Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Research Article 1260

Dmitry Yu. Murzin1,*


Techno-Economic Analysis for Production
Emilien Daigue1
Robert Slotte1 of L-Arabitol from L-Arabinose
Dmitry A. Sladkovskiy2
Tapio Salmi1 The process design for synthesis of arabitol by hydrogenation of arabinose on a
supported ruthenium catalyst is described. Aspen HYSYS software was used to
design an arabitol production plant for subsequent generation of hydrogen
through aqueous phase reforming (APR). The process design included hydrogen
recycling requiring a flash drum and recompression. The total costs of arabitol
were estimated to be substantially dependent on the feedstock costs and are close-
ly related to arabinose feed cost. Feasible production of hydrogen using APR
requires efficient extraction of hemicellulose from lignocellulosic biomass and
subsequent hydrolysis.
This is an open access article under
the terms of the Creative Commons Keywords: Arabinose, Arabitol, Hydrogenation, Hydrogen production, Techno-economics
Attribution License, which permits
use, distribution and reproduction Received: March 11, 2020; revised: April 11, 2020; accepted: April 21, 2020
in any medium, provided the
original work is properly cited. DOI: 10.1002/ceat.202000125

1 Introduction with oxygen followed by downstream processing can be also


an option of hydrogen generation from biomass apparently
Utilization of biomass as a feedstock for making fuels and requiring also high temperatures.
chemicals has seen an enormous growth in the recent years In any case, hydroprocessing of a biomass-derived feedstock
[1–5]. This is related to a potential replacement of the fossil can involve on-site methane steam reforming for hydrogen
feedstock at least partially. Renewable feedstock can be utilized generation, (lignocellulosic) biomass gasification or can rely on
for generation of diesel fuels from triglycerides [1, 4], conver- purchase from external suppliers. An interesting alternative to
sion of cellulose and hemicelluloses present in lignocellulosic high-temperature steam reforming or biomass gasification is
biomass to biofuels [3], as well as synthesis of high-value-added aqueous phase reforming (APR) of various polyols, requiring
biobased products through so-called platform chemicals [2, 5]. much lower temperatures (225–250 °C) and thus resulting in
An important concern in application of future biorefineries lower energy consumption [15–17].
is related to their cost efficiency in comparison with established In the previous work of the authors, techno-economic analy-
processes in oil refining and petrochemistry based on crude oil sis was performed for APR of xylitol [18] and sorbitol [19]
and natural gas. As a result of such concerns, there have been using Aspen HYSYS software to design a 500 kg h–1 hydrogen
recently several studies addressing techno-economic evaluation production plant. The capacity was selected based on an esti-
of various process schemes related to different biorefinery pro- mated amount needed for 100 000 t of green diesel annual pro-
cesses [6–13]. Analysis of economic efficiency covered different duction from tall oil. Such facilities have been recently installed
levels of integration [6], production of biodiesel [7], ethanol in Finland [20]. The total costs of hydrogen were estimated to
biorefinery integrated with hydrothermal liquefaction [9], syn- be 2–3 fold higher than the costs of hydrogen from natural gas
thesis of succinic acid [12], hydrolysis strategy [13] as well as steam reforming, with the polyol price being the main cost
more general issues of economic feasibility of cellulosic bio- item (ca. 92 %). In a pursuit for more cost-efficient production
refineries. of polyols from lignocellulosic biomass it was decided to focus
In particular interesting is sustainable production of hydro- on the processes upstream of APR, namely, hydrogenation of
gen from renewable resources, e.g., from bioethanol by steam sugars and hydrolysis of hemicelluloses.
reforming [11]. In fact, hydrogen is required in substantial
amounts in biorefineries for various hydrodeoxygenation reac-
tions. A clear need for hydrogen is related with a high oxygen –
content in lignocellulosic biomass [1–3]. For example, hydro- 1
Prof. Dr. Dr. habil. Dmitry Yu. Murzin, Emilien Daigue, Robert Slotte,
treatment of rapeseed oil triglycerides requires H2 input of Prof. Dr. Tapio Salmi
7–16 moles depending on the dominating pathway (hydro- dmurzin@abo.fi
deoxygenation or decarboxylation) [14]. Hydrogen can be Åbo Akademi University, Laboratory of Industrial Chemistry and
produced by a number of methods including, e.g., electrolysis Reaction Engineering, Biskopsgatan 8, 20500 Turku/Åbo, Finland.
or steam reforming of natural gas. The latter is the most cost- 2
Dr. Dmitry A. Sladkovskiy
effective method [14] even if high temperatures are needed. St. Petersburg State Institute of Technology (Technical University),
Gasification of biomass with steam and subsequent reforming Moskovski pr. 26, 190013 St. Petersburg, Russia.

