
Gómez-Ríos, 2016) followed by acid solubilization of inorganic salts using hydrochloric acid
5.2% w/w, at room temperature and constant stirring (Escobar et al., 2013). The resulting solid

4
fraction is mainly chitin, which can be either commercialized or used as raw material for the
synthesis of chitosan.

2.2.2 Isolation of chitin by a combination of fermentative and physical-chemical (FPC) methods.


Chitin can also be obtained from fermentation of shrimp shell waste using lactic acid bacteria
(L.A. Cira et al., 2002; Duan et al., 2012; Pacheco, 2010; Sini et al., 2007). In this case, a mix of
Lactobacillus bulgaricus and Streptococcus thermophiles, activated as indicated by Duan et al
(Duan et al., 2012) was inoculated to different fermentation media. The best results were
obtained for the medium prepared with glucose 4% w/w supplemented with salts, at pH 6.4,
using a solid-liquid ratio (g/mL) of 1:15 (Gómez-Ríos, 2016).

Lactic acid bacteria use sugars available in the medium for biomass and lactic acid biosynthesis.
The accumulation of lactic acid promotes solubility of inorganic salts present in shrimp waste,
thus turning into a demineralization stage. The mineral extraction proceeds completely in this
fermentative stage (Gómez-Ríos, 2016; Pacheco, 2010). Furthermore, proteolytic enzymes allow
bacteria to use peptides and proteins as nitrogen and carbon sources, thereby performing
deproteinization of raw material (Luis A Cira et al., 2002; Pacheco, 2010). Nonetheless, protein
removal achieved through fermentation varies from 60 to 68% (Gómez-Ríos, 2016; Pacheco et
al., 2009); likewise, pigment content is reduced around 50%. Hence, extraction of pigments and
proteins in the FPC methods should be completed by PC methods as described in section 2.2.1.

2.2.3 Chemical deacetylation (CDA) of chitin. High quality chitosan contains approximately
90% of glucosamine units, i.e., a deacetylation degree (%DD) of 90% (Weska et al., 2007). From
lab-scale experiments it was observed that the deacetylation degree could be further improved
when temperature and pressure are adjusted while the chemical reaction advances in certain
extent. Accordingly, chitin was suspended into an aqueous solution of sodium hydroxide 50%
w/w, in a solid-liquid ratio (g/mL) of 1:24, with continuous stirring at 90°C and atmospheric
pressure. When 75% of reaction was completed, the reactor temperature and pressure were
increased to 120°C and 1.2barg, respectively. At the end of the reaction, the solid fraction was
characterized by FTIR (Fourier Transform Infrared Spectroscopy), potentiometric titration with
sodium hydroxide (Czechowska-biskup et al., 2012) and physical chemical tests.
Characterization of chitosan, obtained in the present study, showed a mean %DD of 89%, a mean
viscosity of 47.7 mPa-s and 50 mesh particle size (Gómez-Ríos, 2016).

2.3 Reaction kinetics. Kinetic power law expressions used in the Aspen Plus® simulations were
adjusted through regression of experimental data acquired in this study, and experimental data
taken from literature (Ahlafi et al., 2013; Ameh et al., 2013; Moura et al., 2011; Pacheco et al.,
2009). Non-elementary chemical reaction and dependence on concentration of a controlling
species was assumed. Table 2 presents the kinetic expressions for reactions involved in the PC
and FPC methods for chitosan isolation, as well as the kinetic expression of the reaction for
chitin deacetylation.

5
Table 2. Reaction kinetics for the stages involved in chitosan synthesis

Reaction Reaction rate k (consistent E (kJ/kmol)


3
expression (kmol/m h) units)

Deacetylation of chitin (CDA) r = kT2.03 e –E/RT CNaOH 3.77x10-5 48760

Demineralization (PC method) r = kT CHCl 5.3x10-8

Deproteinization (PC method) r = kT2 Cprot 2.92x10-11

Demineralization (FPC method) r = ke –E/RT Cmin 6.8x103 40092

Deproteinization (FPC method) r = ke –E/RT Cprot 1.5x105 50499

Lactic fermentation (FPC method) r = ke –E/RT Csugar 3x106 44912

Molar concentrations expressed in kmol/m3; Universal constant R in kJ/kmol K, and temperature


T in K. Cprot = total protein concentration; Cmin = total mineral concentration

3. PROCESSES SIMULATION

In this work, two methods were evaluated as feasible alternatives for chitosan production from
Penaeus Vannamei shrimp industrial waste. The first method considered PC isolation of chitin
followed by CDA for chitosan production (PC-CDA). The second process considered FPC for
chitin isolation and CDA for chitosan production (FPC-CDA). Both methods were simulated in
Aspen Plus® v8.0. Considering that chitosan production is essentially a combination of
independent batch sub-processes, the simulation strategy included individual units of processing
rather than a continuous plant, i.e., the feeding stream composition for every batch-stage was
defined according to the output stream and results of simulation for the previous batch-stage.

