SlideShare ist ein Scribd-Unternehmen logo
1 von 11
Downloaden Sie, um offline zu lesen
Propylene glycol from glycerol: Process evaluation and break-even
price determination
Roberto X. Jimenez a
, Andre F. Young b, *
, Heloisa L.S. Fernandes a
a
Department of Chemical Engineering, Federal University of Rio de Janeiro, Av. Athos da Silveira Ramos, 149, Rio de Janeiro, Brazil
b
Department of Chemical and Petroleum Engineering, Federal Fluminense University, Rua Passo da Patria, 156, Niteroi, Brazil
a r t i c l e i n f o
Article history:
Received 22 November 2019
Received in revised form
27 April 2020
Accepted 23 May 2020
Available online 31 May 2020
Keywords:
Glycerol
Process simulation
Propylene glycol
Hydrogen
Reforming
a b s t r a c t
Propylene glycol can be produced from the hydrogenolysis of glycerol. If hydrogen comes from a
renewable source, the produced propylene glycol can be considered completely renewable. The objective
of this work is to perform a technical and economic evaluation of a propylene glycol production process
which uses glycerol as the main raw material. Two possibilities regarding the hydrogen source are
explored: it may come from an external source or be produced locally. It is shown that the break-even
price of the partially renewable propylene glycol is lower than the market price of propylene glycol,
which indicates that this process is economically viable. It is also shown that the price of hydrogen
produced via steam reforming is not competitive with its price if produced by other sources. Conse-
quently, the fully renewable propylene glycol production process yields a lower profit, but it is still viable
and depends on the market.
© 2020 Elsevier Ltd. All rights reserved.
1. Introduction
Glycerol is the main by-product of biodiesel production [1e3].
Along with the increased supply of biodiesel, glycerol production
has also increased. In 2017, there were produced about 374.5
thousand cubic meters of glycerol, which meant a 9.5% increase in
comparison with 2016 [4].
The increasing glycerol surplus has motivated the development
of chemical processes which employ it as the main raw material. As
a consequence, the number of published scientific articles on the
use of glycerol as the main raw material in chemical processes has
increased. For example, articles related to glycerol in the Scopus
database have increased by almost 7 times per year between 2008
and 2016 [3].
Among the products that can be obtained from glycerol there
are propanediols (1,2- and 1,3-propanediol), propene, epichloro-
hydrin, methanol, ethanol, lactic acid, citric acid, propanoic acid
and hydrogen [5e7]. Monteiro [3] and Freitas [6] identified that 1,3-
propanediol, hydrogen and 1,2-propanediol or propylene glycol
(PG) are the products formed from glycerol on which researchers
have published more articles. According to these authors, the
number of patents or patent applications on PG production from
glycerol was larger than those on 1,3-propanediol or hydrogen
production until 2013.
Propylene glycol can be used as a surfactant, a moisturizer, an
antifreeze agent, a solvent, a preservative and in detergent for-
mulations [6,8]. It is usually obtained by the hydrolysis of propylene
oxide [8,9]. Propylene oxide is produced from propylene hydro-
chloride, which comes from the petrochemical industry.
In Brazil, the only registered producer of propylene glycol is
Dow®. It has a chemical plant located in the industrial complex of
Aratu, in the state of Bahia. The company also produces propylene
oxide, so propylene glycol is still produced by the conventional
route from petrochemicals. It is noteworthy that this is the only
propylene glycol production plant in Latin America [10]. A simpli-
fied flowsheet of this process technology is shown in Fig. 1.
Alternatively, PG can be produced by a more ecologically sound
route, which employs glycerol as a carbon and hydrogen source,
and which relies less on fossil sources. Some scientific articles about
this route were published, but there are only a few patents (granted
or under application) on it. Currently, if USPTO and Lens databases
are searched for patents on processes involving the keywords
propylene glycol, glycerol/glycerin and 1,2-propanediol, glycerol/
glycerin, only 12 results are retrieved. Among them, 10 patents
describe processes with liquid-phase reactors, one that employs a
vapor-phase reactor and one with a two-phase reactor (reactions* Corresponding author.
E-mail address: ayoung@id.uff.br (A.F. Young).
Contents lists available at ScienceDirect
Renewable Energy
journal homepage: www.elsevier.com/locate/renene
https://doi.org/10.1016/j.renene.2020.05.126
0960-1481/© 2020 Elsevier Ltd. All rights reserved.
Renewable Energy 158 (2020) 181e191
start in the liquid phase and generate acetol, which is transferred to
the vapor phase to react). It is interesting to note that, although the
search was not limited by date, all the results date to the last 8
years.
These results indicate that this technology is new and may
require more studies until consolidation. Considering this, the
objective of this work is to perform a technical and economic
assessment of the partially renewable and fully renewable PG
production processes from glycerol, to find out if these processes
are economically viable in the current Brazilian economic scenario.
Equipment design and process simulations were performed with
the aid of the Aspen HYSYS® v8.8 process simulator.
2. Process simulation and equipment design
In order to proceed to the economic analysis, the reaction ki-
netics and vapor-liquid equilibria must be correctly reproduced in
the simulations so that the equipment can be properly designed.
The properties of all components involved in the system are
available in the Aspen HYSYS® simulator database.
2.1. Propylene glycol production from glycerol
Glycerol (or glycerin) obtained from biodiesel is available in
different purity levels. Crude glycerin is defined as the glycerol-rich
mixture (around 75% glycerol) obtained after the separation of
biodiesel. After neutralization, glycerin is usually called blonde; its
weight percent composition is 85.0% glycerol, 11.7% water, 3.2%
methanol, and 0.2% biodiesel. After distillation, USP grade glycerol
or bidistilled glycerol (99.5% glycerol) is obtained [11]. The advan-
tage of using blonde glycerin as a raw material is that, as the puri-
fication process has a cost, it ends up being cheaper. However,
bidistilled glycerol is regarded as the process raw material in this
work, since kinetic data for the hydrogenolysis and reforming re-
actions refer to glycerol-water mixtures only [12].
Furthermore, according to Resolution no. 30 published by ANP
(Brazilian National Agency of Petroleum, Natural Gas and Biofuels),
which regulates biodiesel production activities in Brazil [4],
including the construction stage, the glycerol purification stage
must be present in every biodiesel production unit. Thus, in all
plants currently operating in Brazil, a stream of composition close
to that of bidistilled glycerol is available, or at least its composition
can be adjusted for this purpose.
In literature, a wide variety of metal catalysts has been used to
promote the CeO bond cleavage reaction of the glycerol molecule.
Cu-based catalysts are widely studied as they offer a high degree of
conversion and selectivity of PG [13e20].
Two mechanisms have been proposed for the production of
propylene glycol from glycerol: according to the dehydration-
hydrogenation mechanism, glycerol is initially dehydrated at
catalyst acid sites to acetol, which is hydrogenated to 1,2-
propanediol [5,13,21]. A three-phase dehydrogenation-dehydra-
tion-hydrogenation mechanism is also suggested [15,17,22,23].
The literature recommends a Langmuir-Hinshelwood-type
expression to model the heterogeneous kinetics of formation of
propylene glycol and by-products (1,3-propanediol, ethylene glycol,
ethanol, methanol, propanol and propanoic acid) through the
dehydration-hydrogenation mechanism [12,19,24,25]. Two types of
catalyst active sites are considered, as observed in Table 1 [12,26].
Zhou et al. [12] employed a CueZnOeAl2O3 catalyst with a
molar ratio of 1:1:0.5 of Cu/Zn/Al, with a particle diameter of
0.17 mm (80e100 mesh), and reported a conversion of 81.5% of the
glycerol and selectivity of 93.4% in terms of propylene glycol. Also,
the authors proposed the rate expressions of the two reaction steps
presented in Table 1, which can be found in Eqs. (1) and (2).
r1 ¼
k1bGCG
ð1 þ bGCG þ bACA þ bPCPÞ
(1)
r2 ¼
k2bGCG
ð1 þ bGCG þ bACA þ bPCPÞ

1 þ
ffiffiffiffiffiffiffiffiffiffiffiffi
bHPH
p 2
(2)
In Eqs. (1) and (2), Ci is the molar concentration of component i,
PH is the partial pressure of hydrogen, ki is the rate constant of
reaction i and bi is the absorption rate constant of component i [12].
The numerical values of these constants are described in Table 2. A
preliminary simulation of the reaction in laboratory conditions was
carried out in Aspen HYSYS®; simulated and experimental data
[12] were compared and relative errors of 3.2% in the glycerol
conversion and 7.0% in the acetol to propylene glycol selectivity
ratio were found. Since average relative errors between experi-
mental and theoretical glycerol and propylene glycol flowrates at
Fig. 1. Simplified propylene oxide hydration process flowsheet.
Table 1
Steps involved in the reaction of propylene glycol formation over a Cu/ZnO/Al2O3
catalyst. G ¼ glycerol, A ¼ acetol, PG ¼ propylene glycol, Q1 type 1 active site, Q2
type 2 active site [12].
Stage Kinetic Adsorption/Desorption
H2 þ 2Q1 #2HQ1 e bH
G þ Q2 #GQ2 e bG
GQ2/AQ2 þ H2O k1 e
2HQ1 þ AQ2 /PGQ2 þ 2Q1 k2 e
AQ2 #A þ Q2 e bÀ1
A
PGQ2 #P þ Q2 e bÀ1
PG
R.X. Jimenez et al. / Renewable Energy 158 (2020) 181e191182
the reactor output are 6.3% and 7.6%, respectively [12], deviations
obtained in the preliminary simulation can be considered accept-
able. Thus, the rate expressions in Eqs. (1) and (2) are used in this
work.
The UNIQUAC model was used for phase equilibria calculations
for the system containing the components involved in the pro-
duction of propylene glycol, since it provides a satisfactory
description of phase equilibria for similar systems [27].
The propylene glycol process flowsheet can be seen in Fig. 2. The
stream properties are presented as Supporting Information,
Table S3. The raw material is fed into the process and is put under
operating conditions before entering the reactor. Excess hydrogen
is used to facilitate mass transfer in the three-phase system within
the reactor. The global reaction is shown in Eq. (3).
C3H8O3 þ H2/C3H8O2 þ H2O
Glycerol PG
(3)
The reaction is carried out at T ¼ 493 K, PH ¼ 4 MPa and a
hydrogen to glycerol molar ratio of 5:1. The catalyst is assumed to
be CueZnOeAl2O3 and the residence time is adjusted to 85
kgCat:h:kg:molÀ1
Glycerol, in order to reproduce the results obtained by
Zhou et al. [12]. Moreover, the authors compared their results at
493 K with others available in the literature [21], and showed
consistency in the values that were reported. Zhou et al. [12] proved
that glycerol conversion increases with higher pressures (testing
3 MPa, 4 MPa and 5 MPa), but at 493 K it remains almost the same,
independent on pressure. A conservative approach was used to
choose a 4 MPa pressure for the reaction simulation. The plant was
designed to operate with a flowrate of 1070 kg/h of glycerol. This
was the average glycerol production in biodiesel plants in Brazil in
2018 [4].
The reactor effluent is mainly composed by propylene glycol,
hydrogen, unreacted glycerol and by-products. The product stream
exchanges heat with the glycerol stream to minimize utility ex-
penses. Then it follows to the separation section, in which excess
hydrogen is recovered with a flash vessel, recycled and mixed with
pure hydrogen, thereby ensuring a 5:1 M ratio of hydrogen to
glycerol at the process inlet. Finally, two distillation columns adjust
the propylene glycol purity to the required value (USP grade,
99.5%).
The literature suggests the use of a trickle-bed-reactor for
glycerol hydrogenolysis in industrial-scale processes [25,28,29].
Inside the reactor, liquid and gas flow downward through the cat-
alytic bed. The net flowrate must be low enough so that the liquid
phase forms a thin film over the fixed bed and descends by gravity
and gas-phase drag [30]. Thus, the contact among the three phases
is maximized. Trickle-bed-reactors have been widely used in other
processes involving hydrogen, mainly in the hydrodesulfurization
of petroleum fractions [30,31].
The catalyst density was not mentioned in the articles reviewed
but is required for the calculation of the fixed-bed volume. The
fixed-bed volume includes the volume occupied by the catalyst
solid mass (30%) and the volume of pore voids and voids among the
catalyst particles (70%), which are generally in a proportion of 3:3:4
[32]. The solid density was calculated from the molar fraction of its
components. Assuming the void fraction in the bed to be equal to
0.4, the densities of solid, pellets and bulk can be estimated. For the
purpose of reactor sizing, the trickle-bed model was simplified to a
plug-flow fixed bed model, as available in Aspen HYSYS®.
The required catalyst mass was obtained by solving the design
equations for a plug-flow reactor. A plug-flow reactor which
operates at conditions similar to those in the work by Zhou et al.
[12] requires 992.8 kg of catalyst. Likewise, a conversion of 98.2%
requires 1489 kg of catalyst, while a conversion of 99.99% requires
an additional 993 kg of catalyst [12]. Consequently, it is not
convenient to try to achieve such high glycerol conversions. Thus,
the plant was designed to operate with an overall glycerol
Table 2
Parameters for the kinetic model of the glycerol hydrogenolysis reaction [12].
Kinetic and absorption constants Parameters
Pre-exponential Factor