Chem. Eng. Technol. 2020, 43, No. 7, 1260–1267 ª 2020 The Authors. Published by WILEY-VCH Verlag GmbH & Co. KGaA. www.cet-journal.com
Research Article 1261

Several examples of various processes technologies involving ited hydrogen solubility, and a need to have a certain granule
hydrogen generation and hydrogenation per se, where techno- size for a reasonable pressure drop, mass transfer in trickle-bed
economic analysis was performed, could be found in the litera- reactors including the diffusion limitations inside the porous
ture describing, e.g., coal conversion to hydrogen [21, 22], car- catalyst particles can be severe [28]. As a consequence, egg-
bon dioxide hydrogenation to methanol [23] or integration of shell catalyst pellets with minimal internal mass transfer and
hydrogenation and dehydrogenation based on dibenzyltoluene structured catalysts can be viewed as viable options for contin-
as a liquid organic hydrogen energy carrier [24]. The authors uous hydrogenation of sugars.
are, however, not aware of any techno-economic analysis of In the current study, hydrogenation of arabinose over ruthe-
arabinose hydrogenation in a continuous reactor, while the nium on carbon was considered because carbons supports are
process flow diagram of xylose hydrogenation in a batch reac- reported to be superior compared to, e.g., alumina [28]. Carbon
tor over a Raney nickel catalyst was addressed in [21]. materials in general are attractive as catalyst supports because
The current work, therefore, considers hydrogenation of a their chemical and textural surface properties can be tuned for
pentose arabinose to arabitol as a case study and a part of the a particular reaction. Apparent drawbacks are poor mechanical
overall technology of hydrogen generation from lignocellulosic strength and large variability of the properties because of varia-
biomass, namely hemicelluloses. Arabinose can be produced by tions in the nature- derived feedstock. An alternative could be
hydrolysis of a hemicellulose-arabinogalactose (Fig. 1a). Hemi- carbon nanotubes or shaped porous carbons made from poly-
celluloses such as arabinogalactans appear in large quantities in mers [34–37]. Polymer-based spherical activated carbon has
larch species (e.g., Larix sibirica). The structural basis of arabi- been used, e.g., in hydrogenation reactions [38] and in APR of
nogalactan is a backbone of b-D-galactopyranose with D-galac- polyols [39].
topyranose and L-arabinofuranose side chains. The average Hydrogenation of arabinose is highly selective, thus for the
molar ratio of galactose to arabinose (Fig. 1b) in arabinogalac- purpose of techno-economic evaluation, a simple reaction stoi-
tan is about 6:1 and the molar mass is 20 000–100 000 g mol–1 chiometry was considered: C5H10O5 + H2 fi C5H12O5.
[26]. Extraction of arabinogalactan can be carried out in water
under moderate conditions on an industrial scale [27].
In fact, such sugars as arabinose and galactose obtained from 2.1 Process Scheme
hemicelluloses can be used in a variety of applications beyond
hydrogen generation, being in general a source for platform First, the mass balance for arabinose hydrogenation should
chemicals production after different transformations. Arabi- be considered. The process model was developed to process
nose (C5H10O5) is an aldopentose, a monosaccharide contain- 10 t h–1 of arabinose to represent a quantity needed to match
ing five carbon atoms and an aldehyde functional group. It is the required capacity in APR for generation of 500 kg h–1 of H2
used in food industry as a noncaloric sweetener being suitable [18, 19]. Taking into account complete conversion of arabinose
for diabetic patients to regulate their blood sugar content. and its solubility in water (834 g L–1 at 25 °C [40]), a minimum
water flow rate of ca. 12 t h–1 for the sugar aqueous feed solu-
tion is required. A certain safety margin should be introduced
2 Hydrogenation of Arabinose to avoid potential problems with solubility of either reactants
or the product, thus it was decided to operate with a sugar
Hydrogenation of sugars with molecular hydrogen is typically solution of 30 wt %. This implies feeding 23.3 t h–1 of water for
done in industrial scale in a batch mode using nickel catalysts, 10 t h–1 of sugar. Arabitol is also soluble in water up to 40 wt %
such as sponge nickel (Raney nickel) [28]. Recently, there is a in the temperature range of the process.
growing interest in application of ruthenium especially in the Hydrogenation was assumed to occur in a plug-flow reactor
continuous mode [28] because Ru exhibits good activity and on a Ru/C catalyst, active in hydrogenation of sugars [28]. Pre-
excellent selectivity and, moreover, is less toxic than Ni. viously batch-wise hydrogenation kinetics was investigated for
Continuous operation in hydrogenation of glucose has been a range of sugars including arabinose [41]. In general, the oper-
demonstrated in trickle-bed reactors using either pelletized or ation conditions for hydrogenation of various sugars (pentoses
structured catalysts, including monoliths and activated carbon and hexoses) were reported to be in the range of 90–130 °C and
cloths [29–33]. Because of a large size of sugar molecules, lim- 30–60 bar. More precisely, 100 °C at the reactor inlet and 50 bar

a) b)

Figure 1. Structure of (a) arabinogalactan and (b) arabinose and galactose.