Subject to raw material availability in the region, it was simulated the batch processing of 1
ton/cycle of fresh shrimp waste, which is equivalent to approximately 272 kg of dry waste, and
yields about 65.8 kg/cycle of chitosan; this production corresponds to a pilot plant production
level in the fine chemistry industry (Kwan et al., 2015). Chitin content in the raw material was
supposed to fit a constant value of 24.2% (Gómez-Ríos, 2016; Shirai, 1999). No pretreatment
was included in the process simulation; thus, the grinding stage was set as the first step in the
processing of shrimp waste. Composition of the feeding stream (solid raw material) was stated
according to the experimental characterization of Penaeus vannamei shrimp shell.

For simulation purposes, protein material in shrimp industrial waste was supposed to be
composed of methionine, phenylalanine, alanine, glutamic acid and lysine peptides. The small

6
fraction of fats present in the raw material was assumed to be methyl-palmitate. Pigment content
in raw material was assumed as pure astaxanthin. Chitin was supposed to be formed only by its
conformational unit D-N-Acetylglucosamine. Solution chemistry was included by using the
“Elect-wizard” assistant from Aspen Plus®; therefore, a detailed electrolyte set of reactions
comprising dissociation, equilibrium and salt formation were included in the simulation of
reactive stages (Aspen Technology Inc., 2015). The Elect-NRTL thermodynamic model was
used in simulation of all processing units. The set of reactions used for chitosan production
simulation and their stoichiometry are presented in Table 3.

Table 3. Reaction sets used in simulation of the stages involved in chitosan synthesis

Reaction set Reaction stoichiometry

Deacetylation (CDA) (1) C8H15NO6 + NaOH C6H13NO5 + C2H3NaO2

Demineralization (PC) (2) CaCO3 + 2 HCl  CaCl2 + H2O + CO2

(3) Na2CO3 + 2 HCl  2 NaCl + H2O + CO2

(4) MgCO3 + 2 HCl  MgCl2 + H2O + CO2

(5) Ca3(PO4)2+ 6 HCl  3 CaCl2 + 2 H3PO4

Deproteinization (PC) (6) C6H12N2O3 + 2 NaOH  2 C3H6NNaO2 + H2O

(7) C10H16N2O7 + 2 NaOH  2 C5H8NNaO4 + H2O

(8) C18H20N2O3 + 2 NaOH  2C9H10NNaO2 + H2O

(9) C10H20N2O3S2 + 2 NaOH  2 C9H10NNaO2 + H2O

(10) C12H26N4O3 + 2 NaOH  2 C6H13N2NaO2 + H2O

Demineralization (FPC) (11) CaCO3 + 2 C3H6O3  C6H10CaO6+ H2O + CO2

(12) Na2CO3 + 2 C3H6O3  2 C3H5NaO3 + H2O + CO2

(13) MgCO3 + 2 C3H6O3  C6H10MgO6 + H2O + CO2

(14) Ca3(PO4)2+ 6 C3H6O3  3 C6H10CaO6 + 2 H3PO

Deproteinization (FPC) (15) C6H12N2O3 + H2O  2 C3H7NO2

(16) C10H16N2O7 + H2O  2 C5H9NO4

7
(17) C18H20N2O3 + H2O  2 C9H11NO2

(18) C10H20N2O3S2 + H2O  2 C9H11NO2

(19) C12H26N4O3 + H2O  2 C6H14N2O2

Lactic fermentation (FPC) (20) C12H22O11  2 C3H6O3

It is known that chitosan production is intensive in water and reactants consumption since it is
required a large liquid-solid ratio in the contact and reaction operations (Gómez-Ríos, 2016;
Khan et al., 2001a; Weska et al., 2007). Reduction of production cost depends markedly on
effective recycling of effluents; hence, the recycling of the deacetylation and pigment extraction
solutions was proposed as alternative for a more cost-efficient process.

Regardless of the process (PC-CDA) or (FPC-CDA), the recycle strategy is same and the mass
rate varies according to requirements of the process. The pigment extraction solvent must be
recovered by single evaporation of the solution at the azeotropic temperature, the vapor is
condensed and purged (5% mass), the resulting concentrate can be discarded as waste or used for
pigment obtention. The sodium hydroxide solution of deacetylation is recycled and purged (10%
mass) without further treatment, since less of 10% of sodium hydroxide reacts effectively in
deacetylation and no significant interference of acetate ion was observed on the reaction kinetics.
The concentration of the recycled solutions is adjusted by adding fresh reactants (ethanol and
sodium hydroxide respectively) according to the purge. If such recycles are avoided the cost of
materials is increased approximately in 70% with an important impact on the production cost.