mol
kg s

Activation Energy

kJ
mol

k1 1.54 Â 107
86.56
k2 7.16 Â 106
57.8
bG 2.22 36.42
bA 8.73 25.94
bPG 5.80 25.77
bH 1.86 Â 10À5
MPa 36.24
Fig. 2. Propylene glycol production flowsheet.
R.X. Jimenez et al. / Renewable Energy 158 (2020) 181e191 183
conversion of 98.2%. Finally, the reactor volume can be calculated
by the ratio of catalyst mass and bulk density.
The reaction must occur under isothermal conditions. To meet
this specification, a cooling jacket was applied, with water at 298 K
as coolant. The heat exchange area was assumed to be approxi-
mately equal to the lateral area of the plug-flow reactor.
In addition to acetol and propylene glycol, the authors identified
the presence of by-products such as ethylene glycol, methanol,
ethanol, propanol and propanoic acid in the product stream, among
which ethylene glycol is the most expressive. Therefore, it is
necessary to simulate the formation of these products so that the
simulation is in agreement with the experimental study.
Ethylene glycol and methanol modeling was performed by
specifying the conversion value found experimentally [12]. Due to
the lack of information on the proportion of the remaining by-
products, their simulation was not feasible. Since the by-products
are lower alcohols (except for propanoic acid), they were regar-
ded as a single representative compound (methanol) in the
simulations.
In a preliminary analysis of the separation system, it was found
that for a better recovery of unreacted hydrogen it is necessary to
cool the product stream from the reactor. To save resources and
accomplish energy integration, it was decided to implement a
shell-and-tube heat exchanger through which this stream and the
glycerol and water feed stream are passed. The heat transfer rate,
the heat exchange area and the global heat transfer coefficient were
obtained with the aid of the simulator, from the values of the inlet
and outlet streams’ temperatures, pressures and flowrates.
A minimum temperature approach of 10 K was used to avoid an
excessively large heat exchange area. The initial estimate of pres-
sure drop on both sides of the exchanger was 10 kPa [33]. The heat
exchanger is designed with floating heads to minimize thermal
expansion effects that may occur for a temperature difference
greater than 30 K between the tube fluid (reactor effluent) and the
shell fluid [34].
The methodology described by Seider et al. [33], which is based
on the relative volatility of the compounds, was adopted for the
synthesis of the separation system. The reactor outlet is a two-
phase stream, which consists of hydrogen, methanol, water, ace-
tol, propylene glycol, ethylene glycol and glycerol (from the most
volatile to the least volatile). Three cuts were required to achieve
the specified purity: Hydrogen/Methanol, Water/Propylene Glycol
and Propylene Glycol/Ethylene Glycol.
Hydrogen is the first component that must be removed, as it is
the excess reagent and can be recycled into the process, thereby
reducing raw materials expenditures. Thus, for the separation of
hydrogen from the reactor effluent stream, a flash vessel was used.
The liquid-gas separator should be such that the vessel diameter
must be large enough so that the gas velocity is less than the liquid
droplet separation velocity and that the vessel height is equal to its
diameter, or 1 m if the diameter is less than 1 m. It is recommended
that hold-up time is about 10 min so that the liquid level inside the
equipment is appropriate. Finally, the feed inlet must be at least
60 cm above the liquid level [35].
To maximize the recovery of pure hydrogen, it is necessary that
the flash vessel also removes heat from the streams, to promote
condensation of the less volatile components. Such heat exchange
was performed by implementing a coil inside the vessel with
cooling water at 298 K. Aspen HYSYS® Adjust block was used so that
the cooling rate varies to ensure that the purity of the hydrogen
obtained is at least 99.5%.
The next two cuts were made through distillation columns. The
first column cuts between water (light key) and propylene glycol
(heavy key), and the second column adjusts the purity of the final
product to the specified value by cutting between propylene glycol
(light key) and ethylene glycol (heavy key). The columns were
simulated rigorously and its input parameters were estimated
based on typical methods.
The minimum number of stages was calculated using the Fenske
equation. The number of theoretical stages was obtained by the
Gilland-Molokanov correlation. The minimum reflux ratio was
estimated from the Underwood equation. Finally, the Kirkbride
equation was used to determine the optimum feed stage [35]. It is
noteworthy that the calculations of these estimates were per-
formed in the simulator itself, using the Shortcut Distillation block.
From the obtained values, it was possible to start the rigorous
design of the distillation columns.
In a first analysis, it was found that the operation under atmo-
spheric pressure demands extreme temperatures in condensers
and reboilers. As a consequence, common utilities such as steam
and cooling water could not be used. Also, degradation or poly-
merization reactions of residual glycerol may occur at elevated
temperatures (about 473 K) [36].
Thus, to avoid such problems, it was decided that the equipment
should operate under vacuum. The operating pressure was set as
close as possible to atmospheric pressure, but the reboilers tem-
peratures were kept at least 15 K below the glycerol degradation
temperature. The literature suggests the use of packed columns for
vacuum distillation and high viscosity mixtures, as in this case [35].
Structured packings offer better separation efficiencies and pres-
sure drops, but are more expensive. Thus, the columns were
designed with random packings of 1 in Pall metal rings, due to the
high number of theoretical stages needed [33].
For the calculation of the packed bed height, the Height Equiv-
alent to a Theoretical Plate (HETP) method was used. For vacuum
distillation, the required column diameter can be estimated from
the flooding condition. This condition should be avoided and can be
estimated from the flooding velocity through the Leva correlation
[33]. The results obtained in the simulation are in agreement with
the work of Gandarias et al. [37].
2.2. Hydrogen production: glycerol steam reforming
Reforming of oxygenated hydrocarbons consists of the cleavage
of the CeC bond to form CO and H2. Then CO and water may react in
a water-gas shift reaction, releasing more H2 and CO2 [38]. Glycerol
reforming needs a catalyst that promotes cleavage of the CeC, OeH
and CeH bonds, and if possible also promote the shift reaction
[38,39]. Reforming can be conducted in both liquid and gaseous
phases. In the latter case, it is necessary to vaporize glycerol, what
requires higher temperatures [40]. The water-gas shift reaction
occurs only in the gas phase.
Glycerol aqueous phase reforming occurs at higher pressures
and lower temperatures, if compared to steam reforming, and of-
fers greater possibilities for using crude glycerin as a raw material
[41]. The major disadvantage of this route is its low reaction rates
[42]. Vapor-phase reforming, in turn, can be conducted at atmo-
spheric pressures and has presented energetic advantages and
higher yield in terms of H2 [43]. That is why this was the route
chosen in this work.
In literature, there are studies about the activity of different
catalysts and operating conditions that maximize the glycerol
reforming and shift reactions, to the detriment of methanation and
coke formation. Such reactions decrease the efficiency of the pro-
cess in terms of H2 generation and impair the catalytic activity by
carbon deposition at the active catalyst sites [44,45].
Catalysts involving noble metals, such as Pt, Ru and Pd, and non-
noble metals, such as Ni and Co, have been successfully tested in
several studies for both aqueous- and vapor-phase reforming
[45e47]. The main disadvantage of nickel-based catalysts is coke
R.X. Jimenez et al. / Renewable Energy 158 (2020) 181e191184
formation, whereas for noble metal catalysts the high cost is an
important drawback.
Kinetic modeling of reforming of oxygenated compounds such
as ethanol and methanol has been satisfactorily done through
power-law expressions [48]. The activation energy depends on the
type of catalyst employed in the reactions. For noble metals (Pt and
Ru), the reaction rate expression depends linearly on glycerol
concentration due to the high activity of such catalysts, while for
other metals it is a function of a fractional power of the glycerol
concentration. The non-dependence of the rate on water concen-
tration is because it is used in excess [49]. Table 3 presents glycerol
reforming kinetic parameters for power-law expressions obtained
for different catalysts.
The Soave-Redlich-Kwong (SRK) equation was chosen to
calculate the thermodynamic properties in this simulation. This
choice is supported by glycerol reforming studies available in the
literature [50e54].
The glycerol steam reforming process for hydrogen production
is designed to meet the hydrogen demand required for hydro-
genolysis and is shown in Fig. 3. The stream properties are pre-
sented as Supporting Information, Table S4. Glycerol and water
enter the process at a 1:9 ratio. The reforming process involves
three reaction steps, carried out in three different reactors. Such
reactors operate under different conditions and employ different
catalysts [55]. The first reactor performs reforming itself, as
described in Eq. (4):
C3H8O3#3 CO þ 4 H2 (4)
The water-gas shift reaction occurs in the second and third re-
actors, as described in Eq. (5):
CO þ H2O#CO2 þ H2 (5)
It is recommended to perform this reaction in two steps because
it is exothermic. So, the first reactor, which is called the high-
temperature shift reactor (HTS), operates at 648 K and atmo-
spheric pressure. The second reactor, which is called the low-
temperature shift reactor (LTS), operates at 498 K and at the same
pressure, in order to shift equilibrium in the products direction
[52].
A Fe2O3 catalyst was used for the HTS reactor. This catalyst has a
low thermal deactivation rate and allows to promote the kinetics by
increasing temperaure. For the LTS reactor, a high-activity CuO
catalyst was used, which increases the products formation rate
under lower temperatures [52].
The last reactor effluent consists mainly of H2, CO2, CO traces
and water vapor, which is condensed, separated and recycled. The
surplus stream enters a Pressure Swing Adsorption (PSA) separa-
tion system, which produces hydrogen at a purity of 99.99%.
The reforming reactor and the two shift reactors are of the fixed-
bed PFR type. Thus, for design, it is first necessary to evaluate the
amount of catalyst to be employed and then calculate the reactor
volume. For the first reactor, the raw material must be heated to the
operational temperature of 973 K and the effluent must be cooled
down to 648 K, at which the HTS reactor operates, and then finally
cooled to 498 K for the LTS [55].
Due to the lack of information on all reactions in a same study
and under the same conditions, the conversion values and opera-
tional conditions adopted by Villaça [52] were used. The conversion
values obtained for methane generation and carbon formation are
much lower than that of the main by-products, so it was decided to
neglect these by-products. The carbon formed will accumulate at
the active catalyst sites over time, thereby decreasing the efficiency
of the process. This issue was considered in the economic evalua-
tion. Methane molar fraction decreases with increasing tempera-
ture, and for the temperature adopted in this work (973 K), no
methane is detected in the reactor output stream. It can also be
observed that the amount of H2 produced almost remains constant
with increasing temperature, which is in agreement with experi-
ments [41].
As the temperature increases, the molar fraction of carbon
monoxide increases. This is because the shift reaction (which
transforms CO to CO2) is exothermic and its yield is higher at low
temperatures. To model the two shift reactors (HTS and LTS), the
Equilibrium Reactor block from Aspen HYSYS® was used to calculate
the output streams’ compositions. This can be justified by the fact
that this reaction indeed tends to equilibrium in the reactor. This
stage of the simulation is different from that presented by Villaça
[52], but is in agreement with the considerations exposed earlier.
Aspen HYSYS® contains in its database the equilibrium constant
value of this reaction as a function of temperature.
Excess water is added to the process. Thus, after the LTS reactor,
water must be condensed in a flash vessel and recirculated.
Aspen HYSYS® does not allow the simulation of adsorption
processes with conventional blocks. Thus, the PSA unit is simplified
to a Splitter block, in which a process stream is separated into
several streams with defined compositions. In the present work,
purity and separation data reported in the work of Villaça [52] and a
pressure of 0.7 MPa were considered.
For sizing, two adsorption towers were considered. The columns
operate in cycles, thus ensuring that the process operates contin-
uously. The adsorption columns are filled with a type 5A alumi-
nosilicate molecular sieve, indicated to separate carbon dioxide
from hydrogen-rich streams [30]. The quantity of adsorbent can be
calculated considering an average adsorption capacity value of
0.255 kg of adsorbent per kg of adsorbate. With the bulk density,
the bed volume is obtained [30,56].
2.3. Economic analysis
The total investment cost was estimated by means of the Lang
method. The FOB price of each equipment was estimated from
correlations available in the literature [33]. These correlations were
calculated for a specific date, so updating is required. It can be done
Table 3
Kinetic parameters of power-law models for glycerol steam reforming [49].
Catalyst Temperature (K) Glycerol Order Steam Order Activation Energy