Chem. Eng. Technol. 2020, 43, No. 7, 1260–1267 ª 2020 The Authors. Published by WILEY-VCH Verlag GmbH & Co. KGaA. www.cet-journal.com
Research Article 1262

were selected for the purpose of the present work as argued be- equation-of-state (EOS) was used [44]. One exception was the
low. The reaction is highly exothermal, thus 118 °C at the reac- separation unit where the predictive Sove-Redlich-Kwong EOS
tor outlet was reached. The temperature difference between the [45] was chosen, as it takes into account the polar interactions
inlet and the outlet is still relatively small not requiring imple- between molecules resulting in better flash calculations. The
mentation of a multitubular reactor. reactant and products L-arabinose and L-arabitol were intro-
While according to stoichiometry one mole of sugar requires duced to simulation as hypocomponents by the hypothetical
one mole of hydrogen to give one mole of the product arabitol, manager using the critical properties (Tab. 1) estimated by the
quite often hydrogenation reactions are conducted in the Joback method [46]. The thermodynamic package also in-
excess of hydrogen acting as a heat sink for highly exothermal cluded evaporation of water and its potential contribution for
reactions [42, 43]. An alternative option of a process without the enthalpy input.
excess of hydrogen would potentially lead to incomplete con-
version, a need to recycle the sugar and catalyst deactivation. Table 1. Reactant and product properties used for tuning the
Therefore, such option was not considered here. A typical thermodynamic basis.
molar excess of hydrogen applied in chemical industry and oil
refining is equal to 10:1 and implies 665 kmol h–1 (or 1.34 t h–1) Arabinose L-Arabinitol
of hydrogen entering the reactor. The excess of hydrogen Molecular weight 150 152
(598 kmol h–1 or 1.2 t h–1) should be separated from the product
stream and then re-pressurized and recycled into the hydro- Normal boiling point [°C] 441.86 494.5
genation reactor. To account for consumed hydrogen, its Ideal liquid density [kg m ]–3
1571 1476.0
make-up stream should be 66 kmol h–1 (134 kg h–1).
A detailed process flow diagram (PFD) with the simulation Critical temperature [C] 617.3 674.0
data is presented in Fig. 2. Critical pressure [kPa] 6578 6785.2
As displayed in Fig. 2, the detailed process scheme consists 3 –1 –1
Critical volume [m kg mol ] 0.343 0.39
of five main parts:
– The feed streams, namely, the sugar aqueous solution (fea- Acentricity 2.3 2.5
turing a pumping operation) and H2 make-up stream.
– A heat exchange unit upstream the reactor to heat up the
incoming sugar solution to the desired temperature and to
cool down the product mixture. While such heat integration 2.2.1 Feed Processing Upstream Reactor
by using a feed-effluent heat exchanger is necessary, it might
not be sufficient requiring an extra heat exchanger as will be The sugar feed stream containing 30 wt % of arabinose is
discussed below in the section on the simulation. pumped to the reactor at a total flow rate of 33.3 t h–1 at STP
– The reactor unit, namely, a trickle-bed reactor, which was (standard temperature and pressure) conditions. A centrifugal
simulated using a plug-flow reactor option available in pump increases the pressure of the sugar solution from atmo-
Aspen HYSYS. spheric pressure up to 50 bar. The properties of the pump were
– A flash tank to separate the gas (H2) from the aqueous solu- selected based on the standard values for industrial centrifugal
tion of arabitol. pumps resulting in the total electric power of 56 kW.
– Hydrogen recycling stream, including a compressing unit. The hydrogen stream at 20 °C and 50 bar accounts for the
hydrogen make-up (134 kg h–1). The hydrogen recycling stream
accounts for the unreacted hydrogen stream (1.22 t h–1) sepa-
2.2 Process Simulations with Aspen rated in the flash and recycled after being recompressed
through a compressor. After mixing the feed by mixers, the
For simulation of the phase equilibrium inside the reactor, a resulting total feed stream enters the heat transfer unit, com-
rigorous thermodynamic model based on the Peng-Robinson prising two heat exchangers. The first one is the feed effluent
heat exchanger (121 m2), simulated
under 25 °C minimum temperature
approach specification, while the
second is the steam preheater
(7 m2), providing an extra heat
needed to reach the specified tem-
perature of 100 °C with a steam
stream at 6.9 bar (164 °C) as a heat
source. To maximize the heating
efficiency and reduce energy con-
sumption, steam is considered to
be totally condensed requiring thus
a steam flow rate of just 472 kg h–1.
In the calculations, the hot stream
Figure 2. Arabinose hydrogenation process flow diagram. (reactor effluent) was cooled to

Chem. Eng. Technol. 2020, 43, No. 7, 1260–1267 ª 2020 The Authors. Published by WILEY-VCH Verlag GmbH & Co. KGaA. www.cet-journal.com
Research Article 1263