3.1 Simulation of chitosan production by the PC-CDA process. CDA of chitin, obtained by
means of PC treatment of shrimp waste, is one of the most used processes for chitosan
production. The process, summarized in Figure 1, involves reactions (2) to (10) for PC chitin
isolation, and reaction (1) for CDA of chitin (see Table 3).

Figure 1. Process diagram for chitosan production using the PC-CDA method

8
Simulation of the size reduction stage was performed assuming a Bond work index of 12.5
kWh/ton. The size reduction unit was simulated as a gyratory crusher with a cut off size ratio of
6. Results for the outlet stream show that more than 80% of the solids have an average particle
diameter of 7.7 mm. The power requirement, estimated by simulation, was 243.3 W for size
reduction of the total raw material stream, i.e., 260 kg of dry solid material.

The pigment extraction stage was simulated as an isothermal batch reactor, fed with an
ethanol-water solution and a solid stream corresponding to the grinder outlet stream. Even
though the extraction does not involve a reaction step, for simulation proposes, it can be assumed
that solubilization of pigments occurs as a chemical reaction that follows a kinetic behavior
dependent on pigment concentration. Simulation results show that the accumulated mass of
astaxanthin in the solution increases with time, giving rise to the higher extraction velocity
during the first 20 minutes.

The ethanol-water solution employed for astaxanthin extraction must be recovered and
recycled after filtration to improve the process economy. The recycle process considered the
extract heating at 87°C in a countercurrent heat exchanger with a calculated heat duty of 438.8
kW and 6.48 m2 of heat transfer area. After heating, 99.6% of the extraction solution is recovered
in a single evaporator, operating at 1 bar and 86°C, followed by condensation in a water cooled
countercurrent heat exchanger with a calculated cooling duty of -380.9kW and 8.12 m2 of heat
transfer area, followed by purge (5%) and concentration with ethanol.

The material is fed to the deproteinization reactor, which was simulated as a batch reactor at
atmospheric pressure. The deproteinization reaction was simulated under a constant heating duty
of 108 kW, with smooth increments of temperature until reaching 70°C. Further
demineralization was simulated in a batch reactor at atmospheric pressure with a constant
cooling duty of -140 kW; this way, the reactor temperature was ranging from 25 to 35 °C.
Simulated results of the demineralization stage were used to fix the feeding stream conditions to
the CDA batch reactor. Considering that the CDA reaction is slightly exothermic, a heating
profile was defined to hold the temperature at specific values according to the reaction stage (see
section 2.2.3). Pressure was also adjusted to fit required values in each reaction stage. Finally,
the solid product is separated through filtration and sent to a drying stage. The filtrate consisting
of aqueous sodium hydroxide–sodium acetate solution is recycled, purged (10%) and
concentrated by adding fresh sodium hydroxide solution.

For simulation of the drying stage, a solid humidity of 10%, after filtration, was assumed. It was
considered a hot air stream with 10% absolute humidity, at 1.4 bar and 20°C of superheating for
drying the load of solids in 20 minutes. As it is showed in Figure 2, there are several filtration
stages along the process. Simulations of the solid-liquid separations in Aspen Plus® were
performed as vacuum rotary drum filters, assuming specific operation conditions, i.e., a
maximum pressure drop of 40 kPa, incompressible cake, a filter medium resistance of 4.05 x

9
109m-1 and a specific cake resistance of 2.0x106 (Khan et al., 2001b). The processing time per
unit, required in the PC-CDA process, is detailed in Table 4. Total processing time in the PC-
CDA process is 495 min.

Table 4. Processing time for chitosan production using the physicochemical method

Equipment Unit Operation Processing Starting time


(Figure 1) time (min)

CR-100 Size reduction 30 Start of batch


processing

T-100 Pigment extraction 20 After processing in CR-


100

F-100 Filtration of depigmented material 30 After processing in T-


100

R-100 Protein hydrolysis reaction 124 After filtration in F-100

F-200 Filtration of deproteinized 20 After processing in R-


material 100

R-200 Mineral solubilization reaction 47 After filtration in F-200

F-300 Filtration of chitin 12 After processing in R-


200

R-300 Deacetylation reaction 1st stage: 120 After filtration in F-300

2nd stage: 60

F-400 Filtration of chitosan 12 After processing in R-


300

D-100 Drying of product 20 After filtration in F-400

HX-100, E- Pigment extraction solvent 60 After filtration in F-100


100, HX- recovery
200

T-200 Pigment extraction solvent 6 After condensation in


adjustment and storage HX-200

10
T-300 Deacetylation solution adjustment 9.5 After filtration in F-400
and storage

Total cycle time (min) 495

Minimum cycle time (min) 190

Total number of batches (min) 346

Annual operating time (h) 4152

3.2 Simulation of chitosan production using the FPC-CDA method. The simulation of the
FPC- CDA process considered the reactions (11) to (20) (See Table 3). Total extraction of
pigments and proteins is achieved through further chemical depigmentation and deproteinization,
involving reactions (6) to (10), prior to the CDA reaction (1). The process diagram is presented
in Figure 2.