kJ
mol

Ru= Al2O3 623e773 1 e 21.2
Pt= C 623e673 1 e e
Co= Al2O3 723e823 0.1 0.4 67.2
Ni= Al2O3 723e823 0.48 0.34 60.0
Co À Ni= Al2O3 773e823 0.25 0.36 63.3
Ni= CeO2 873e923 0.233 e 103.4
Ni À ZrO2=
CeO2
973 0.3 e 43.4
R.X. Jimenez et al. / Renewable Energy 158 (2020) 181e191 185
with the Plant Cost Index, published periodically by Chemical En-
gineering Magazine. The value of the Plant Cost Index for the year
on which the correlations were based is 394. For 2018 it is 603.1
[57].
For non-corrosive environments with hydrogen, Seider et al.
[33] recommend the use of a 1% Cr and 0.5% Mo steel alloy (ASME
code SA-387B). This material was applied only in large equipment
used in operations involving the presence of hydrogen.
For the calculation of production costs, it was considered that
the plant operates continuously 24 h a day for 330 days a year.
Prices for raw materials, utilities and catalysts are summarized in
Table 4.
It was not possible to stablish an average price for hydrogen,
because several production processes employ a variety of raw
materials and energy sources. For instance, for hydrogen produced
via water electrolysis the cost may vary from 1.28 to 4.14 US$/kg,
these values being the most recent that could be found in the
literature [58]. Negro et al. [59] presented the different costs of
hydrogen in Brazil if it was produced via electrolysis, fossil fuel
reforming, or renewable fuel reforming. However, these values are
out of date and may affect the veracity of the economic analysis. For
this reason, these costs were updated using the Brazilian National
Consumer Price Index [60], that represents the accumulated infla-
tion in the country over time. The results can be observed in Table 5.
Thus, in the present work, the hydrogen price range is 0.23e4.14
US$/kg, considering the lowest price of hydrogen if it is obtained by
natural gas reforming and the highest price if it is obtained by water
electrolysis. The price of hydrogen produced by electrolysis using
solar energy was disregarded, since its value is very different from
others and the price range obtained would be too wide [59].
Rajkhowa et al. [25] quantified the dominant factors in the loss
of catalytic activity for a Cu-based catalyst during the glycerol
hydrogenolysis reaction, concluding that for this copper catalyst,
sintering occurs due to the agglomeration of Cu atoms, decreasing
the effective area of the catalytic surface. The authors evaluated the
stability of the catalyst for periods of 80e90 h under maximum
operability conditions and different raw material compositions. For
a feed of pure glycerol, the catalyst showed high stability
throughout the experiments. However, for a feed stream containing
pollutants such as sulfur, chlorine, and glycerides from biodiesel
production, its activity decreases. This is one more reason to
consider bidistilled glycerol as the raw material. Thus, the loss of
catalytic activity was neglected.
In the hydrogen production process, Levalley et al. [61] indicate
that copper and iron catalysts for the shift reaction keep a good
performance from 2 to 4 years. An average time of 3 years was
assumed for the economic analysis. Berman and Epstein [62]
proved that RueAl2O3 catalysts suffer significant thermal decom-
position above 1373 K. As the reforming reactor operates at 973 K, it
was considered that thermal deactivation is not critical, and the
catalyst may be replaced together with the other catalysts.
The annual consumption of raw materials and utilities was ob-
tained directly from Aspen HYSYS®. For the calculation of
manpower, the Brazilian average wage for industrial operators was
considered [63]. The number of workers per shift was estimated
from the number of equipment units, considering five shifts as
recommended by Seider et al. [33]. The remaining production costs
were estimated from correlations available in the literature.
3. Results and discussion
A brief description of the process equipment size and cost can be
found in Table 6, while the relative costs of equipment pieces are
shown in Fig. 4. To produce 821 kg/h of propylene glycol, 22 kg/h of
hydrogen, 1070 kg/h of glycerol, and 267 kg/h of water are needed.
Producing 23 kg/h of hydrogen requires 174 kg/h of glycerol and
127 kg/h of water. It is clear that the distillation columns are the
most expensive equipment in the propylene glycol process. This is
due to the high purity specification of PG, which leads to larger
distillation columns. Due to the high steam demand, the steam
generation unit costs represent more than half of the equipment
costs for the steam reforming process, as can be seen in Fig. 5.
Figs. 6 and 7 show the costs distribution of the utilities
consumed in each process. Cooling water and steam are the main
utilities used in the PG process because of the distillation columns’
Fig. 3. Glycerol steam reforming flowsheet.
Table 4
Prices of raw materials, utilities and products involved in the glycerol hydro-
genolysis and glycerol steam reforming processes.
Raw Material Unit Price (US$/kg) Reference
Glycerol 0.4918 [64]
Hydrogen 1.28e4.14 [58]
Process Water 0.0008 [65]
Utilities
450 psi Vapor 0.0130 [66] a
Cooling Water 0.0008 [65]
Fuel (US$/m3
) 0.9768 [67]
Electricity (US$/kWh) 0.1032 [68]
Effluent Treatment 0.0560 [69]
Catalysts
Cu= ZnO= Al2O3 1.8000 [70]
Ru= Al2O3 21000 [70]
Fe2O3 13.900 [70]
CuO 6.7000 [70]
Products
Propylene Glycol 1.4500 [10]
a
Updated to 2018 with the Plant Cost Index.
R.X. Jimenez et al. / Renewable Energy 158 (2020) 181e191186
reboilers and condensers. Moreover, there is also a high electricity
demand due to the compressor.
Considering the production of partially renewable propylene
glycol (with hydrogen purchased from an external source) with an
average price of 2.185 US$/kg for H2, the break-even price for pro-
pylene glycol is 1.17 US$/kg, with a total investment of 4,722,678.07
US$. The break-even price of hydrogen production via glycerol
steam reforming is 9.01 US$/kg, with a total investment of
1,441,653.78 US$. The detailed cost statement of the processes is
presented as Supporting Information, Table S1.
It was found that the break-even price of propylene glycol is 1.36
US$/kg for a fully renewable PG production process (that is, a
propylene glycol production process integrated with a glycerol
reforming process to produce hydrogen). Therefore, the generation
of a 100% green PG resulted in a 16% increase in its final price.
Table 5
Prices of hydrogen from different sources in 2003 and 2018 [59].
Process Energy/Raw Material Source 2003 Price (US$/kg) 2018 Price (US$/kg)
Water
Electrolysis
Nuclear Energy 1.98 3.62
Hydraulic Energy 0.79 1.45
Natural Gas 1.58 2.90
Solar 20.76 38.10
Fossil Fuels Reforming Natural Gas 0.12 0.23
Gasoline 1.35 2.48
Methanol 1.19 2.19
Renewable Fuels Reforming Biogas 1.71 3.15
Ethanol 1.50 2.74
Table 6
Propylene glycol and glycerol reforming processes equipment sizing and cost estimates.
Equipment Equipment Size Size Variable Cost (US$)
Propylene Glycol Process
Heat Exchanger 58.81 Total Heat Transfer Area (m2
) 59,733.88
Reactor 1.20 Volume (m3
) 26,967.90
Distillation Colum 1 18 Theoretical Stages 16,450.77
Distillation Colum 2 60 Theoretical Stages 137,091.41
Flash Vessel 1.60 Height (m) 405,172.96
Glycerol Steam Reforming Process
Vapor Generation Unit e e 91,355.20
Heat Exchanger 1 2.86 Total Heat Transfer Area (m2
) 4403.80
Heat Exchanger 1 2.86 Total Heat Transfer Area (m2
) 4140.31
Reforming Reactor 0.0119 Volume (m3
) 7908.29
HTS Reactor 0.0055 Volume (m3
) 10,604.89
LTS Reactor 0.0052 Volume (m3
) 9592.25
Flash Vessel 1.60 Volume (m3
) 7908.29
PSA 0.1435 Volume (m3
) 13,404.60
Fig. 4. Propylene glycol process equipment cost distribution.
Fig. 5. Steam reforming process equipment cost distribution.
R.X. Jimenez et al. / Renewable Energy 158 (2020) 181e191 187
However, this additional expense can be offset when the expanding
market shares of more sustainable products are taken into account.
The costs distributions of the evaluated processes are shown in
Table 7. It can be seen that raw material and catalysts correspond to
55e60% of the total production cost. The renewable PG production
cost is higher in nearly US$ 1.2 million, due to the price of renew-
able hydrogen.
The main source of uncertainty in the economic analysis for the
propylene glycol production process considering that the hydrogen
is purchased from an external source is its purchase price, which
depends on the raw material used to produce hydrogen, the
quantity purchased and the distance between the PG plant and
hydrogen plant.
For this reason, a sensitivity analysis was performed to assess
the impact of the price of H2 on the PG production cost. It was found
that a 1% increase in the price of hydrogen implies a 0.045% increase
in the PG production cost. It follows that the price of propylene
glycol is not greatly influenced by small variations in the price of
hydrogen. Considering the limits of the hydrogen price range [59], a
price range for propylene glycol of 1.15e1.23 US$/kg is obtained.
Thus, it can be said that even for large variations in the price of
hydrogen, the propylene glycol production cost is not significantly
affected.
If a similar sensitivity analysis concerning the glycerol price is
performed, it is concluded that for a 1% increase in the price of
glycerol the glycerol reforming process cost increases by 0.41% and
the propylene glycol production process cost increases by 0.57%.
The selling price of PG and the quantity sold is subject to
changes due to market offer and demand. Fig. 8 shows the variation
of annual net profit with the variation of the PG price. This analysis
was performed for a constant propylene glycol production of
824 kg/h, which was the amount obtained by processing all the
glycerol generated by an average biodiesel plant in Brazil.
A similar analysis is shown in Fig. 9, but in this case the amount
of PG produced is varied and its price is kept constant. For the
renewable propylene glycol, the minimum quantity to be produced
to achieve a zero annual net profit is approximately 776 kg/h. For
the partially renewable propylene glycol with hydrogen obtained
Fig. 6. Propylene glycol process utilities cost distribution.
Fig. 7. Steam reforming process utilities cost distribution.
Table 7
Production costs distribution for the partially renewable PG and renewable PG
processes.
Production Costs Distribution Partially Renewable PG Renewable PG
Raw Materials and Catalyst 59.7% 55.7%
Direct Costs 7.0% 10.7%
Indirect Costs 4.7% 6.8%
Utilities 18.7% 16.9%
Devaluation 5.0% 5.6%
Other Expenses 4.9% 4.3%
Fig. 8. Variation in the net annual profit as a function of PG selling price.
R.X. Jimenez et al. / Renewable Energy 158 (2020) 181e191188
from an external source (and a lower production cost), the mini-
mum quantity is 670 kg/h. So, it takes 106 kg/h more PG to make
green propylene glycol feasible, which represents a relatively
broader market share.
A comparison between the conventional process, which uses
propylene oxide as the main raw material, and the renewable and
partially renewable processes was carried out. Raw material and
utility consumption for the conventional process were estimated by
simplified mass and energy balances in Aspen HYSYS®, as shown in
Fig. 1. The stream properties are presented as Supporting Infor-
mation, Table S5. In this technology, the reaction consists of pro-
pylene oxide hydration in the presence of an acid catalyst (sulfuric
acid), and methanol is used as a solvent. The reaction is normally
carried out in a continuous stirred tank [71].
There are some clear disadvantages in the conventional process,
such as petroleum-based raw materials, toxic solvents and corro-
sion problems. Propylene oxide is commonly obtained from pro-
pylene (petroleum based), while methanol is produced from
natural gas. Corrosion problems must be taken in consideration,
because of the use of H2SO4, which is dissolved in water in 0,1% wt.
To produce the same amount of PG, 4000 kg/h of water, 40 kg/h of
H2SO4, 643 kg/h of methanol and 675 kg/h of propylene oxide are
needed. In comparison, the partially green process uses only 267 kg/
h of water and the fully green process consumes a total of 394 kg/h
of process water. Both processes require none of the other com-
ponents. Aspen HYSYS® also gives an estimation of the carbon
emissions. Its 2237 kg/h for the conventional process, against
467 kg/h for the propylene glycol production stage and 141 kg/h for
steam reforming.
Fig. 10 (A) shows that the conventional process’s raw material
cost is greater in nearly 25% than the partially renewable PG process
and 16% than the renewable PG process. This difference is caused by
the high demand of solvent, which is used in a propylene oxide-
methanol equivolume mixture. In Fig. 10 (B) the utilities demand
of each process can be compared. The propylene oxide process
utilities consumption is almost 3.5 times larger than in the
renewable PG process. This high consumption is due to the
requirement to keep the reactor temperature under 325 K and the
separation steps needed to achieve the same purity that has been
reached in the other processes. The reactor temperature is a critical
variable, because of propylene oxide’s low boiling point. Addi-
tionally, it is important to notice that the renewable PG process
consumes more utilities than the partially renewable PG process,
due to steam reforming.
Finally, these results suggest that both green and partially green
PG are economically feasible, because both break-even prices of
propylene glycol were lower than the reported market price, even
though its price depends on the existing demand. The environ-
mental advantages of these products seem clear, but, for future
work, it would be interesting to perform a formal environmental
analysis of the green propylene glycol production process, such as a
Life Cycle Assessment (LCA), and to compare it to the conventional
PG production process to have more elucidative results in this
sense.
4. Conclusion
The present work simulated the propylene glycol production
Fig. 9. Variation in the net annual profit as a function of PG production.
Fig. 10. Comparison of the PG production processes. (A) Raw material costs. (B) Util-
ities demands.
R.X. Jimenez et al. / Renewable Energy 158 (2020) 181e191 189
process using glycerol from biodiesel production as the main raw
material, in a Brazilian scenario. A total production cost of 1.17 US$/
kg was obtained when PG is produced using hydrogen from an
external source. When it is produced using hydrogen obtained from
glycerol steam reforming, its price increases to 1.36 US$/kg. A
hydrogen production cost of 9.01 US$/kg was calculated when it is
obtained from glycerol steam reforming, which is much higher
when compared to the market prices of hydrogen obtained by other
processes. It has caused a nearly 70% drop in the annual net profit of
propylene glycol production in the base scenario, although the
process remains economically viable. Fully renewable PG is
approximately 16% more expensive than partially renewable PG,