50 °C. However, to reach a specified final temperature of 30 °C, value of 0.01 mole min–1gRu–1 was achieved after 80 h time-on-
an extra cooler is required. stream (TOS) in hydrogenation of glucose. Because of mass
Analysis of the heat exchange performance using the tem- transfer limitations, the value of activation energy in the con-
perature-enthalpy diagram (Fig. 3) showed that the process can tinuous reactor in that case was somewhat lower than in the
operate without an external heat supply at a minimum temper- batch mode. The same approach was adopted in the current
ature approach lower than 18.9 °C and that the overlap of the work setting the value of activation energy to be 42 kJ mol–1,
composite curves leads only to the cooling utility. The amount which is lower that the value of 57.2 kJ mol–1 reported for
of the heat source (3.7 MW) is higher than the heat sink hydrogenation of arabinose in the batch reactor [36]. This ad-
(3.0 MW). Nevertheless, a minimum temperature approach of justment of the value of activation energy is not influencing the
25 °C was used in the process design which requires just a low- heat balance determined by the overall conversion and has an
capacity steam heater and, in any case, is needed for the start- impact only the temperature profile along the reactor length.
up operations. Fig. 4 besides displaying this temperature profile also displays
the concentration profiles using the pseudo-first-order kinetic
option in Aspen HYSYS.
The reactor inlet temperature of 100 °C was selected for the
process design being a compromise between the time to reach
complete conversion and a need not to use too high tempera-
tures. The adiabatic temperature rise was 18 °C. Such small
temperature difference allows implementation of a cheaper
reactor option, i.e., an adiabatic reactor compared to a more
expensive multitubular counterpart. Conversion close to 100 %
can be reached at arabinose weight hourly space velocity
(WHSV) of 1.6 h–1 for the specified mass flow rate giving the
catalyst volume of 6.25 m3. More precisely, an arabinose con-
Figure 3. Temperature-enthalpy diagram of the heat recovery version of 99.98 % gives an outlet stream of 34 690 kg h–1 con-
system.
taining arabinose and arabitol with the flows 2.5 kg h–1 and
10 132 kg h–1, respectively. The rest is water (23 344 kg h–1) and
2.2.2 Reactor hydrogen (1207 kg h–1).
Taking into account the catalyst bulk density of 400 kg m–3, a
As mentioned above, because of high selectivity a simple reac- catalyst load in the reactor is 2.5 t. The reactor was considered
tion stoichiometry was considered: C5H10O5 + H2 fi C5H12O5. as a vertical cylinder with a typical length-to-diameter ratio of
Previously, it was reported that kinetic profiles for hydrogena- 3, which thus results in the length of 4.2 m and the diameter of
tion of different sugars are very close to each other [41]. This 1.4 m.
allows utilization of the experimental data on continuous The pressure drop of 4.4 kPa was calculated by the Ergun
hydrogenation of another sugar – glucose over a Ru/C catalyst equation [47] setting the bed voidage at 0.4. The particle shape
in the plug-flow reactor [31] to adjust the catalyst productivity factor value was considered to be 0.5, while the diameter was
in the case of arabinose. It was demonstrated in [31] that the 7 mm.

Figure 4. Concentration and temperature profiles in the arabinose hydrogenation fixed-bed reactor.

Chem. Eng. Technol. 2020, 43, No. 7, 1260–1267 ª 2020 The Authors. Published by WILEY-VCH Verlag GmbH & Co. KGaA. www.cet-journal.com
Research Article 1264

2.2.3 Separation and Recycle 3 Process Cost Estimations


The aqueous solution of arabitol was separated from hydrogen The aim of the techno-economical analysis was to estimate the
using a flash drum resulting in the vapor stream and the liquid total costs of arabinose hydrogenation including both opera-
stream (Fig. 4). To perform the flash separation with Aspen, tional costs and capital investment. All the calculated values of
the predictive Sove-Redlich-Kwong (PSRK) model [45, 46] was costs are related to the specified input of arabinose of 10 t h–1.
used accounting for the polar interactions between molecules The fixed capital costs were estimated according to the Lang
to obtain more accurate results. factorial method [48]:
It seems preferable to operate the flash at a low temperature X 
of 30 °C and at 48.3 bar to save energy for hydrogen recycling C¼F Ce (1)
stream recompression. The flash mass balance under these
conditions (Tab. 2) shows that 99.8 % of hydrogen is recovered where C1) is the total plant capital cost, SCe is the total deliv-
into the vapor phase (V) to be recycled in the reactor after re- ered cost of all major equipment items: reactors, tanks, heat
compression. Tab. 2 also illustrates that all arabitol entering the exchangers, pumps, etc., and F is the installation (Lang) factor.
flash drum along with water and traces of unreacted arabinose The purchased equipment costs Ce referred to U.S. Gulf
is recovered in the liquid phase. A liquid residence time of Coast basis, January 2006 (CE index = 478.6, NF refinery infla-
15 min was set for the flash tank. The total tank volume is tion index = 1961.6) were calculated based on the following
15.5 m3 with the liquid load of 50 %. correlation:
The remaining element of the flow diagram is a compressor
to increase the pressure of the hydrogen recycle stream. The CE ¼ a þ bSn (2)
following values for the pressure drop were adopted for differ-
ent pieces of equipment: heat exchangers per pass 50 kPa, where a and b are the cost constants according to the type of
steam heater and cooler 25 kPa, separator 10 kPa. equipment, S is characteristic size parameter related to the
A centrifugal compressor can be used to increase the hydro- equipment, and n is the exponent for the type of equipment.
gen pressure from 48.3 to 50 bar. The simulations in HYSYS Constants for Eq. (2) were taken from [48]. The installed
were performed using a polytropic model with the values for equipment costs CIE were estimated from:
polytropic and mechanical efficiencies of respectively 75 % and
CIE ¼ CE fi ð575=478:6Þ (3)
95 % typical for industrial centrifugal compressors. The tem-
perature increase through the compressor is ca. 4 °C, while the
using the Hand’s installing factors fi (4 for pumps and vessels;
total power required by the compressor was 20.4 kW.
3.5 for exchangers; 2.5 for compressors [44]) and adjusted to
the year 2014 by using the CE index of 575. The reactor total
Table 2. Capital expenses calculations. purchase cost includes the cost of the shell (vertical pressure
vessel, including distributors) and the cost of the catalyst bed.
Item Cost [$] Fraction of FCI The catalyst costs were assumed to be 16 000 $ m–3, which
might be an underestimation. The costs of Ru were not, how-
Catalyst 100 000 2.4 ever, explicitly accounted for as it can be recovered, allowing
Reactor 786 824 19.0 full costs compensation after the end of the catalyst lifetime.
Based on these assumptions, the capital costs for the reactor
Feed-effluent heat exchanger 136 350 3.3
were 786 824 $, while the capital costs of the catalyst are respec-
Steam heater 69 797 1.7 tively 100 000 $. Values for other cost items are presented in
Tab. 2.
Separator 490 097 11.8
While fixed capital investments were based on the estimation
Compressor 171 206 4.1 of the purchase costs of the main equipment, the other costs
are estimated as factors of the total equipment cost. Addition
Pumps 219 940 5.3
of capital expenditure (CAPEX) and offsite costs (40 % of in-
Total ISBL 1 974 213 47.6 side battery limits (ISBL)) gave combined inside and outside
OSBL 789 685 19.0
battery limits (OSBL) costs of 2.7 MM$. On top of these costs,
engineering and construction costs (30 % of ISBL+OSBL) and
Engineering costs 829 169 20.0 contingency charges (20 % of ISBL+OSBL) should be included
Contingency charges 552 780 13.3 giving the fixed capital investment. Finally, the fixed capital an-
nual charge on 15-year basis and the working capital (addition-
Fixed capital investment 4 145 847 100.0 al money for starting the plant up and running it until some
Working capital 414 585 income is generated) as 15 % of ISBL+OSBL were also included
in the calculations. The total required investment was calcu-
Total required investment 5 942 380 lated to be 5.9 MM$ with 30 % precision (Tab. 3). Such calcula-