Figure 2. Process diagram for chitosan production using the FPC-CDA method

The fermentative process was simulated to occur in a batch reactor. Loading to the reactor was
defined as formed by an aqueous fermentation medium with a sugar concentration of 4% w/w,
and grinded shrimp waste 6.25% w/w. From lab-scale experiments it was observed that mineral
and protein material from shrimp waste decreases noticeably within the first 36 hours of
fermentation; therefore, the time of bio-pretreatment was set to 41 hours. The process was
simulated as isothermal at 32.5°C. According to simulations, around 70% of protein content is
extracted through fermentation, while minerals are completely dissolved as a consequence of pH
reduction caused by the accumulation of lactic acid.

The additional physical-chemical treatment for enhancement of %DD in the FPC-CDA method
was simulated following the same reaction kinetics as for the PC-CDA method (see Table 2). In
the FPC-CDA method the pigment extraction takes place after the fermentation cycle; the
amount of solvent required is reduced in a 50% with respect to the PC-CDA method, and the
pigment extraction time is reasonably reduced to 17 minutes. Furthermore, processing time for

11
completion of protein extraction decreased to 40 minutes and quantity of sodium hydroxide
solution required is reduced in a 50% with respect to the PC-CDA method. The summary of
processing times per unit is presented in Table 5. Total processing time in the FPC-CDA process
is 2861 min

Table 5. Processing time for chitosan production by using the fermentative and physicochemical
methods

Equipment Unit Operation Process time Start time


(Figure 2) (min)

CR-101 Size reduction 30 Start of batch processing

R-101 Fermentation of solids 2460 (41 hours) After processing in CR-1

F-101 Filtration of fermented solids 60 After fermentation in R-


101

T-101 Pigment extraction 17 After filtration in F-101

F-201 Filtration of depigmented 30 After processing in T-101


material

R-201 Protein hydrolysis reaction 40 After filtration in F-201

F-301 Filtration of chitin 12 After processing in R-


201

R-301 Deacetylation reaction 1st stage: 120 After filtration in F-301

2nd stage: 60

F-401 Filtration of chitosan 12 After processing in R-


301

D-101 Drying of product 20 After filtration in F-401

HX-101, E- Pigment extraction solvent 30 After filtration in F-201


101, HX- recovery
201

T-201 Pigment extraction solvent 4 After condensation in


adjustment and storage HX-201

12
T-301 Deacetylation solution 9.5 After filtration in F-401
adjustment and storage

Total cycle time (min) 2861

Minimum cycle time (min) 2460

Total number of batches 116

Annual operating time (h) 8300

A comparison between the mass balance per component and the energy requirements for the PC-
CDA, and the FPC-CDA method, is presented in Table 6 and Table 7, respectively.

Table 6. Mass balance per component for the simulated processes of chitosan production

CDA-PC CDA-FPC

Input Output Input Output


Component
(kg/cycle) (kg/cycle) (kg/cycle) (kg/cycle)

Chitosan Effluents Chitosan Effluents

Proteins/Peptides 95.16 4.924x10-4 -- 95.16 1.178x10-3 --

Amino-acid solution -- -- 101.894 -- -- 101.857

Fats 16.64 1.18x10-6 16.639 16.64 3.39x10-11 16.640

Astaxanthin 1.066 7.58x10-8 1.066 1.066 2.17x10-12 1.066

Sodium acetate -- -- 23.490 -- -- 23.490

Water 8665.1 -- 8665 6129 -- 6129

Ethanol 47.84 -- 47.84 23.92 -- 23.92

Sodium hydroxide 277.72 -- 199.77 186.1 -- 137.7

Hydrochloric acid 143.3 -- 63.125 -- -- --

Sugar -- -- -- 134.78 -- 2.28

13
Calcium chloride -- 0.017 79.795 -- -- --

Sodium chloride -- 0.044 21.010 -- -- --

Phosphoric acid -- -- 35.228 -- -- 35.228

Carbon dioxide -- -- 24.476 -- -- 24.476

Magnesium chloride -- 0.039 18.724 -- -- --

Chitin 77.91 -- 14.188 77.91 -- 14.188

Deacetylated chitin -- -- 51.613 -- -- 51.613

Sodium hydroxide (50% -- -- 1781.2 -- -- 1781.2


w/w-recycle)

Ethanol (80%w/w - -- -- 1136.2 -- -- 568.1


recycle)