but it is possible, depending on the market.
Declaration of competing interest
The authors declare that they have no known competing
financial interests or personal relationships that could have
appeared to influence the work reported in this paper.
CRediT authorship contribution statement
Roberto X. Jimenez: Conceptualization, Investigation, Software,
Writing - original draft, Writing - review  editing. Andre F. Young:
Conceptualization, Methodology, Supervision, Visualization,
Writing - original draft, Writing - review  editing. Heloisa L.S.
Fernandes: Conceptualization, Supervision, Writing - original draft,
Writing - review  editing.
Appendix A. Supplementary data
Supplementary data to this article can be found online at
https://doi.org/10.1016/j.renene.2020.05.126.
References
[1] A.F. Young, F.L.P. Pessoa, E.M. Queiroz, Biodiesel Production Technologies -
Supercritical and Enzymatic Production, Novas Ediç~oes Acad^emicas, Saar-
brücken, Germany, 2015 (in Portuguese).
[2] M.C.S. de Mello, H.G.D. Villardi, A.F. Young, F.L.P. Pessoa, A.M. Salgado, Life
cycle assessment of biodiesel produced by the methylic-alkaline and ethylic-
enzymatic routes, Fuel 208 (2017) 329e336, https://doi.org/10.1016/
j.fuel.2017.07.014.
[3] M.R. Monteiro, C.L. Kugelmeier, R.S. Pinheiro, M.O. Batalha, A. da Silva Cesar,
Glycerol from biodiesel production: technological paths for sustainability,
Renew. Sustain. Energy Rev. 88 (2018) 109e122, https://doi.org/10.1016/
j.rser.2018.02.019.
[4] ANP - Brazilian National Agency of Petroleum, Natural Gas and Biofuels.
Biodisel, 2018 accessed 10.1.19, http://www.anp.gov.br/producao-de-
biocombustiveis/biodiesel/simp-biodisel (in Portuguese).
[5] C.J.A. Mota, Pinto, Catalytic transformations of glycerol for innovation in the
chemical industry, Rev. Virtual Quim. 9 (2017) 135e149, https://doi.org/
10.21577/1984-6835.20170011.
[6] Z. Freitas, Glycerin as raw material for chemical industry: Glicerina como
materia-prima para indústria química: evaluation of research efforts and
commercial initiatives, Federal University of Rio de Janeiro, Rio de Janeiro,
Brazil, 2013 (in Portuguese).
[7] S. Veluturla, N. Archna, D. Subba Rao, N. Hezil, I.S. Indraja, S. Spoorthi, Catalytic
valorization of raw glycerol derived from biodiesel: a review, Biofuels 9 (2018)
305e314, https://doi.org/10.1080/17597269.2016.1266234.
[8] H. Mitta, P.K. Seelam, S. Ojala, R.L. Keiski, P. Balla, Tuning Y-zeolite based
catalyst with copper for enhanced activity and selectivity in vapor phase
hydrogenolysis of glycerol to 1,2-propanediol, Appl. Catal. Gen. 550 (2018)
308e319, https://doi.org/10.1016/j.apcata.2017.10.019.
[9] Y. Nakagawa, K. Tomishige, Heterogeneous catalysis of the glycerol hydro-
genolysis, Catal. Sci. Technol. 1 (2011) 179, https://doi.org/10.1039/
c0cy00054j.
[10] Abiquim - Brazilian Chemical Industry Association, Annual Report, 2018 (in
Portuguse).
[11] Y. Zhang, M.A. Dube, D.D. McLean, M. Kates, Biodiesel production from waste
cooking oil: 1. Process design and technological assessment, Bioresour.
Technol. 89 (2003) 1e16, https://doi.org/10.1016/S0960-8524(03)00040-3.
[12] Z. Zhou, X. Li, T. Zeng, W. Hong, Z. Cheng, W. Yuan, Kinetics of hydrogenolysis
of glycerol to propylene glycol over Cu-ZnO-Al2O3 catalysts, Chin. J. Chem.
Eng. 18 (2010) 384e390, https://doi.org/10.1016/S1004-9541(10)60235-2.
[13] M.A. Dasari, P.-P. Kiatsimkul, W.R. Sutterlin, G.J. Suppes, Low-pressure
hydrogenolysis of glycerol to propylene glycol, Appl. Catal. Gen. 281 (2005)
225e231, https://doi.org/10.1016/J.APCATA.2004.11.033.
[14] M.N. Gatti, M.D. Mizrahi, J.M. Ramallo-Lopez, F. Pompeo, G.F. Santori,
N.N. Nichio, Improvement of the catalytic activity of Ni/SiO2-C by the modi-
fication of the support and Zn addition: bio-propylene glycol from glycerol,
Appl. Catal. Gen. 548 (2017) 24e32, https://doi.org/10.1016/
J.APCATA.2017.08.037.
[15] E.P. Maris, R.J. Davis, Hydrogenolysis of glycerol over carbon-supported Ru
and Pt catalysts, J. Catal. 249 (2007) 328e337, https://doi.org/10.1016/
J.JCAT.2007.05.008.
[16] T. Miyazawa, Y. Kusunoki, K. Kunimori, K. Tomishige, Glycerol conversion in
the aqueous solution under hydrogen over Ru/C þ an ion-exchange resin and
its reaction mechanism, J. Catal. 240 (2006) 213e221, https://doi.org/10.1016/
J.JCAT.2006.03.023.
[17] C. Montassier, J.C. Menezo, L.C. Hoang, C. Renaud, J. Barbier, Aqueous polyol
conversions on ruthenium and on sulfur-modified ruthenium, J. Mol. Catal. 70
(1991) 99e110, https://doi.org/10.1016/0304-5102(91)85008-P.
[18] S.R. Schmidt, S.K. Tanielyan, N. Marin, G. Alvez, R.L. Augustine, Selective
conversion of glycerol to propylene glycol over fixed bed Raney® Cu catalysts,
Top. Catal. 53 (2010) 1214e1216, https://doi.org/10.1007/s11244-010-9565-
x.
[19] R.V. Sharma, P. Kumar, A.K. Dalai, Selective hydrogenolysis of glycerol to
propylene glycol by using Cu:Zn:Cr:Zr mixed metal oxides catalyst, Appl.
Catal. Gen. 477 (2014) 147e156, https://doi.org/10.1016/
J.APCATA.2014.03.007.
[20] E.S. Vasiliadou, A.A. Lemonidou, Glycerol transformation to value added C3
diols: reaction mechanism, kinetic, and engineering aspects, Wiley Inter-
discip. Rev. Energy Environ. 4 (2015) 486e520, https://doi.org/10.1002/
wene.159.
[21] G.J. Suppes, W.R. Sutterlin, M. Dasari, Method of Producing Lower Alcohols
from Glycerol, 2007. US7943805B2.
[22] C. Montassier, D. Giraud, J. Barbier, Polyol conversion by liquid phase het-
erogeneous catalysis over metals, Stud. Surf. Sci. Catal. 41 (1988) 165e170,
https://doi.org/10.1016/S0167-2991(09)60811-9.
[23] S. Wang, Y. Zhang, H. Liu, Selective hydrogenolysis of glycerol to propylene
glycol on Cu-ZnO composite catalysts: structural requirements and reaction
mechanism, Chem. Asian J. 5 (2010) 1100e1111, https://doi.org/10.1002/
asia.200900668.
[24] D. Lahr, B. Shanks, Kinetic analysis of the hydrogenolysis of lower polyhydric
Alcohols: glycerol to glycols, Ind. Eng. Chem. Res. 42 (2003) 5467e5472,
https://doi.org/10.1021/IE030468L.
[25] T. Rajkhowa, G.B. Marin, J.W. Thybaut, A comprehensive kinetic model for Cu
catalyzed liquid phase glycerol hydrogenolysis, Appl. Catal. B Environ. 205
(2017) 469e480, https://doi.org/10.1016/J.APCATB.2016.12.042.
[26] C.K. Cheng, S.Y. Foo, A.A. Adesina, Glycerol steam reforming over bimetallic
Co-Ni/Al2O3, Ind. Eng. Chem. Res. 49 (2010) 10804e10817, https://doi.org/
10.1021/ie100462t.
[27] A. Lancia, D. Musmarra, F. Pepe, Vapor-liquid equilibria for mixtures of
ethylene glycol, propylene glycol, and water between 98º and 122ºC, J. Chem.
Eng. Jpn. 29 (1996) 449e455, https://doi.org/10.1252/jcej.29.449.
[28] E.S. Vasiliadou, A.A. Lemonidou, Kinetic study of liquid-phase glycerol
hydrogenolysis over Cu/SiO2 catalyst, Chem. Eng. J. 231 (2013) 103e112,
https://doi.org/10.1016/j.cej.2013.06.096.
[29] Y. Xi, J.E. Holladay, J.G. Frye, A.A. Oberg, J.E. Jackson, D.J. Miller, A kinetic and
mass transfer model for glycerol hydrogenolysis in a trickle-bed reactor, Org.
Process Res. Dev. 14 (2010) 1304e1312, https://doi.org/10.1021/op900336a.
[30] R. Perry, D. Green, Perry’s Chemical Engineers’ Handbook, McGraw-Hil, New
York, United States, 1997.
[31] C.N. Satterfield, Trickle-bed reactors, AIChE J. 21 (1975) 209e228, https://
doi.org/10.1002/aic.690210202.
[32] R.J. Farrauto, C.H. Bartholomew, Fundamentals of Industrial Catalytic Pro-
cesses, Blackie Academic  Professional, London, United Kingdom, 1997.
[33] W.D. Seider, J.D. Seader, D.R. Lewin, Product and Process Design Principles -
Synthesis, Analysis, and Evaluation, Second, John Wiley and Sons, Inc, Inter-
national Edition, 2003.
[34] D. Woods, Rules of Thumb in Engineering Practice, Wiley, Darmstadt, Ger-
many, 2007.
[35] R. Sinnott, G. Towler, Chemical Engineering Design, Elsevier Inc., United
States, 2008.
[36] C. Ferreira, Process Flowsheet Proposal for Purification of Glycerol from Bio-
diesel Production, Federal University of Rio de Janeiro, Rio de Janeiro, Brazil,
2018 (in Portuguese).
[37] I. Gandarias, P.L. Arias, I. Agirrezabal-Telleria, Economic assessment for the
production of 1,2-Propanediol from bioglycerol hydrogenolysis using molec-
ular hydrogen or hydrogen donor molecules, Environ. Prog. Sustain. Energy 35
(2016) 447e454, https://doi.org/10.1002/ep.12232.
[38] R.R. Davda, J.W. Shabaker, G.W. Huber, R.D. Cortright, J.A. Dumesic, A review
of catalytic issues and process conditions for renewable hydrogen and alkanes
by aqueous-phase reforming of oxygenated hydrocarbons over supported
metal catalysts, Appl. Catal. B Environ. 56 (2005) 171e186, https://doi.org/
10.1016/J.APCATB.2004.04.027.
[39] A. Iriondo, V.L. Barrio, J.F. Cambra, P.L. Arias, M.B. Güemez, R.M. Navarro,
M.C. Sanchez-Sanchez, J.L.G. Fierro, Hydrogen production from glycerol over
R.X. Jimenez et al. / Renewable Energy 158 (2020) 181e191190
nickel catalysts supported on Al2O3 modified by Mg, Zr, Ce or La, Top. Catal. 49
(2008) 46e58, https://doi.org/10.1007/s11244-008-9060-9.
[40] P.D. Vaidya, A.E. Rodrigues, Glycerol reforming for hydrogen production: a
review, Chem. Eng. Technol. 32 (2009) 1463e1469, https://doi.org/10.1002/
ceat.200900120.
[41] R.L. Manfro, N.F.P. Ribeiro, M.M.V.M. Souza, Production of hydrogen from
steam reforming of glycerol using nickel catalysts supported on Al2O3, CeO2
and ZrO2, Catal. Sustain. Energy. 1 (2013) 60e70, https://doi.org/10.2478/cse-
2013-0001.
[42] I. Iliuta, H.R. Radfarnia, M.C. Iliuta, Hydrogen production by sorption-
enhanced steam glycerol reforming: sorption kinetics and reactor simula-
tion, AIChE J. 59 (2013) 2105e2118, https://doi.org/10.1002/aic.13979.
[43] R.L. Manfro, M.M.V.M. Souza, Catalyst for hydrogen production through
glycerol reforming review, Curr. Top. Catal. 11 (2014) 37e54.
[44] S. Adhikari, S. Fernando, S.R. Gwaltney, S.D. Filip To, R. Mark Bricka,
P.H. Steele, A. Haryanto, A thermodynamic analysis of hydrogen production by
steam reforming of glycerol, Int. J. Hydrogen Energy 32 (2007) 2875e2880,
https://doi.org/10.1016/J.IJHYDENE.2007.03.023.
[45] B. Zhang, X. Tang, Y. Li, Y. Xu, W. Shen, Hydrogen production from steam
reforming of ethanol and glycerol over ceria-supported metal catalysts, Int. J.
Hydrogen Energy 32 (2007) 2367e2373, https://doi.org/10.1016/
J.IJHYDENE.2006.11.003.
[46] A.O. Menezes, M.T. Rodrigues, A. Zimmaro, L.E.P. Borges, M.A. Fraga, Produc-
tion of renewable hydrogen from aqueous-phase reforming of glycerol over Pt
catalysts supported on different oxides, Renew. Energy 36 (2011) 595e599,
https://doi.org/10.1016/J.RENENE.2010.08.004.
[47] K.S. Avasthi, R.N. Reddy, S. Patel, Challenges in the production of hydrogen
from glycerol e a biodiesel byproduct via steam reforming process, Procedia
Eng. 51 (2013) 423e429, https://doi.org/10.1016/J.PROENG.2013.01.059.
[48] S. Adhikari, S.D. Fernando, A. Haryanto, Kinetics and reactor modeling of
hydrogen production from glycerol via steam reforming process over Ni/CeO2
catalysts, Chem. Eng. Technol. 32 (2009) 541e547, https://doi.org/10.1002/
ceat.200800462.
[49] R. Sundari, P.D. Vaidya, Reaction kinetics of glycerol steam reforming using a
Ru/Al2O3 catalyst, Energy Fuels 26 (2012) 4195e4204, https://doi.org/
10.1021/ef300658n.
[50] H. Chen, Y. Ding, N.T. Cong, B. Dou, V. Dupont, M. Ghadiri, P.T. Williams,
A comparative study on hydrogen production from steam-glycerol reforming:
thermodynamics and experimental, Renew. Energy 36 (2011) 779e788,
https://doi.org/10.1016/j.renene.2010.07.026.
[51] Y. Liu, R. Farruto, A. Lawal, Autothermal reforming of glycerol in a dual layer
monolith catalyst, Chem. Eng. Sci. 89 (2013) 31e39, https://doi.org/10.1016/
J.CES.2012.11.030.
[52] J. Villaça, Proposal and Simulation of the Glycerol Steam Reforming Process for
Hydrogen Obtaining, Federal University of Rio de Janeiro, Rio de Janeiro,
Brazil, 2018 (in Portuguese).
[53] X. Wang, S. Li, H. Wang, B. Liu, X. Ma, Thermodynamic analysis of glycerin
steam reforming, Energy Fuels 22 (2008) 4285e4291, https://doi.org/10.1021/
ef800487r.
[54] G. Yang, H. Yu, F. Peng, H. Wang, J. Yang, D. Xie, Thermodynamic analysis of
hydrogen generation via oxidative steam reforming of glycerol, Renew. En-
ergy 36 (2011) 2120e2127, https://doi.org/10.1016/J.RENENE.2011.01.022.
[55] M.M.V.M. Souza, Hydrogen Technology, Synergia, Rio de Janeiro, Brazil, 2009
(in Portuguese).
[56] A. Hines, R. Maddox, Mass Transfer Fundamentals and Applications, Prentice-
Hall, Michigan, United States, 1985.
[57] S. Jenkins, Economic Indicators, Chem. Eng, 2019. https://www.
chemengonline.com/cepci-updates-january-2018-prelim-and-december-
2017-final/?printmode¼1. accessed 10.3.18.
[58] L. Bonfim-Rocha, M.L. Gimenes, S.H. Bernardo de Faria, R.O. Silva, L.J. Esteller,
Multi-objective design of a new sustainable scenario for bio-methanol pro-
duction in Brazil, J. Clean. Prod. 187 (2018) 1043e1056, https://doi.org/
10.1016/J.JCLEPRO.2018.03.267.
[59] M.L.M. Negro, R.C. Garner, M. Linardi, Mass production of hydrogen in Brazil,
world clim, Energy Event 3 (2003) 293e298.
[60] IBGE - Brazilian Institute of Geography and Statistics, National Consumer Price
Index, 2019. https://www.ibge.gov.br/estatisticas/economicas/precos-e-
custos/9256-indice-nacional-de-precos-ao-consumidor-amplo.html?
=t=series-historicas. accessed 9.29.18.
[61] T.L. LeValley, A.R. Richard, M. Fan, The progress in water gas shift and steam
reforming hydrogen production technologies e a review, Int. J. Hydrogen
Energy 39 (2014) 16983e17000, https://doi.org/10.1016/
j.ijhydene.2014.08.041.
[62] A. Berman, M. Epstein, Ruthenium catalysts for high temperature solar
reforming of methane, in: Hydrogen Power: Theoretical and Engineering
Solutions, Springer, Dordrecht, Netherlands, 1998, pp. 213e218, https://
doi.org/10.1007/978-94-015-9054-9_26.
[63] Glassdoor, Industrial Operator Salaries, 2018 accessed 9.20.18, https://www.
lovemondays.com.br/salarios/cargo/salario-operador-industrial (in
Portuguese).
[64] Comex Stat, 2018. http://comexstat.mdic.gov.br/pt/home. accessed 9.20.18.
[65] SNIS - National System of Information about Sanitation, Diagnosis of the
Water and Sewage Services - 2017, Brasilia, Brazil, 2019.
[66] C.A.G. Perlingeiro, Process Systems Engineering, Blucher, S~ao Paulo, Brazil,
2005 (in Portuguese).
[67] Comgas - S~ao Paulo Gas Company, Pipe Natural Gas Tariffs, 2018 accessed
10.1.18, https://www.comgas.com.br/tarifas/comercial/ (in Portuguse).
[68] ANEEL - National Electricity Regulatory Agency, Energy consumption report
by sector, accessed 9.20.18, http://relatorios.aneel.gov.br/_layouts/xlviewer.
aspx?id=/RelatoriosSAS/RelSampRegCC.xlsxSource=http://relatorios.aneel.
gov.br/RelatoriosSAS/Forms/AllItems.aspxDefaultItemOpen=1, 2018 (in
Portuguese).
[69] R. Turton, R. Bailie, W. Whiting, J. Shaeiwitz, Analysis, Synthesis and Design of
Chemical Processes, Prentice Hall, Boston, United States, 2008.
[70] Sigma-Aldrich, Product Directory, 2018. https://www.sigmaaldrich.com/
technical-service-home/product-catalog.html. accessed 10.1.18.
[71] H.S. Fogler, Elements of Chemical Engineering, fourth ed., Prentice Hall, New
Jersey, United States, 2009.
R.X. Jimenez et al. / Renewable Energy 158 (2020) 181e191 191