1) List of symbols at the end of the paper.

Chem. Eng. Technol. 2020, 43, No. 7, 1260–1267 ª 2020 The Authors. Published by WILEY-VCH Verlag GmbH & Co. KGaA. www.cet-journal.com
Research Article 1265

Table 3. Operational expenses calculations. or glucose) for production of the corresponding hexitols (galac-
titol and sorbitol).
Costs Cost [$ a–1] % of TCOP Different options were considered in [19] for production of
Variable costs of production hydrogen from sorbitol, which according to stoichiometry
should give the same amount of hydrogen and essentially fol-
Arabinose feedstock 50 060 000 89.5 lows the same pattern as galactitol [50]. Hydrogen costs were
Hydrogen 3 288 000 5.9 evaluated for different yields of hydrogen reflecting distribution
of carbon stored in the feedstock among different phases.
Electricity 55 408 0.1 When the reaction is driven to complete conversion of sorbitol
Steam 47 787 0.1 (or galactitol) to gaseous phase products without formation of
liquid products and alkanes, a so-called best-case scenario is
VCOP total 53 451 195 95.5 realized affording the yield of hydrogen of 13 moles per mole
Fixed costs of production 0.0 of a hexitol. This case affords the hydrogen costs of ca. 3 $ kg–1
when the sorbitol price is ca. 300 $ t–1. A more conservative
Operating labor 675 000 1.2
hydrogen yield of 7.3 moleH2molesorb–1 resulting in 10 % of
Supervision 168 750 0.3 carbon yield in the liquid phase requires that the sorbitol or
galactitol price for an APR process competitive in terms of
Direct salary overhead 337 500 0.6
hydrogen generation with methane steam reforming should be
Maintenance 98 712 0.2 lower than 150 $ t–1. This analysis highlights a need to perform
techno-economic analysis of the upstream processes to hydro-
Property taxes and insurance 39 485 0.1
genation, namely, extraction of hemicelluloses from lignocellu-
Rent of land 55 279 0.1 losic biomass and subsequent hydrolysis, which is currently in
General plant overhead 831 975 1.5
progress.