Calcium carbonate 13.325 -- -- 13.325 -- --

Calcium phosphate 33.53 -- -- 33.53 -- --

Sodium carbonate 6.812 -- -- 6.812 -- --

Magnesium carbonate 3.96 -- -- 3.96 -- --

Calcium lactate -- -- -- -- -- 99.821

Sodium lactate -- -- -- -- -- 14.40

Magnesium lactate -- -- -- -- -- 9.51

Lactic acid -- -- -- -- -- 29.16

Table 7. Energy requirements for the simulated processes of chitosan production

CDA-PC CDA-FPC
Operation
(MJ/cycle) (MJ/cycle)

Size reduction 0.879 0.879

14
Fermentation -- -389.66

Protein extraction reaction 777.6 388.8

Mineral extraction reaction -386.4 --

Deacetylation (1st stage) 176.69 176.69

Deacteylation (2nd stage) 82.94 82.94

Pumping & compression 189.66 179.9

Reactor agitation 6.70 10.44

Solvent evaporation 1570.21 763.04

Solvent condensation -1570.21 -763.04

4. TECHNO-ECONOMIC ANALYSIS

Techno-economic analysis was used for assessment of the operation costs, projection of
incomes, equipment specification and investment (among others), for the two alternatives of
chitosan production from shrimp cell waste, i.e., PC-CDA and FPC-CDA methods.
Manufacturing costs, investment, profits, taxes and other relevant information were determined
throughout a planning horizon of ten operating years. 13 For comparison purposes, the economic
assessment was performed under the same macroeconomic and demanding conditions, price
structure, debt and interest rate.

In order to compare both processes yielding the same annual production, requirement of
processing 1 batch/day was imposed. In the case of the FPC-CDA process, a systematic
debottlenecking study (Athimulam et al., 2006; Koulouris et al., 2000; Tan et al., 2006) was
necessary due to the large batch time (41h) for the fermentation stage. The fermenter presents the
largest combined utilization (73%) in the plant, therefore it was the target unit for application of
the debottlenecking strategy. Debottlenecking strategies in batch processes are focused on
reducing either the size or time utilization of the target procedure or equipment through
increasing the number of cycles per batch for the limiting procedure, rearrangement of the
equipment assignment or introducing new equipment (Koulouris et al., 2000).

The debottlenecking study on the FPC-CDA process was applied as indicated by Tan et al (Tan
et al., 2006), showing that two additional fermenters were required in order to reduce the
combined utilization of the first fermenter to 24%, yielding production levels equivalent to those

15
of the PC-CDA process i.e. 346 batches/year. The scheduling summary for the PC-CDA and
FPC-CDA processes are shown in the operation Gantt Charts (Figures 3 and 4).

Figure 3. Operations Gantt chart for PC-CDA process.

Figure 4. Operations Gantt chart for PC-CDA process.

Price of reactants and materials, considered in the first year of the plant operation, are presented
in Table 8. The prices were assigned to the inlet and outlet streams of the plant, based on trading
prices in the local market. The shrimp waste was considered as a "zero cost raw material” due to
its availability as residue from the shrimp culture industry; additionally, it was assumed that the
raw material is available at the plant location, i.e., it is supposed that the chitosan plant is located

16
nearby a shrimp processing company. Based on market studies from the literature, 4, 7 the selling
price of chitosan was fixed at 35 USD/kg.

Table 8. Reactant costs for the chitosan production

Material Price (USD/kg)

Ethanol-water solution (Depigmentation) 0.80

Sodium hydroxide solution (Deproteinization) 0.04

Sodium hydroxide solution (Deacetylation) 0.32

Hydrochloric acid solution (Demineralization) 0.05

Sugar aqueous solution (Fermentation) 0.16

The energy and utility costs were calculated according to the energetic requirements showed in
Table 7. For the first year, heating cost was projected in 0.026 USD/MJ, cooling cost in 0.035
USD/MJ and electricity in 0.07 USD/MJ.

Direct labor cost was assumed to be similar for both projects, since the required personnel is
supposed to be independent of the process. Operative expenses were calculated relative to sales
with the following distribution: 6% for management, 5% for marketing, 10% for national trading
and logistics, 13% for international trading and logistics. Taxes were accounted as 34% over
profits while interest rate as 26% over profits. Leverage equivalent to 60% of investment was
assumed. The results of costs, incomes and investment for the processes under study are
summarized in Tables 9 and 10.