Weitere ähnliche Inhalte

Was ist angesagt?

Synthesis of Oxygenated Fuel Additives via Acetylation of Bio-Glycerol over H...
Synthesis of Oxygenated Fuel Additives via Acetylation of Bio-Glycerol over H...Synthesis of Oxygenated Fuel Additives via Acetylation of Bio-Glycerol over H...
Synthesis of Oxygenated Fuel Additives via Acetylation of Bio-Glycerol over H...crimsonpublisherspps
 
Synthesis of Oxygenated Fuel Additives via Acetylation of Bio-Glycerol over H...
Synthesis of Oxygenated Fuel Additives via Acetylation of Bio-Glycerol over H...Synthesis of Oxygenated Fuel Additives via Acetylation of Bio-Glycerol over H...
Synthesis of Oxygenated Fuel Additives via Acetylation of Bio-Glycerol over H...DanesBlake
 
Phase equilibrium feasibility studies of free fatty acids extraction from pal...
Phase equilibrium feasibility studies of free fatty acids extraction from pal...Phase equilibrium feasibility studies of free fatty acids extraction from pal...
Phase equilibrium feasibility studies of free fatty acids extraction from pal...Alexander Decker
 
Presentation on a novel accelerated composting process.
Presentation on a novel accelerated composting process.Presentation on a novel accelerated composting process.
Presentation on a novel accelerated composting process.Kiran Parmar
 
132952928 4th-yr-dsgn-pro
132952928 4th-yr-dsgn-pro132952928 4th-yr-dsgn-pro
132952928 4th-yr-dsgn-proBatuhanKse1
 
Application of Hydrotrophy in HPLC Hydrotrophic Solution, A Novel Cost Effect...
Application of Hydrotrophy in HPLC Hydrotrophic Solution, A Novel Cost Effect...Application of Hydrotrophy in HPLC Hydrotrophic Solution, A Novel Cost Effect...
Application of Hydrotrophy in HPLC Hydrotrophic Solution, A Novel Cost Effect...ijtsrd
 
IRJET- Hydrothermal Liquefaction Process (HTL) of Sugarcane Bagasse for the P...
IRJET- Hydrothermal Liquefaction Process (HTL) of Sugarcane Bagasse for the P...IRJET- Hydrothermal Liquefaction Process (HTL) of Sugarcane Bagasse for the P...
IRJET- Hydrothermal Liquefaction Process (HTL) of Sugarcane Bagasse for the P...IRJET Journal
 
1 s2.0-s0960852414013169-main
1 s2.0-s0960852414013169-main1 s2.0-s0960852414013169-main
1 s2.0-s0960852414013169-mainEly Maestria
 
IRJET- Study of the Products Obtained by Treating Glutamic Acid with Tertiary...
IRJET- Study of the Products Obtained by Treating Glutamic Acid with Tertiary...IRJET- Study of the Products Obtained by Treating Glutamic Acid with Tertiary...
IRJET- Study of the Products Obtained by Treating Glutamic Acid with Tertiary...IRJET Journal
 
334529321 production-of-acetic-acid-pptx
334529321 production-of-acetic-acid-pptx334529321 production-of-acetic-acid-pptx
334529321 production-of-acetic-acid-pptxBatuhanKse1
 
Recycling Technology - A Cost Effective Approach for the Synthesis of Alterna...
Recycling Technology - A Cost Effective Approach for the Synthesis of Alterna...Recycling Technology - A Cost Effective Approach for the Synthesis of Alterna...
Recycling Technology - A Cost Effective Approach for the Synthesis of Alterna...Associate Professor in VSB Coimbatore
 
PLANT DESIGN FOR MANUFACTURING OF HYDROGEN BY STEAM METHANE REFORMING (SMR)
PLANT DESIGN FOR MANUFACTURING OF HYDROGEN BY STEAM METHANE REFORMING (SMR)PLANT DESIGN FOR MANUFACTURING OF HYDROGEN BY STEAM METHANE REFORMING (SMR)
PLANT DESIGN FOR MANUFACTURING OF HYDROGEN BY STEAM METHANE REFORMING (SMR)Priyam Jyoti Borah
 
Biomass to olefins cracking of renewable naphtha
Biomass to olefins    cracking of renewable naphthaBiomass to olefins    cracking of renewable naphtha
Biomass to olefins cracking of renewable naphthapxguru
 
SPE IPTC Poster Gas Hydrate
SPE IPTC Poster Gas HydrateSPE IPTC Poster Gas Hydrate
SPE IPTC Poster Gas HydrateArdian Nengkoda
 

Was ist angesagt? (20)

Synthesis of Oxygenated Fuel Additives via Acetylation of Bio-Glycerol over H...
Synthesis of Oxygenated Fuel Additives via Acetylation of Bio-Glycerol over H...Synthesis of Oxygenated Fuel Additives via Acetylation of Bio-Glycerol over H...
Synthesis of Oxygenated Fuel Additives via Acetylation of Bio-Glycerol over H...
 
Synthesis of Oxygenated Fuel Additives via Acetylation of Bio-Glycerol over H...
Synthesis of Oxygenated Fuel Additives via Acetylation of Bio-Glycerol over H...Synthesis of Oxygenated Fuel Additives via Acetylation of Bio-Glycerol over H...
Synthesis of Oxygenated Fuel Additives via Acetylation of Bio-Glycerol over H...
 
Acetic Acid Process Plant Design
Acetic Acid Process Plant DesignAcetic Acid Process Plant Design
Acetic Acid Process Plant Design
 
Phase equilibrium feasibility studies of free fatty acids extraction from pal...
Phase equilibrium feasibility studies of free fatty acids extraction from pal...Phase equilibrium feasibility studies of free fatty acids extraction from pal...
Phase equilibrium feasibility studies of free fatty acids extraction from pal...
 
tA04 04 0106
tA04 04 0106tA04 04 0106
tA04 04 0106
 
Presentation on a novel accelerated composting process.
Presentation on a novel accelerated composting process.Presentation on a novel accelerated composting process.
Presentation on a novel accelerated composting process.
 
132952928 4th-yr-dsgn-pro
132952928 4th-yr-dsgn-pro132952928 4th-yr-dsgn-pro
132952928 4th-yr-dsgn-pro
 
Application of Hydrotrophy in HPLC Hydrotrophic Solution, A Novel Cost Effect...
Application of Hydrotrophy in HPLC Hydrotrophic Solution, A Novel Cost Effect...Application of Hydrotrophy in HPLC Hydrotrophic Solution, A Novel Cost Effect...
Application of Hydrotrophy in HPLC Hydrotrophic Solution, A Novel Cost Effect...
 
IRJET- Hydrothermal Liquefaction Process (HTL) of Sugarcane Bagasse for the P...
IRJET- Hydrothermal Liquefaction Process (HTL) of Sugarcane Bagasse for the P...IRJET- Hydrothermal Liquefaction Process (HTL) of Sugarcane Bagasse for the P...
IRJET- Hydrothermal Liquefaction Process (HTL) of Sugarcane Bagasse for the P...
 
T33108112
T33108112T33108112
T33108112
 
1 s2.0-s0960852414013169-main
1 s2.0-s0960852414013169-main1 s2.0-s0960852414013169-main
1 s2.0-s0960852414013169-main
 
IRJET- Study of the Products Obtained by Treating Glutamic Acid with Tertiary...
IRJET- Study of the Products Obtained by Treating Glutamic Acid with Tertiary...IRJET- Study of the Products Obtained by Treating Glutamic Acid with Tertiary...
IRJET- Study of the Products Obtained by Treating Glutamic Acid with Tertiary...
 
334529321 production-of-acetic-acid-pptx
334529321 production-of-acetic-acid-pptx334529321 production-of-acetic-acid-pptx
334529321 production-of-acetic-acid-pptx
 
Recycling Technology - A Cost Effective Approach for the Synthesis of Alterna...
Recycling Technology - A Cost Effective Approach for the Synthesis of Alterna...Recycling Technology - A Cost Effective Approach for the Synthesis of Alterna...
Recycling Technology - A Cost Effective Approach for the Synthesis of Alterna...
 