Environmental charges 27 639 0.0

Depreciation 276 394 0.5


4 Conclusion
FCOP total 2 510 734 4.5 The present study is focused on detailed techno-economic
Annual capital charge 396 159 0.7 analysis of pentoses and hexoses hydrogenation for production
of polyols, which can be subsequently used in hydrogen gener-
Total costs of production (TCOP) 55 961 929 100.0 ation by APR. Besides the process simulation also cost estima-
–1
Arabitol cost [$ t ] 700 tions of the major equipment (reactor, heat-exchangers, etc.)
were performed.
Reactor modeling was done based on the experimental data
for hydrogenation of arabinose at 100 °C and 50 bar of hydro-
tions might underestimate the real costs because mass transfer gen over 5%Ru/C catalyst. Techno-economical analysis using
limitations diminishing the catalyst effectiveness factor might Aspen HYSYS was performed to design a ca. 10 t h–1 arabitol
be important, increasing thereby the catalyst volume. A rigor- plant corresponding to 500 kg h–1 hydrogen production plant
ous modeling of a trickle-bed reactor for hydrogenation of serving as an option for H2 generation in remote locations
arabinose similar to the case of glucose reported recently [49] where lignocellulosic biomass is readily available. The total
is, however, outside of the scope of the present work. costs of arabitol are ca. 700 $ t–1 for an annual capacity of ca.
Variable costs of production were determined for 8000 h of 80 kt a–1, and are closely related to arabinose feed cost (ca.
production with the following cost factors: 22.4 $ MWh–1 for 89.5 %) when the arabinose price is 625 $ t–1. Generation of
medium pressure steam and 60 $ MWh–1 for electricity. The hydrogen from lignocellulosic biomass, which can be competi-
arabinose feedstock price was set at 625 $ t–1 (dry basis). tive with steam reforming of natural gas, requires a two-three
Hydrogen costs were calculated based on Department of Ener- fold decrease of the polyols costs and thus efficient extraction
gy (DOE) Hydrogen and Fuel Cells Program Record as of of hemicelluloses and subsequent hydrolysis.
September 2012 as 3 $ kg–1. The fixed costs include operating
labor (three shift positions and three shifts, 75 000 $ per person
on annual basis), supervision (25 % of operating labor), direct Acknowledgment
salary overhead (60 % of the operating labor), and maintenance
(5 % of ISBL investment). Other cost items are specified in DAS acknowledges funding through the grant of the Govern-
Tab. 3. ment of the Russian Federation for support of scientific re-
Costs for arabitol production are ca. 700 $ t–1 for an annual search conducted under the supervision of leading scientists at
capacity of ca. 80 000 t, and are closely related to the arabinose Russian institutions of higher education, research institutions
feed cost (ca. 89.5 %). of State Academies of Sciences, and State Research Center of
Based on similarities of kinetic data [41] it can be suggested the Russian Federation on March 19, 2014, No. 14.Z50.31.0013.
that the process design reported here is similar not only to pen-
toses (i.e., arabinose and xylose), but also to hexoses (galactose The authors have declared no conflict of interest.

Chem. Eng. Technol. 2020, 43, No. 7, 1260–1267 ª 2020 The Authors. Published by WILEY-VCH Verlag GmbH & Co. KGaA. www.cet-journal.com
Research Article 1266