Table 9. Costs and incomes for the simulated processes of chitosan production

PC-CDA FPC-CDA
Item
(USD/kg chitosan) (USD/kg chitosan)

Materials 7.520 7.560

Packaging 0.053 0.053

Labor 1.870 1.870

17
Energy & utilities 2.100 1.190

Total Production cost 11.540 10.680

Gross margin 68% 71%

Total Operative cost 7.790 7.790

Total costs 19.320 18.470

Total income 35 35

Table 10. Fixed investment required for the simulated processes of chitosan production

PC-CDA FPC-CDA
Investment
(Thousands USD/kg chitosan) (Thousands USD/kg chitosan)

Land 1.927 2.440

Building 2.890 3.660

Machinery and equipment 3.883 4.439

Preoperative 0.874 0.999

Vehicles and furniture 2.280 2.280

Total 11.855 13.941

When compared the PC-CDA process with the FPC-CDA process, it is noticed that the FPC-
CDA method necessitates lower amount of specific reactants and less energy requirements, e.g.,
water usage decreases in 38.66 L/kg-chitosan; the consumption of sodium hydroxide is reduced
in 25% and the hydrochloric acid requirement is eliminated (see Tables 6 and 7). The FPC-CDA
requires larger area and buildings, which increases 15% the investment in fixed assets (see Table
10) regarding the PC-CDA process.

Indicators of economic feasibility are presented in Table 11. It is observed that the estimated
cash flow performance in both projects is quite similar, as indicated by the net present value
(NPV), the payback period and the internal rate of return (IRR) values. Along with the planning
horizon, the economic assessment favored the chitosan production using the PC-CDA method,

18
because its NPV and IRR values are higher, while the payback period is 1 year lower than that in
the FPC-CDA method. However, chitosan production by the FPC-CDA method is more feasible
if the project is extended beyond 10 years (perpetuity); in such case, the net present value is
higher because the lower production costs makes up the larger investment requirements.
Regarding the product quality, both alternatives would render similar characteristics as was
confirmed by the experimental probes.

Table 11. Indicators of economic feasibility for the simulated processes of chitosan production

Indicator PC-CDA FPC-CDA

Cost of capital 13.3% 13.3%

Total capital cost (Millions USD) 0.7865 0.9166

NPV (Millions USD) 0.4977 0.4789

Perpetuity NPV (Millions USD) 1.0021 1.0596

IRR 26.6% 24.4%

Perpetuity IRR 30.8% 29%

Payback (years) 5 6

Gross margin 68% 71%

The major constraints for profitability of chitosan production are the material cost, investment
required and processing time; all of them are directly related with the quality of the final product.
Chitosan price depends on product quality given mainly by its deacetylation degree, water
solubility, viscosity, molecular weight and residual pigment, mineral and protein. Regarding the
price projection of 35 USD/kg-chitosan FOB and 68% of gross margin for the most feasible
process, the economic feasibility of a potential chitosan plant could be higher since the market
price for high quality chitosan ranges from 50 to 100 USD/kg-chitosan, 4, 7 having a significant
impact on profit projection and plant economic feasibility.

5. CONCLUSIONS

A techno-economic analysis for chitosan production from shrimp shell processing waste was
performed, using experimental lab-scale data and plant simulation results, using Aspen Plus®
v8.0 as simulation tool. Two different processing methods were compared, both of them with a
gross margin around 70%; thus, it can be argued that the implementation of one of these

19
alternatives would significantly contribute to give added value to the shrimp shell waste, partially
exploited in Latin American countries like Colombia.

Chitosan production presents drawbacks such as low volumetric yields with regard to raw
material and high water requirements. Application of water optimization techniques as the
proposed by Lee and Foo (2017) or Gouws et al. (2010) can lead to a more environmentally
friendly process. Fermentation and enzymatic processes also appear as prospective alternatives
with interesting economic and environmental performances for long-term projects related to
chitin and chitosan production. Even though techno-economic analyses reveal that chitosan
production from shrimp wastes is profitable and the specific processes can be further improved
for a more cost-competitive and efficient production. In addition to water minimization,
equipment and process intensification, the commercialization of valuable sub-products such as
pigments, protein hydrolysates and mineral salts must be considered. Notwithstanding, in the
Colombian and Latin American context, the low cost and large availability of raw material, as
well as the unsatisfied market, represent a stimulating perspective for the establishment of
chitosan industries in this geographical region.

ACKNOWLEDGMENT

This work was partially funded by Universidad Antonio Nariño Project No. 2015049. The
experimental procedures were funded by Departmento Administrativo de Ciencia, Tecnología e
Innovación – Colciencias - Colombia Project No. 111550227815.

ABBREVIATIONS

TEA, Techno-Economic Analysis; CAPE, Computer-Aided Process Engineering; PC, physical-


chemical; FPC, fermentative and physical-chemical; CDA, chemical deacetylation; DD,
deacetylation degree; FTIR, Fourier Transform Infrared Spectroscopy; prot, protein; min,
mineral; NPV, net present value; IRR, internal rate of return; FOB, free on board;

AUTHOR CONTRIBUTIONS
The manuscript was written through contributions of all authors. All authors have given approval
to the final version of the manuscript.