PLANT DESIGN FOR MANUFACTURING OF HYDROGEN BY STEAM METHANE REFORMING (SMR)
PLANT DESIGN FOR MANUFACTURING OF HYDROGEN BY STEAM METHANE REFORMING (SMR)PLANT DESIGN FOR MANUFACTURING OF HYDROGEN BY STEAM METHANE REFORMING (SMR)
PLANT DESIGN FOR MANUFACTURING OF HYDROGEN BY STEAM METHANE REFORMING (SMR)
 
BITE10281
BITE10281BITE10281
BITE10281
 
Biomass to olefins cracking of renewable naphtha
Biomass to olefins    cracking of renewable naphthaBiomass to olefins    cracking of renewable naphtha
Biomass to olefins cracking of renewable naphtha
 
Jurnal inter 2
Jurnal inter 2Jurnal inter 2
Jurnal inter 2
 
20180926230120_73171.pdf
20180926230120_73171.pdf20180926230120_73171.pdf
20180926230120_73171.pdf
 
SPE IPTC Poster Gas Hydrate
SPE IPTC Poster Gas HydrateSPE IPTC Poster Gas Hydrate
SPE IPTC Poster Gas Hydrate
 

Ähnlich wie 2020 young-ren ea

Parameters Analysis of the Assisted Combustion of Residual Biodiesel Glycerol
Parameters Analysis of the Assisted Combustion of Residual Biodiesel Glycerol Parameters Analysis of the Assisted Combustion of Residual Biodiesel Glycerol
Parameters Analysis of the Assisted Combustion of Residual Biodiesel Glycerol drboon
 
Optimization of key factors affecting biogas production from milk waste using...
Optimization of key factors affecting biogas production from milk waste using...Optimization of key factors affecting biogas production from milk waste using...
Optimization of key factors affecting biogas production from milk waste using...Lasbet Mohamed
 
BP Dalian Energy Innovation Laboratory 2013 brochure
BP Dalian Energy Innovation Laboratory 2013 brochureBP Dalian Energy Innovation Laboratory 2013 brochure
BP Dalian Energy Innovation Laboratory 2013 brochureSteve Wittrig
 
Supercritical fluid technology in biodiesel production
Supercritical fluid technology in biodiesel productionSupercritical fluid technology in biodiesel production
Supercritical fluid technology in biodiesel productionAlexander Decker
 
11.supercritical fluid technology in biodiesel production
11.supercritical fluid technology in biodiesel production11.supercritical fluid technology in biodiesel production
11.supercritical fluid technology in biodiesel productionAlexander Decker
 
GC-MS and FTIR analysis of bio-oil obtained from freshwater algae (spirogyra)...
GC-MS and FTIR analysis of bio-oil obtained from freshwater algae (spirogyra)...GC-MS and FTIR analysis of bio-oil obtained from freshwater algae (spirogyra)...
GC-MS and FTIR analysis of bio-oil obtained from freshwater algae (spirogyra)...Agriculture Journal IJOEAR
 
IJSRED-V2I1P18
IJSRED-V2I1P18IJSRED-V2I1P18
IJSRED-V2I1P18IJSRED
 
Oh functionalization by glycerol
Oh functionalization by glycerolOh functionalization by glycerol
Oh functionalization by glycerolAnthony Maputi
 
Propylene Production by Propane Dehydrogenation (PDH)
Propylene Production by Propane Dehydrogenation (PDH)Propylene Production by Propane Dehydrogenation (PDH)
Propylene Production by Propane Dehydrogenation (PDH)Amir Razmi
 
A REVIEW ON LIGNOCELLULOSIC BIOMASS CONVERSION TO DIESEL COMPATIBLE AND USEFU...
A REVIEW ON LIGNOCELLULOSIC BIOMASS CONVERSION TO DIESEL COMPATIBLE AND USEFU...A REVIEW ON LIGNOCELLULOSIC BIOMASS CONVERSION TO DIESEL COMPATIBLE AND USEFU...
A REVIEW ON LIGNOCELLULOSIC BIOMASS CONVERSION TO DIESEL COMPATIBLE AND USEFU...IAEME Publication
 
JBEI Science Highlights - February 2023
JBEI Science Highlights - February 2023JBEI Science Highlights - February 2023
JBEI Science Highlights - February 2023SaraHarmon5
 
A sustainability assessment of biofuel supply chain.pdf
A sustainability assessment of biofuel supply chain.pdfA sustainability assessment of biofuel supply chain.pdf
A sustainability assessment of biofuel supply chain.pdfAmber Ford
 
BIOMETHANE PRODUCTION USING ANAEROBIC DIGESTION OF FOOD WASTE BY IGBUDU SALVA...
BIOMETHANE PRODUCTION USING ANAEROBIC DIGESTION OF FOOD WASTE BY IGBUDU SALVA...BIOMETHANE PRODUCTION USING ANAEROBIC DIGESTION OF FOOD WASTE BY IGBUDU SALVA...
BIOMETHANE PRODUCTION USING ANAEROBIC DIGESTION OF FOOD WASTE BY IGBUDU SALVA...SalvationIgbudu1
 
Comparative Ethanol Productivities of Two Different Recombinant Fermenting St...
Comparative Ethanol Productivities of Two Different Recombinant Fermenting St...Comparative Ethanol Productivities of Two Different Recombinant Fermenting St...
Comparative Ethanol Productivities of Two Different Recombinant Fermenting St...IJERA Editor
 
Seminar defences 2
Seminar defences 2Seminar defences 2
Seminar defences 2Majur Mading
 
(Liquid liquid) equilibrium of systems involved in the stepwise ethanolysis o...
(Liquid liquid) equilibrium of systems involved in the stepwise ethanolysis o...(Liquid liquid) equilibrium of systems involved in the stepwise ethanolysis o...
(Liquid liquid) equilibrium of systems involved in the stepwise ethanolysis o...Josemar Pereira da Silva
 

Ähnlich wie 2020 young-ren ea (20)

Parameters Analysis of the Assisted Combustion of Residual Biodiesel Glycerol
Parameters Analysis of the Assisted Combustion of Residual Biodiesel Glycerol Parameters Analysis of the Assisted Combustion of Residual Biodiesel Glycerol
Parameters Analysis of the Assisted Combustion of Residual Biodiesel Glycerol
 
Optimization of key factors affecting biogas production from milk waste using...
Optimization of key factors affecting biogas production from milk waste using...Optimization of key factors affecting biogas production from milk waste using...
Optimization of key factors affecting biogas production from milk waste using...
 
Alghe oil
Alghe oilAlghe oil
Alghe oil
 
BP Dalian Energy Innovation Laboratory 2013 brochure
BP Dalian Energy Innovation Laboratory 2013 brochureBP Dalian Energy Innovation Laboratory 2013 brochure
BP Dalian Energy Innovation Laboratory 2013 brochure
 
Supercritical fluid technology in biodiesel production
Supercritical fluid technology in biodiesel productionSupercritical fluid technology in biodiesel production
Supercritical fluid technology in biodiesel production
 
11.supercritical fluid technology in biodiesel production
11.supercritical fluid technology in biodiesel production11.supercritical fluid technology in biodiesel production
11.supercritical fluid technology in biodiesel production
 
Jk3616291636
Jk3616291636Jk3616291636
Jk3616291636
 
GC-MS and FTIR analysis of bio-oil obtained from freshwater algae (spirogyra)...
GC-MS and FTIR analysis of bio-oil obtained from freshwater algae (spirogyra)...GC-MS and FTIR analysis of bio-oil obtained from freshwater algae (spirogyra)...
GC-MS and FTIR analysis of bio-oil obtained from freshwater algae (spirogyra)...
 
Formulationassessment
FormulationassessmentFormulationassessment
Formulationassessment
 
IJSRED-V2I1P18
IJSRED-V2I1P18IJSRED-V2I1P18
IJSRED-V2I1P18
 
Oh functionalization by glycerol
Oh functionalization by glycerolOh functionalization by glycerol
Oh functionalization by glycerol
 
Propylene Production by Propane Dehydrogenation (PDH)
Propylene Production by Propane Dehydrogenation (PDH)Propylene Production by Propane Dehydrogenation (PDH)
Propylene Production by Propane Dehydrogenation (PDH)
 
A REVIEW ON LIGNOCELLULOSIC BIOMASS CONVERSION TO DIESEL COMPATIBLE AND USEFU...
A REVIEW ON LIGNOCELLULOSIC BIOMASS CONVERSION TO DIESEL COMPATIBLE AND USEFU...A REVIEW ON LIGNOCELLULOSIC BIOMASS CONVERSION TO DIESEL COMPATIBLE AND USEFU...
A REVIEW ON LIGNOCELLULOSIC BIOMASS CONVERSION TO DIESEL COMPATIBLE AND USEFU...
 
JBEI Science Highlights - February 2023
JBEI Science Highlights - February 2023JBEI Science Highlights - February 2023
JBEI Science Highlights - February 2023
 
Gc taiwan paper 2
Gc taiwan paper 2Gc taiwan paper 2
Gc taiwan paper 2
 
A sustainability assessment of biofuel supply chain.pdf
A sustainability assessment of biofuel supply chain.pdfA sustainability assessment of biofuel supply chain.pdf
A sustainability assessment of biofuel supply chain.pdf
 
BIOMETHANE PRODUCTION USING ANAEROBIC DIGESTION OF FOOD WASTE BY IGBUDU SALVA...
BIOMETHANE PRODUCTION USING ANAEROBIC DIGESTION OF FOOD WASTE BY IGBUDU SALVA...BIOMETHANE PRODUCTION USING ANAEROBIC DIGESTION OF FOOD WASTE BY IGBUDU SALVA...
BIOMETHANE PRODUCTION USING ANAEROBIC DIGESTION OF FOOD WASTE BY IGBUDU SALVA...
 
Comparative Ethanol Productivities of Two Different Recombinant Fermenting St...
Comparative Ethanol Productivities of Two Different Recombinant Fermenting St...Comparative Ethanol Productivities of Two Different Recombinant Fermenting St...
Comparative Ethanol Productivities of Two Different Recombinant Fermenting St...
 
Seminar defences 2
Seminar defences 2Seminar defences 2
Seminar defences 2
 
(Liquid liquid) equilibrium of systems involved in the stepwise ethanolysis o...
(Liquid liquid) equilibrium of systems involved in the stepwise ethanolysis o...(Liquid liquid) equilibrium of systems involved in the stepwise ethanolysis o...
(Liquid liquid) equilibrium of systems involved in the stepwise ethanolysis o...
 

Kürzlich hochgeladen

Generative AI or GenAI technology based PPT
Generative AI or GenAI technology based PPTGenerative AI or GenAI technology based PPT
Generative AI or GenAI technology based PPTbhaskargani46
 
Bhubaneswar🌹Call Girls Bhubaneswar ❤Komal 9777949614 💟 Full Trusted CALL GIRL...
Bhubaneswar🌹Call Girls Bhubaneswar ❤Komal 9777949614 💟 Full Trusted CALL GIRL...Bhubaneswar🌹Call Girls Bhubaneswar ❤Komal 9777949614 💟 Full Trusted CALL GIRL...
Bhubaneswar🌹Call Girls Bhubaneswar ❤Komal 9777949614 💟 Full Trusted CALL GIRL...Call Girls Mumbai
 
Thermal Engineering Unit - I & II . ppt
Thermal Engineering  Unit - I & II . pptThermal Engineering  Unit - I & II . ppt
Thermal Engineering Unit - I & II . pptDineshKumar4165
 
Hostel management system project report..pdf
Hostel management system project report..pdfHostel management system project report..pdf
Hostel management system project report..pdfKamal Acharya
 
A CASE STUDY ON CERAMIC INDUSTRY OF BANGLADESH.pptx
A CASE STUDY ON CERAMIC INDUSTRY OF BANGLADESH.pptxA CASE STUDY ON CERAMIC INDUSTRY OF BANGLADESH.pptx
A CASE STUDY ON CERAMIC INDUSTRY OF BANGLADESH.pptxmaisarahman1
 
kiln thermal load.pptx kiln tgermal load
kiln thermal load.pptx kiln tgermal loadkiln thermal load.pptx kiln tgermal load
kiln thermal load.pptx kiln tgermal loadhamedmustafa094
 
Learn the concepts of Thermodynamics on Magic Marks
Learn the concepts of Thermodynamics on Magic MarksLearn the concepts of Thermodynamics on Magic Marks
Learn the concepts of Thermodynamics on Magic MarksMagic Marks
 
Unleashing the Power of the SORA AI lastest leap
Unleashing the Power of the SORA AI lastest leapUnleashing the Power of the SORA AI lastest leap
Unleashing the Power of the SORA AI lastest leapRishantSharmaFr
 
Navigating Complexity: The Role of Trusted Partners and VIAS3D in Dassault Sy...
Navigating Complexity: The Role of Trusted Partners and VIAS3D in Dassault Sy...Navigating Complexity: The Role of Trusted Partners and VIAS3D in Dassault Sy...
Navigating Complexity: The Role of Trusted Partners and VIAS3D in Dassault Sy...Arindam Chakraborty, Ph.D., P.E. (CA, TX)
 
Double Revolving field theory-how the rotor develops torque
Double Revolving field theory-how the rotor develops torqueDouble Revolving field theory-how the rotor develops torque
Double Revolving field theory-how the rotor develops torqueBhangaleSonal
 
HAND TOOLS USED AT ELECTRONICS WORK PRESENTED BY KOUSTAV SARKAR
HAND TOOLS USED AT ELECTRONICS WORK PRESENTED BY KOUSTAV SARKARHAND TOOLS USED AT ELECTRONICS WORK PRESENTED BY KOUSTAV SARKAR
HAND TOOLS USED AT ELECTRONICS WORK PRESENTED BY KOUSTAV SARKARKOUSTAV SARKAR
 
Introduction to Serverless with AWS Lambda
Introduction to Serverless with AWS LambdaIntroduction to Serverless with AWS Lambda
Introduction to Serverless with AWS LambdaOmar Fathy
 
Minimum and Maximum Modes of microprocessor 8086
Minimum and Maximum Modes of microprocessor 8086Minimum and Maximum Modes of microprocessor 8086
Minimum and Maximum Modes of microprocessor 8086anil_gaur
 
Air Compressor reciprocating single stage
Air Compressor reciprocating single stageAir Compressor reciprocating single stage
Air Compressor reciprocating single stageAbc194748
 
S1S2 B.Arch MGU - HOA1&2 Module 3 -Temple Architecture of Kerala.pptx
S1S2 B.Arch MGU - HOA1&2 Module 3 -Temple Architecture of Kerala.pptxS1S2 B.Arch MGU - HOA1&2 Module 3 -Temple Architecture of Kerala.pptx
S1S2 B.Arch MGU - HOA1&2 Module 3 -Temple Architecture of Kerala.pptxSCMS School of Architecture
 
Computer Lecture 01.pptxIntroduction to Computers
Computer Lecture 01.pptxIntroduction to ComputersComputer Lecture 01.pptxIntroduction to Computers
Computer Lecture 01.pptxIntroduction to ComputersMairaAshraf6
 
+97470301568>> buy weed in qatar,buy thc oil qatar,buy weed and vape oil in d...
+97470301568>> buy weed in qatar,buy thc oil qatar,buy weed and vape oil in d...+97470301568>> buy weed in qatar,buy thc oil qatar,buy weed and vape oil in d...
+97470301568>> buy weed in qatar,buy thc oil qatar,buy weed and vape oil in d...Health
 
Rums floating Omkareshwar FSPV IM_16112021.pdf
Rums floating Omkareshwar FSPV IM_16112021.pdfRums floating Omkareshwar FSPV IM_16112021.pdf
Rums floating Omkareshwar FSPV IM_16112021.pdfsmsksolar
 

Kürzlich hochgeladen (20)

Generative AI or GenAI technology based PPT
Generative AI or GenAI technology based PPTGenerative AI or GenAI technology based PPT
Generative AI or GenAI technology based PPT
 
Bhubaneswar🌹Call Girls Bhubaneswar ❤Komal 9777949614 💟 Full Trusted CALL GIRL...
Bhubaneswar🌹Call Girls Bhubaneswar ❤Komal 9777949614 💟 Full Trusted CALL GIRL...Bhubaneswar🌹Call Girls Bhubaneswar ❤Komal 9777949614 💟 Full Trusted CALL GIRL...
Bhubaneswar🌹Call Girls Bhubaneswar ❤Komal 9777949614 💟 Full Trusted CALL GIRL...
 