Symbols used nery integrated with a hydrothermal liquefaction process to


convert lignin into biochemicals, Biofuels Bioprod. Bioref.
a [–] cost constant in Eq. (2) 2018, 12, 497–509.
b [–] cost constant in Eq. (2) [10] S. Gunukula, S. J. W. Klein, H. P. Pendse, W. J. De Sisto,
C [$] total plant capital cost M. C. Wheeler, Techno-economic analysis of thermal deoxy-
SCe [$] total delivered cost of all major genation based biorefineries for the coproduction of fuels
equipment items and chemicals, Appl. Energy 2018, 214, 16–23.
[11] M. Compagnoni, E. Mostafavi, A. Tripodi, N. Mahinpey,
CIE [$] installed equipment costs
I. Rossetti, Techno-economic analysis of a bioethanol to hy-
F [–] installation (Lang) factor
drogen centralized plant, Energy Fuels 2017, 31, 12988–
fi [–] Hand’s installing factors
12996.
S [–] characteristic size parameter related
[12] B. C. Klein, J. F. L. Silva, T. L. Junqueira, S. C. Rabelo, P. V.
to the equipment
Arruda, J. L. Ienczak, P. E. Mantelatto, J. G. C. Pradella, S. V.
Junior, A. Bonomi, Process development and techno-eco-
Superscript
nomic analysis of bio-based succinic acid derived from pen-
n exponent for the equipment time toses integrated to a sugarcane biorefinery, Biofuels Bioprod.
Bioref. 2017, 11, 1051–1064.
Abbreviations [13] W. Won, A. H. Motagamwala, J. A. Dumesic, C. T. Marave-
APR aqueous phase reforming lias, A co-solvent hydrolysis strategy for the production of
CAPEX capital expenditure biofuels: process synthesis and technoeconomic analysis,
DOE Department of Energy React. Chem. Eng. 2017, 2, 397–405.
[14] Green Vegetable Oil Processing (Eds: N. C. Easterbrook, W. E.
FCOP fixed costs of production
Farr), Elsevier, New York 2013.
ISBL inside battery limits
[15] R. Cortright, R. Davda, J. A. Dumesic, Hydrogen from cata-
OSBL outside battery limits
lytic reforming of biomass-derived hydrocarbons in liquid
TCOP total costs of production
water, Nature 2002, 418, 964–976.
VCOP variable costs of production
[16] R. R. Davda, J. W. Shabaker, G. W. Huber, R. D. Cortright,
J. A. Dumesic, A review of catalytic issues and process con-
ditions for renewable hydrogen and alkanes by aqueous-
References phase reforming of oxygenated hydrocarbons over support-
ed metal catalysts, Appl. Catal., B 2005, 56, 171–186.
[1] A. Srivastava, R. Prasad, Triglycerides-based diesel fuels, [17] I. Coronado, M. Stekrova, M. Reinikainen, P. Simell, L. Lef-
Renewable Sustainable Energy Rev. 2000, 4, 111–133. ferts, J. Lehtonen, A review of catalytic aqueous-phase
[2] J. J. Bozell, G. R. Petersen, Technology development for the
reforming of oxygenated hydrocarbons derived from bio-
production of biobased products from biorefinery carbo-
refinery water fractions, Int. J. Hydrogen Energy 2016, 41,
hydrates – US Department of Energy’s ‘‘Top 10’’ revisited,
11003–11032.
Green Chem. 2010, 12, 539–554. [18] D. Yu. Murzin, S. Garcia, V. Russo, T. Kilpiö, L. I. Godina,
[3] D. M. Alonso, J. Q. Bond, J. A. Dumesic, Catalytic conver-
A. V. Tokarev, A. V. Kirilin, I. L. Simakova, S. Poulston,
sion of biomass to biofuels, Green Chem. 2010, 12, 1493–
D. A. Sladkovskiy, J. Wärnå, Kinetics, modelling and process
1513.
design of hydrogen production by aqueous-phase reforming
[4] D. Y. Murzin, I. L. Simakova, Catalysis in biomass conver-
of xylitol, Ind. Eng. Chem. Res. 2017, 56, 13240–13253.
sion, in Comprehensive Inorganic Chemistry II: From Ele-
[19] D. A. Sladkovskiy, L. I. Godina, K. V. Semikin, E. V. Sladkov-
ments to Applications, (Eds: J. Reedijk, K. Poeppelmeier),
skaya, D. A. Smirnova, D. Yu. Murzin, Process design and
Vol. 7, Elsevier, Amsterdam 2013, 559–586.
techno-economical analysis of hydrogen production by
[5] A. Corma, S. Iborra, A. Velty, Chemical routes for the trans-
aqueous phase reforming of sorbitol, Chem. Eng. Res. Des.
formation of biomass into chemicals, Chem. Rev. 2007, 107,
2018, 134, 104–116.
2411–2502.
[20] J. Laurikko, N. Nylund, P. Aakko-Saksa, S. Mannonen,
[6] C. A. Garcia-Velasquez, V. Aristizabal-Marulanda, C. A. Car-
V. Vauhkonen, P. Roslund, Crude tall oil-based renewable
dona, Analysis of bioenergy production at different levels of
diesel in passenger car field test, SAE Tech. Paper 2014,
integration in energy-driven biorefineries, Clean Technol.
2014-01-2774.
Environ. Policy 2018, 20, 1599–1613.
[21] M. Aziz, I. Nuran Zaini, T. Oda, A. Morihara, T. Kashiwagi,
[7] C. D. B. Gutierrez, D. L. R. Serna, C. A. C. Alzate, A compre-
Energy conservative brown coal conversion to hydrogen and
hensive review on the implementation of the biorefinery
power based on enhanced process integration: Integrated
concept in biodiesel production plants, Biofuel Res. J. 2017,
drying, coal direct chemical looping, combined cycle and hy-
4, 691–703.
drogenation, Int. J. Hydrogen Energy 2017, 42, 2904–2913.
[8] T. R. Brown, Price uncertainty, policy, and the economic fea-
[22] M. Aziz, F. B. Juangsa, W. Kurniawan, B. A. Budiman, Clean
sibility of cellulosic biorefineries, Biofuels, Bioprod. Bioref.
co-production of H2 and power from low rank coal, Energy
2018, 12, 485–496.
2016, 116, 489–497.
[9] D. Bbosa, M. Mba-Wright, R. C. Brown, More than ethanol:
a techno-economic analysis of a corn stover ethanol biorefi-

Chem. Eng. Technol. 2020, 43, No. 7, 1260–1267 ª 2020 The Authors. Published by WILEY-VCH Verlag GmbH & Co. KGaA. www.cet-journal.com
Research Article 1267