REFERENCES

Ahlafi, H., Moussout, H., Boukhlifi, F., Echetna, M., Bennani, M.N., Slimane, S.M., 2013.
Kinetics of N-Deacetylation of Chitin Extracted from Shrimp Shells Collected from Coastal
Area of Morocco. Mediterr. J. Chem. 2, 503–513. doi:10.13171/mjc.2.3.2013.22.01.20

Alshekhli, O., Foo, D.C.Y., Hii, C.L., Law, C.L., 2011. Process simulation and debottlenecking
for an industrial cocoa manufacturing process. Food Bioprod. Process. 89, 528–536.
doi:10.1016/j.fbp.2010.09.013

20
Ameh, A.O., Isa, M.T., Adeleye, T.J., Adama, K.K., 2013. Kinetics of demineralization of
shrimp exoskeleton in chitin and chitosan synthesis. J. Chem. Eng. Mater. Sci. 4, 32–37.
doi:10.5897/JCEMS13.0154

Aspen Technology Inc., 2015. Aspen Plus ® User Guide, Aspen Technology, Inc.

Athimulam, A., Kumaresan, S., Foo, D.C.Y., Sarmidi, M.R., Aziz, R.A., 2006. Modelling and
Optimization of Eurycoma longifolia Water Extract Production. Food Bioprod. Process. 84,
139–149. doi:10.1205/fbp.06004

Bajaj, M., Winter, J., Gallert, C., 2011. Effect of deproteination and deacetylation conditions on
viscosity of chitin and chitosan extracted from Crangon crangon shrimp waste. Biochem.
Eng. J. 56, 51–62. doi:10.1016/j.bej.2011.05.006

Cira, L.A., Huerta, S., Hall, G.M., Shirai, K., 2002. Pilot scale lactic acid fermentation of shrimp
wastes for chitin reco v ery. Process Biochem. 37, 1359–1366.

Cira, L.A., Huerta, S., Shirai, K., 2002. Fermentacion lactica de cabezas de camaron (. Rev.
Mex. Ing. Quim. 1, 45–48.

Czechowska-biskup, R., Jarosińska, D., Rokita, B., Ulański, P., Rosiak, J.M., 2012.
Determination of Degree of Deacetylation of Chitosan - Comparaison of Methods. Prog.
Chem. Appl. chitin its Deriv. 17, 5–20.

Duan, S., Li, L., Zhuang, Z., Wu, W., Hong, S., Zhou, J., 2012. Improved production of chitin
from shrimp waste by fermentation with epiphytic lactic acid bacteria. Carbohydr. Polym.
89, 1283–1288. doi:10.1016/j.carbpol.2012.04.051

Escobar, D., Ossa, C., Quintana, M., Ospina, W., 2013. Optimización de un protocolo de
extracción de quitina y quitosano desde caparazones de crustáceos. (Spanish). Optim.
Protoc. chitin chitosan Extr. from Crustac. shells. 18, 260–266.

Escobar, D., Urrea, C., Gutiérrez, M., Zapata, P., 2011. Producción de matrices de quitosano
extraído de crustáceos. Rev. Ing. Biomed. 5, 20–25.

Ghorbel-Bellaaj, O., Hmidet, N., Jellouli, K., Younes, I., Ma??lej, H., Hachicha, R., Nasri, M.,
2011. Shrimp waste fermentation with Pseudomonas aeruginosa A2: Optimization of chitin
extraction conditions through Plackett-Burman and response surface methodology
approaches. Int. J. Biol. Macromol. 48, 596–602. doi:10.1016/j.ijbiomac.2011.01.024

Global Industry Analysts Inc., 2014. Chitin and Chitosan - Global Strategic Business Report. San
José, California.

Gómez-Ríos, D., 2016. Technical and economic feasibility analysis of a plant process for
production of chitosan from shrimp shells in Colombia. Universidad de Antioquia.

Gouws, J.F., Majozi, T., Foo, D.C.Y., Chen, C.L., Lee, J.Y., 2010. Water minimization
techniques for batch processes. Ind. Eng. Chem. Res. 49, 8877–8893.

21
doi:10.1021/ie100130a

Goycoolea, F., Agulló, E., Mato, R., 2004. Fuentes y procesos de obtención, in: Pastor, A. (Ed.),
Quitina Y Quitosano: Obtención, Caracterización Y Aplicaciones. Pontificia Universidad
Católica del Perú, Lima, pp. 105–154.

Hernández, H., Águila, E., Flores, O., Viveros, E.., Ramos, E., 2009. Obtención y caracterización
de quitosano a partir de exoesqueletos de camarón. Superf. y Vacío 22, 57–60.

Herrera, M.A., Sánchez, D.I., López, J., Núñez, J.A., Moreno, O.H., 2011. Extracción de la
astaxantina y su estabilidad. Rev. Latinoam. Recur. Nat. 7, 21–27.