Thermal Engineering Unit - I & II . ppt
Thermal Engineering  Unit - I & II . pptThermal Engineering  Unit - I & II . ppt
Thermal Engineering Unit - I & II . ppt
 
Hostel management system project report..pdf
Hostel management system project report..pdfHostel management system project report..pdf
Hostel management system project report..pdf
 
A CASE STUDY ON CERAMIC INDUSTRY OF BANGLADESH.pptx
A CASE STUDY ON CERAMIC INDUSTRY OF BANGLADESH.pptxA CASE STUDY ON CERAMIC INDUSTRY OF BANGLADESH.pptx
A CASE STUDY ON CERAMIC INDUSTRY OF BANGLADESH.pptx
 
kiln thermal load.pptx kiln tgermal load
kiln thermal load.pptx kiln tgermal loadkiln thermal load.pptx kiln tgermal load
kiln thermal load.pptx kiln tgermal load
 
Learn the concepts of Thermodynamics on Magic Marks
Learn the concepts of Thermodynamics on Magic MarksLearn the concepts of Thermodynamics on Magic Marks
Learn the concepts of Thermodynamics on Magic Marks
 
Call Girls in South Ex (delhi) call me [🔝9953056974🔝] escort service 24X7
Call Girls in South Ex (delhi) call me [🔝9953056974🔝] escort service 24X7Call Girls in South Ex (delhi) call me [🔝9953056974🔝] escort service 24X7
Call Girls in South Ex (delhi) call me [🔝9953056974🔝] escort service 24X7
 
Unleashing the Power of the SORA AI lastest leap
Unleashing the Power of the SORA AI lastest leapUnleashing the Power of the SORA AI lastest leap
Unleashing the Power of the SORA AI lastest leap
 
Cara Menggugurkan Sperma Yang Masuk Rahim Biyar Tidak Hamil
Cara Menggugurkan Sperma Yang Masuk Rahim Biyar Tidak HamilCara Menggugurkan Sperma Yang Masuk Rahim Biyar Tidak Hamil
Cara Menggugurkan Sperma Yang Masuk Rahim Biyar Tidak Hamil
 
Navigating Complexity: The Role of Trusted Partners and VIAS3D in Dassault Sy...
Navigating Complexity: The Role of Trusted Partners and VIAS3D in Dassault Sy...Navigating Complexity: The Role of Trusted Partners and VIAS3D in Dassault Sy...
Navigating Complexity: The Role of Trusted Partners and VIAS3D in Dassault Sy...
 
Double Revolving field theory-how the rotor develops torque
Double Revolving field theory-how the rotor develops torqueDouble Revolving field theory-how the rotor develops torque
Double Revolving field theory-how the rotor develops torque
 
HAND TOOLS USED AT ELECTRONICS WORK PRESENTED BY KOUSTAV SARKAR
HAND TOOLS USED AT ELECTRONICS WORK PRESENTED BY KOUSTAV SARKARHAND TOOLS USED AT ELECTRONICS WORK PRESENTED BY KOUSTAV SARKAR
HAND TOOLS USED AT ELECTRONICS WORK PRESENTED BY KOUSTAV SARKAR
 
Introduction to Serverless with AWS Lambda
Introduction to Serverless with AWS LambdaIntroduction to Serverless with AWS Lambda
Introduction to Serverless with AWS Lambda
 
Minimum and Maximum Modes of microprocessor 8086
Minimum and Maximum Modes of microprocessor 8086Minimum and Maximum Modes of microprocessor 8086
Minimum and Maximum Modes of microprocessor 8086
 
Air Compressor reciprocating single stage
Air Compressor reciprocating single stageAir Compressor reciprocating single stage
Air Compressor reciprocating single stage
 
S1S2 B.Arch MGU - HOA1&2 Module 3 -Temple Architecture of Kerala.pptx
S1S2 B.Arch MGU - HOA1&2 Module 3 -Temple Architecture of Kerala.pptxS1S2 B.Arch MGU - HOA1&2 Module 3 -Temple Architecture of Kerala.pptx
S1S2 B.Arch MGU - HOA1&2 Module 3 -Temple Architecture of Kerala.pptx
 
Computer Lecture 01.pptxIntroduction to Computers
Computer Lecture 01.pptxIntroduction to ComputersComputer Lecture 01.pptxIntroduction to Computers
Computer Lecture 01.pptxIntroduction to Computers
 
+97470301568>> buy weed in qatar,buy thc oil qatar,buy weed and vape oil in d...
+97470301568>> buy weed in qatar,buy thc oil qatar,buy weed and vape oil in d...+97470301568>> buy weed in qatar,buy thc oil qatar,buy weed and vape oil in d...
+97470301568>> buy weed in qatar,buy thc oil qatar,buy weed and vape oil in d...
 
Rums floating Omkareshwar FSPV IM_16112021.pdf
Rums floating Omkareshwar FSPV IM_16112021.pdfRums floating Omkareshwar FSPV IM_16112021.pdf
Rums floating Omkareshwar FSPV IM_16112021.pdf
 