[23] A. Alsayegh, J. R. Johnson, B. Ohs, M. Wessling, Methanol [36] Q. Wang, X. Liang, W. Qiao, C. Liu, X. Liu, L. Zhan, L. Ling,
production via direct carbon dioxide hydrogenation using Preparation of polystyrene-based activated carbon spheres
hydrogen from photocatalytic water splitting: Process devel- with high surface area and their adsorption to dibenzothio-
opment and techno-economic analysis, J. Cleaner Prod. phene, Fuel Process. Technol. 2009, 90, 381–387.
2019, 208, 1446–1458. [37] A. J. Romero-Anaya, M. Ouzzine, M. A. Lillo-Ródenas,
[24] M. H. Hamayun, M. Hussain, I. M. Maafa, R. Aslam, Inte- A. Linares-Solano, Spherical carbons: Synthesis, character-
gration of hydrogenation and dehydrogenation based on ization and activation processes, Carbon 2014, 68, 296–307.
dibenzyltoluene as liquid organic hydrogen energy carrier, [38] H. Klefer, M. Munoz, A. Modrow, B. Böhringer, P. Wasser-
Int. J. Hydrogen Energy 2019, 44, 5345–5354. scheid, B. J. M. Etzold, Polymer-based spherical activated
[25] A. D. Mountraki, K. R. Koutsospyros, B. Benjelloun Mlayah, carbon as easy-to-handle catalyst support for hydrogenation
A. C. Kokossis, Selection of biorefinery routes: the case of reactions, Chem. Eng. Technol. 2016, 39, 276–284.
xylitol and its integration with an organosolv process, Waste [39] L. I. Godina, A. V. Tokarev, I. L. Simakova, P. Mäki-Arvela,
Biomass Valorization 2017, 8, 2283–2300. E. Krzymyk, J. Gläsel, L. Kronberg, B. Etzold, D. Yu. Murzin,
[26] S. Willför, R. Sjöholm, C. Laine, B. Holmbom, Structural fea- Aqueous phase reforming of alcohols with three carbon
tures of water-soluble arabinogalactans from Norway spruce atoms on carbon supported Pt, Catal. Today 2018, 301, 78–
and Scots pine heartwood, Wood Sci. Technol. 2002, 36, 101– 89.
110. [40] R. C. Weast, CRC Handbook of Chemistry and Physics, 62nd
[27] S. Willför, B. Holmbom, Isolation and characterisation of ed., CRC Press, Boca Raton, FL 1981, p. C-110.
water-soluble polysaccharides from Norway spruce and [41] V. A. Sifontes Herrera, O. Oladele, K. Kordas, K. Eränen,
Scots pine, Wood Sci. Technol. 2004, 38, 173–179. J.-P. Mikkola, D. Yu. Murzin, T. Salmi, Sugar hydrogenation
[28] D. Yu. Murzin, A. Duque, K. Arve, V. Sifontes, A. Aho, over a Ru/C catalyst, J. Chem. Tech. Biotechnol. 2011, 86,
K. Eränen, T. Salmi, Catalytic hydrogenation of sugars, in 658–668.
RSC Biomass Sugars for Non-Fuel Applications (Ed.: D. Yu. [42] Chemical Process Technology, 2nd ed. (Eds: J. A. Moulijn,
Murzin, O. Simakova), RSC Green Chem. 2016, 44, 89–133. M. Makkee, A. E. van Diepen), Wiley, New York 2013.
[29] J. Wisniak, R. Simon, Hydrogenation of glucose, fructose, [43] D. Yu. Murzin, Chemical Reaction Technology, de Gruyter,
and their mixture, Ind. Eng. Chem. Prod. Res. Dev. 1979, 18, Berlin 2015.
50–57. [44] D.-Y. Peng, D. B. Robinson, A new two-constant equation of
[30] N. Déchamp, A. Gamez, A. Perrard, P. Gallezot, Kinetics of state, Ind. Eng. Chem. Fundam. 1976, 15, 59–64.
glucose hydrogenation in a trickle-bed reactor, Catal. Today [45] K. Fischer, J. Gmehling, Further development, status and re-
1995, 24, 29–34. sults of the PSRK method for the prediction of vapor-liquid
[31] A. Aho, S. Roggan, K. Eränen, T. Salmi, D. Yu. Murzin, Con- equilibria and gas solubilities, Fluid Phase Equilib. 1995, 112,
tinuous hydrogenation of glucose with ruthenium on carbon 1–22.
nanotubes catalysts, Catal. Sci. Technol. 2015, 5, 953–959. [46] K. G. Joback, R. C. Reid, Estimation of pure-component
[32] J. Hilgert, S. Lima, A. Aho, K. Eränen, D. Yu. Murzin, R. Ri- properties from group contributions, Chem. Eng. Commun.
naldi, On the impact of salts by the neutralisation of (ligno)- 1987, 57, 233–243.
cellulose hydrolysates on the hydrogenation of sugars, [47] S. Ergun, Fluid flow through packed columns, Chem. Eng.
ChemCatChem 2018, 10, 2409–2416. Prog. 1952, 48, 89–94.
[33] B. J. Arena, Deactivation of ruthenium catalysts in continu- [48] G. Towler, R. Sinnott, Chemical Engineering Design – Princi-
ous glucose hydrogenation, Appl. Catal., A 1992, 87, 219– ples, Practice and Economics of Plant and Process Design,
229. Elsevier, New York 2008.
[34] Z. Zhu, A. Li, S. Zhong, F. Liu, Q. Zhang, Preparation and [49] T. Kilpiö, A. Aho, D. Murzin, T. Salmi, Experimental and
characterization of polymer-based spherical activated car- modeling study of catalytic hydrogenation of glucose to sor-
bons with tailored pore structure, J. Appl. Polym. Sci. 2008, bitol in a continuously operating packed bed reactor, Ind.
109, 1692–1698. Eng. Chem. Res. 2013, 52, 7590–7770.
[35] B. Böhringer, O. Guerra Gonzalez, I. Eckle, M. Müller, J.-M. [50] L. I. Godina, A. V. Kirilin, A. V. Tokarev, D. Yu. Murzin,
Giebelhausen, C. Schrage, S. Fichtner, Polymer-based spheri- Aqueous phase reforming of industrially relevant sugar alco-
cal activated carbons from adsorptive properties to filter per- hols with different chirality, ACS Catal. 2015, 5, 2989–3005.
formance, Chem. Ing. Tech. 2011, 83, 53–60.

Chem. Eng. Technol. 2020, 43, No. 7, 1260–1267 ª 2020 The Authors. Published by WILEY-VCH Verlag GmbH & Co. KGaA. www.cet-journal.com

You might also like