Instituto Colombiano Agropecuario, 2012. El sector camaronicultor colombiano: Evolución y


admisibilidad.

Khan, A.., Shibata, H., Kresnowati, P., Tai, S.L., 2001a. Conceptual process design: Production
of chitin and chitosan from shrimp shells. Delft University of Technology.

Khan, A.., Shibata, H., Kresnowati, P., Tai, S.L., 2001b. Production of chitin and chitosan from
shrimp shells. Delft University of Technology.

Koulouris, A., Calandranis, J., Petrides, D.P., 2000. Throughput analysis and debottlenecking of
integrated batch chemical processes. Comput. Chem. Eng. 24, 1387–1394.
doi:10.1016/S0098-1354(00)00382-3

Kwan, T.H., Pleissner, D., Lau, K.Y., Venus, J., Pommeret, A., Lin, C.S.K., 2015. Techno-
economic analysis of a food waste valorization process via microalgae cultivation and co-
production of plasticizer, lactic acid and animal feed from algal biomass and food waste.
Bioresour. Technol. 198, 292–299. doi:10.1016/j.biortech.2015.09.003

Lauer, M., 2008. Methodology guideline on techno economic assessment ( TEA ), ThermalNet
WP3B Economics.

Lee, J.-Y., Foo, D.C.Y., 2017. Simultaneous Targeting and Scheduling for Batch Water
Networks. Ind. Eng. Chem. Res. 56, 1559–1569. doi:10.1021/acs.iecr.6b03714

Moura, C.M. de, Moura, J.M. de, Soares, N.M., Pinto, L.A. de A., 2011. Evaluation of molar
weight and deacetylation degree of chitosan during chitin deacetylation reaction: Used to
produce biofilm. Chem. Eng. Process. Process Intensif. 50, 351–355.
doi:10.1016/j.cep.2011.03.003

Pacheco, N., 2010. Extracción biotecnológica de quitina para la producción de quitosanos :


caracterización y aplicación.

Pacheco, N., Garnica-González, M., Ramírez-Hernández, J.Y., Flores-Albino, B., Gimeno, M.,
Bárzana, E., Shirai, K., 2009. Effect of temperature on chitin and astaxanthin recoveries
from shrimp waste using lactic acid bacteria. Bioresour. Technol. 100, 2849–2854.
doi:10.1016/j.biortech.2009.01.019

22
Peniche, C., Argüelles-Monal, W., Goycoolea, F.M., 2008. Chitin and Chitosan: Major Sources,
Properties and Applications, in: Monomers, Polymers and Composites from Renewable
Resources. pp. 517–534. doi:10.1016/B978-0-08-045316-3.00013-2

Percot, A., Viton, C., Domard, A., 2003. Optimization of chitin extraction from shrimp shells.
Biomacromolecules 4, 12–18. doi:10.1021/bm025602k

Piccolo, C., Bezzo, F., 2009. A techno-economic comparison between two technologies for
bioethanol production from lignocellulose. Biomass and Bioenergy 33, 478–491.
doi:10.1016/j.biombioe.2008.08.008

Pillai, C.K.S., Paul, W., Sharma, C.P., 2009. Chitin and chitosan polymers: Chemistry, solubility
and fiber formation. Prog. Polym. Sci. 34, 641–678.
doi:10.1016/j.progpolymsci.2009.04.001

Rødde, R.H., Einbu, A., Vårum, K.M., 2008. A seasonal study of the chemical composition and
chitin quality of shrimp shells obtained from northern shrimp (Pandalus borealis).
Carbohydr. Polym. 71, 388–393. doi:10.1016/j.carbpol.2007.06.006

Shirai, K., 1999. Utilización de desperdicios de camarón para recuperación de quitina , proteínas
y pigmentos por vía microbiana. Universidad Autonoma Metropolitana Iztapalapa.

Sini, T.K., Santhosh, S., Mathew, P.T., 2007. Study on the production of chitin and chitosan
from shrimp shell by using Bacillus subtilis fermentation. Carbohydr. Res. 342, 2423–2429.
doi:10.1016/j.carres.2007.06.028

Sinnott, R., Towler, G., 2009. Costing and project evaluation, in: Chemical Engineering Design.
Elsevier, pp. 291–388.

Tan, J., Foo, D.C.Y., Kumaresan, S., Aziz, R.A., 2006. Debottlenecking of a batch
pharmaceutical cream production. Pharm. Eng. 26, 1–9.

Weska, R.F., Moura, J.M., Batista, L.M., Rizzi, J., Pinto, L.A.A., 2007. Optimization of
deacetylation in the production of chitosan from shrimp wastes: Use of response surface
methodology. J. Food Eng. 80, 749–753. doi:10.1016/j.jfoodeng.2006.02.006

23

You might also like