2020 young-ren ea

  • 1. Propylene glycol from glycerol: Process evaluation and break-even price determination Roberto X. Jimenez a , Andre F. Young b, * , Heloisa L.S. Fernandes a a Department of Chemical Engineering, Federal University of Rio de Janeiro, Av. Athos da Silveira Ramos, 149, Rio de Janeiro, Brazil b Department of Chemical and Petroleum Engineering, Federal Fluminense University, Rua Passo da Patria, 156, Niteroi, Brazil a r t i c l e i n f o Article history: Received 22 November 2019 Received in revised form 27 April 2020 Accepted 23 May 2020 Available online 31 May 2020 Keywords: Glycerol Process simulation Propylene glycol Hydrogen Reforming a b s t r a c t Propylene glycol can be produced from the hydrogenolysis of glycerol. If hydrogen comes from a renewable source, the produced propylene glycol can be considered completely renewable. The objective of this work is to perform a technical and economic evaluation of a propylene glycol production process which uses glycerol as the main raw material. Two possibilities regarding the hydrogen source are explored: it may come from an external source or be produced locally. It is shown that the break-even price of the partially renewable propylene glycol is lower than the market price of propylene glycol, which indicates that this process is economically viable. It is also shown that the price of hydrogen produced via steam reforming is not competitive with its price if produced by other sources. Conse- quently, the fully renewable propylene glycol production process yields a lower profit, but it is still viable and depends on the market. © 2020 Elsevier Ltd. All rights reserved. 1. Introduction Glycerol is the main by-product of biodiesel production [1e3]. Along with the increased supply of biodiesel, glycerol production has also increased. In 2017, there were produced about 374.5 thousand cubic meters of glycerol, which meant a 9.5% increase in comparison with 2016 [4]. The increasing glycerol surplus has motivated the development of chemical processes which employ it as the main raw material. As a consequence, the number of published scientific articles on the use of glycerol as the main raw material in chemical processes has increased. For example, articles related to glycerol in the Scopus database have increased by almost 7 times per year between 2008 and 2016 [3]. Among the products that can be obtained from glycerol there are propanediols (1,2- and 1,3-propanediol), propene, epichloro- hydrin, methanol, ethanol, lactic acid, citric acid, propanoic acid and hydrogen [5e7]. Monteiro [3] and Freitas [6] identified that 1,3- propanediol, hydrogen and 1,2-propanediol or propylene glycol (PG) are the products formed from glycerol on which researchers have published more articles. According to these authors, the number of patents or patent applications on PG production from glycerol was larger than those on 1,3-propanediol or hydrogen production until 2013. Propylene glycol can be used as a surfactant, a moisturizer, an antifreeze agent, a solvent, a preservative and in detergent for- mulations [6,8]. It is usually obtained by the hydrolysis of propylene oxide [8,9]. Propylene oxide is produced from propylene hydro- chloride, which comes from the petrochemical industry. In Brazil, the only registered producer of propylene glycol is Dow®. It has a chemical plant located in the industrial complex of Aratu, in the state of Bahia. The company also produces propylene oxide, so propylene glycol is still produced by the conventional route from petrochemicals. It is noteworthy that this is the only propylene glycol production plant in Latin America [10]. A simpli- fied flowsheet of this process technology is shown in Fig. 1. Alternatively, PG can be produced by a more ecologically sound route, which employs glycerol as a carbon and hydrogen source, and which relies less on fossil sources. Some scientific articles about this route were published, but there are only a few patents (granted or under application) on it. Currently, if USPTO and Lens databases are searched for patents on processes involving the keywords propylene glycol, glycerol/glycerin and 1,2-propanediol, glycerol/ glycerin, only 12 results are retrieved. Among them, 10 patents describe processes with liquid-phase reactors, one that employs a vapor-phase reactor and one with a two-phase reactor (reactions* Corresponding author. E-mail address: ayoung@id.uff.br (A.F. Young). Contents lists available at ScienceDirect Renewable Energy journal homepage: www.elsevier.com/locate/renene https://doi.org/10.1016/j.renene.2020.05.126 0960-1481/© 2020 Elsevier Ltd. All rights reserved. Renewable Energy 158 (2020) 181e191
  • 2. start in the liquid phase and generate acetol, which is transferred to the vapor phase to react). It is interesting to note that, although the search was not limited by date, all the results date to the last 8 years. These results indicate that this technology is new and may require more studies until consolidation. Considering this, the objective of this work is to perform a technical and economic assessment of the partially renewable and fully renewable PG production processes from glycerol, to find out if these processes are economically viable in the current Brazilian economic scenario. Equipment design and process simulations were performed with the aid of the Aspen HYSYS® v8.8 process simulator. 2. Process simulation and equipment design In order to proceed to the economic analysis, the reaction ki- netics and vapor-liquid equilibria must be correctly reproduced in the simulations so that the equipment can be properly designed. The properties of all components involved in the system are available in the Aspen HYSYS® simulator database. 2.1. Propylene glycol production from glycerol Glycerol (or glycerin) obtained from biodiesel is available in different purity levels. Crude glycerin is defined as the glycerol-rich mixture (around 75% glycerol) obtained after the separation of biodiesel. After neutralization, glycerin is usually called blonde; its weight percent composition is 85.0% glycerol, 11.7% water, 3.2% methanol, and 0.2% biodiesel. After distillation, USP grade glycerol or bidistilled glycerol (99.5% glycerol) is obtained [11]. The advan- tage of using blonde glycerin as a raw material is that, as the puri- fication process has a cost, it ends up being cheaper. However, bidistilled glycerol is regarded as the process raw material in this work, since kinetic data for the hydrogenolysis and reforming re- actions refer to glycerol-water mixtures only [12]. Furthermore, according to Resolution no. 30 published by ANP (Brazilian National Agency of Petroleum, Natural Gas and Biofuels), which regulates biodiesel production activities in Brazil [4], including the construction stage, the glycerol purification stage must be present in every biodiesel production unit. Thus, in all plants currently operating in Brazil, a stream of composition close to that of bidistilled glycerol is available, or at least its composition can be adjusted for this purpose. In literature, a wide variety of metal catalysts has been used to promote the CeO bond cleavage reaction of the glycerol molecule. Cu-based catalysts are widely studied as they offer a high degree of conversion and selectivity of PG [13e20]. Two mechanisms have been proposed for the production of propylene glycol from glycerol: according to the dehydration- hydrogenation mechanism, glycerol is initially dehydrated at catalyst acid sites to acetol, which is hydrogenated to 1,2- propanediol [5,13,21]. A three-phase dehydrogenation-dehydra- tion-hydrogenation mechanism is also suggested [15,17,22,23]. The literature recommends a Langmuir-Hinshelwood-type expression to model the heterogeneous kinetics of formation of propylene glycol and by-products (1,3-propanediol, ethylene glycol, ethanol, methanol, propanol and propanoic acid) through the dehydration-hydrogenation mechanism [12,19,24,25]. Two types of catalyst active sites are considered, as observed in Table 1 [12,26]. Zhou et al. [12] employed a CueZnOeAl2O3 catalyst with a molar ratio of 1:1:0.5 of Cu/Zn/Al, with a particle diameter of 0.17 mm (80e100 mesh), and reported a conversion of 81.5% of the glycerol and selectivity of 93.4% in terms of propylene glycol. Also, the authors proposed the rate expressions of the two reaction steps presented in Table 1, which can be found in Eqs. (1) and (2). r1 ¼ k1bGCG ð1 þ bGCG þ bACA þ bPCPÞ (1) r2 ¼ k2bGCG ð1 þ bGCG þ bACA þ bPCPÞ 1 þ ffiffiffiffiffiffiffiffiffiffiffiffi bHPH p 2 (2) In Eqs. (1) and (2), Ci is the molar concentration of component i, PH is the partial pressure of hydrogen, ki is the rate constant of reaction i and bi is the absorption rate constant of component i [12]. The numerical values of these constants are described in Table 2. A preliminary simulation of the reaction in laboratory conditions was carried out in Aspen HYSYS®; simulated and experimental data [12] were compared and relative errors of 3.2% in the glycerol conversion and 7.0% in the acetol to propylene glycol selectivity ratio were found. Since average relative errors between experi- mental and theoretical glycerol and propylene glycol flowrates at Fig. 1. Simplified propylene oxide hydration process flowsheet. Table 1 Steps involved in the reaction of propylene glycol formation over a Cu/ZnO/Al2O3 catalyst. G ¼ glycerol, A ¼ acetol, PG ¼ propylene glycol, Q1 type 1 active site, Q2 type 2 active site [12]. Stage Kinetic Adsorption/Desorption H2 þ 2Q1 #2HQ1 e bH G þ Q2 #GQ2 e bG GQ2/AQ2 þ H2O k1 e 2HQ1 þ AQ2 /PGQ2 þ 2Q1 k2 e AQ2 #A þ Q2 e bÀ1 A PGQ2 #P þ Q2 e bÀ1 PG R.X. Jimenez et al. / Renewable Energy 158 (2020) 181e191182
  • 3. the reactor output are 6.3% and 7.6%, respectively [12], deviations obtained in the preliminary simulation can be considered accept- able. Thus, the rate expressions in Eqs. (1) and (2) are used in this work. The UNIQUAC model was used for phase equilibria calculations for the system containing the components involved in the pro- duction of propylene glycol, since it provides a satisfactory description of phase equilibria for similar systems [27]. The propylene glycol process flowsheet can be seen in Fig. 2. The stream properties are presented as Supporting Information, Table S3. The raw material is fed into the process and is put under operating conditions before entering the reactor. Excess hydrogen is used to facilitate mass transfer in the three-phase system within the reactor. The global reaction is shown in Eq. (3). C3H8O3 þ H2/C3H8O2 þ H2O Glycerol PG (3) The reaction is carried out at T ¼ 493 K, PH ¼ 4 MPa and a hydrogen to glycerol molar ratio of 5:1. The catalyst is assumed to be CueZnOeAl2O3 and the residence time is adjusted to 85 kgCat:h:kg:molÀ1 Glycerol, in order to reproduce the results obtained by Zhou et al. [12]. Moreover, the authors compared their results at 493 K with others available in the literature [21], and showed consistency in the values that were reported. Zhou et al. [12] proved that glycerol conversion increases with higher pressures (testing 3 MPa, 4 MPa and 5 MPa), but at 493 K it remains almost the same, independent on pressure. A conservative approach was used to choose a 4 MPa pressure for the reaction simulation. The plant was designed to operate with a flowrate of 1070 kg/h of glycerol. This was the average glycerol production in biodiesel plants in Brazil in 2018 [4]. The reactor effluent is mainly composed by propylene glycol, hydrogen, unreacted glycerol and by-products. The product stream exchanges heat with the glycerol stream to minimize utility ex- penses. Then it follows to the separation section, in which excess hydrogen is recovered with a flash vessel, recycled and mixed with pure hydrogen, thereby ensuring a 5:1 M ratio of hydrogen to glycerol at the process inlet. Finally, two distillation columns adjust the propylene glycol purity to the required value (USP grade, 99.5%). The literature suggests the use of a trickle-bed-reactor for glycerol hydrogenolysis in industrial-scale processes [25,28,29]. Inside the reactor, liquid and gas flow downward through the cat- alytic bed. The net flowrate must be low enough so that the liquid phase forms a thin film over the fixed bed and descends by gravity and gas-phase drag [30]. Thus, the contact among the three phases is maximized. Trickle-bed-reactors have been widely used in other processes involving hydrogen, mainly in the hydrodesulfurization of petroleum fractions [30,31]. The catalyst density was not mentioned in the articles reviewed but is required for the calculation of the fixed-bed volume. The fixed-bed volume includes the volume occupied by the catalyst solid mass (30%) and the volume of pore voids and voids among the catalyst particles (70%), which are generally in a proportion of 3:3:4 [32]. The solid density was calculated from the molar fraction of its components. Assuming the void fraction in the bed to be equal to 0.4, the densities of solid, pellets and bulk can be estimated. For the purpose of reactor sizing, the trickle-bed model was simplified to a plug-flow fixed bed model, as available in Aspen HYSYS®. The required catalyst mass was obtained by solving the design equations for a plug-flow reactor. A plug-flow reactor which operates at conditions similar to those in the work by Zhou et al. [12] requires 992.8 kg of catalyst. Likewise, a conversion of 98.2% requires 1489 kg of catalyst, while a conversion of 99.99% requires an additional 993 kg of catalyst [12]. Consequently, it is not convenient to try to achieve such high glycerol conversions. Thus, the plant was designed to operate with an overall glycerol Table 2 Parameters for the kinetic model of the glycerol hydrogenolysis reaction [12]. Kinetic and absorption constants Parameters Pre-exponential Factor mol kg s Activation Energy kJ mol k1 1.54 Â 107 86.56 k2 7.16 Â 106 57.8 bG 2.22 36.42 bA 8.73 25.94 bPG 5.80 25.77 bH 1.86 Â 10À5 MPa 36.24 Fig. 2. Propylene glycol production flowsheet. R.X. Jimenez et al. / Renewable Energy 158 (2020) 181e191 183
  • 4. conversion of 98.2%. Finally, the reactor volume can be calculated by the ratio of catalyst mass and bulk density. The reaction must occur under isothermal conditions. To meet this specification, a cooling jacket was applied, with water at 298 K as coolant. The heat exchange area was assumed to be approxi- mately equal to the lateral area of the plug-flow reactor. In addition to acetol and propylene glycol, the authors identified the presence of by-products such as ethylene glycol, methanol, ethanol, propanol and propanoic acid in the product stream, among which ethylene glycol is the most expressive. Therefore, it is necessary to simulate the formation of these products so that the simulation is in agreement with the experimental study. Ethylene glycol and methanol modeling was performed by specifying the conversion value found experimentally [12]. Due to the lack of information on the proportion of the remaining by- products, their simulation was not feasible. Since the by-products are lower alcohols (except for propanoic acid), they were regar- ded as a single representative compound (methanol) in the simulations. In a preliminary analysis of the separation system, it was found that for a better recovery of unreacted hydrogen it is necessary to cool the product stream from the reactor. To save resources and accomplish energy integration, it was decided to implement a shell-and-tube heat exchanger through which this stream and the glycerol and water feed stream are passed. The heat transfer rate, the heat exchange area and the global heat transfer coefficient were obtained with the aid of the simulator, from the values of the inlet and outlet streams’ temperatures, pressures and flowrates. A minimum temperature approach of 10 K was used to avoid an excessively large heat exchange area. The initial estimate of pres- sure drop on both sides of the exchanger was 10 kPa [33]. The heat exchanger is designed with floating heads to minimize thermal expansion effects that may occur for a temperature difference greater than 30 K between the tube fluid (reactor effluent) and the shell fluid [34]. The methodology described by Seider et al. [33], which is based on the relative volatility of the compounds, was adopted for the synthesis of the separation system. The reactor outlet is a two- phase stream, which consists of hydrogen, methanol, water, ace- tol, propylene glycol, ethylene glycol and glycerol (from the most volatile to the least volatile). Three cuts were required to achieve the specified purity: Hydrogen/Methanol, Water/Propylene Glycol and Propylene Glycol/Ethylene Glycol. Hydrogen is the first component that must be removed, as it is the excess reagent and can be recycled into the process, thereby reducing raw materials expenditures. Thus, for the separation of hydrogen from the reactor effluent stream, a flash vessel was used. The liquid-gas separator should be such that the vessel diameter must be large enough so that the gas velocity is less than the liquid droplet separation velocity and that the vessel height is equal to its diameter, or 1 m if the diameter is less than 1 m. It is recommended that hold-up time is about 10 min so that the liquid level inside the equipment is appropriate. Finally, the feed inlet must be at least 60 cm above the liquid level [35]. To maximize the recovery of pure hydrogen, it is necessary that the flash vessel also removes heat from the streams, to promote condensation of the less volatile components. Such heat exchange was performed by implementing a coil inside the vessel with cooling water at 298 K. Aspen HYSYS® Adjust block was used so that the cooling rate varies to ensure that the purity of the hydrogen obtained is at least 99.5%. The next two cuts were made through distillation columns. The first column cuts between water (light key) and propylene glycol (heavy key), and the second column adjusts the purity of the final product to the specified value by cutting between propylene glycol (light key) and ethylene glycol (heavy key). The columns were simulated rigorously and its input parameters were estimated based on typical methods. The minimum number of stages was calculated using the Fenske equation. The number of theoretical stages was obtained by the Gilland-Molokanov correlation. The minimum reflux ratio was estimated from the Underwood equation. Finally, the Kirkbride equation was used to determine the optimum feed stage [35]. It is noteworthy that the calculations of these estimates were per- formed in the simulator itself, using the Shortcut Distillation block. From the obtained values, it was possible to start the rigorous design of the distillation columns. In a first analysis, it was found that the operation under atmo- spheric pressure demands extreme temperatures in condensers and reboilers. As a consequence, common utilities such as steam and cooling water could not be used. Also, degradation or poly- merization reactions of residual glycerol may occur at elevated temperatures (about 473 K) [36]. Thus, to avoid such problems, it was decided that the equipment should operate under vacuum. The operating pressure was set as close as possible to atmospheric pressure, but the reboilers tem- peratures were kept at least 15 K below the glycerol degradation temperature. The literature suggests the use of packed columns for vacuum distillation and high viscosity mixtures, as in this case [35]. Structured packings offer better separation efficiencies and pres- sure drops, but are more expensive. Thus, the columns were designed with random packings of 1 in Pall metal rings, due to the high number of theoretical stages needed [33]. For the calculation of the packed bed height, the Height Equiv- alent to a Theoretical Plate (HETP) method was used. For vacuum distillation, the required column diameter can be estimated from the flooding condition. This condition should be avoided and can be estimated from the flooding velocity through the Leva correlation [33]. The results obtained in the simulation are in agreement with the work of Gandarias et al. [37]. 2.2. Hydrogen production: glycerol steam reforming Reforming of oxygenated hydrocarbons consists of the cleavage of the CeC bond to form CO and H2. Then CO and water may react in a water-gas shift reaction, releasing more H2 and CO2 [38]. Glycerol reforming needs a catalyst that promotes cleavage of the CeC, OeH and CeH bonds, and if possible also promote the shift reaction [38,39]. Reforming can be conducted in both liquid and gaseous phases. In the latter case, it is necessary to vaporize glycerol, what requires higher temperatures [40]. The water-gas shift reaction occurs only in the gas phase. Glycerol aqueous phase reforming occurs at higher pressures and lower temperatures, if compared to steam reforming, and of- fers greater possibilities for using crude glycerin as a raw material [41]. The major disadvantage of this route is its low reaction rates [42]. Vapor-phase reforming, in turn, can be conducted at atmo- spheric pressures and has presented energetic advantages and higher yield in terms of H2 [43]. That is why this was the route chosen in this work. In literature, there are studies about the activity of different catalysts and operating conditions that maximize the glycerol reforming and shift reactions, to the detriment of methanation and coke formation. Such reactions decrease the efficiency of the pro- cess in terms of H2 generation and impair the catalytic activity by carbon deposition at the active catalyst sites [44,45]. Catalysts involving noble metals, such as Pt, Ru and Pd, and non- noble metals, such as Ni and Co, have been successfully tested in several studies for both aqueous- and vapor-phase reforming [45e47]. The main disadvantage of nickel-based catalysts is coke R.X. Jimenez et al. / Renewable Energy 158 (2020) 181e191184
  • 5. formation, whereas for noble metal catalysts the high cost is an important drawback. Kinetic modeling of reforming of oxygenated compounds such as ethanol and methanol has been satisfactorily done through power-law expressions [48]. The activation energy depends on the type of catalyst employed in the reactions. For noble metals (Pt and Ru), the reaction rate expression depends linearly on glycerol concentration due to the high activity of such catalysts, while for other metals it is a function of a fractional power of the glycerol concentration. The non-dependence of the rate on water concen- tration is because it is used in excess [49]. Table 3 presents glycerol reforming kinetic parameters for power-law expressions obtained for different catalysts. The Soave-Redlich-Kwong (SRK) equation was chosen to calculate the thermodynamic properties in this simulation. This choice is supported by glycerol reforming studies available in the literature [50e54]. The glycerol steam reforming process for hydrogen production is designed to meet the hydrogen demand required for hydro- genolysis and is shown in Fig. 3. The stream properties are pre- sented as Supporting Information, Table S4. Glycerol and water enter the process at a 1:9 ratio. The reforming process involves three reaction steps, carried out in three different reactors. Such reactors operate under different conditions and employ different catalysts [55]. The first reactor performs reforming itself, as described in Eq. (4): C3H8O3#3 CO þ 4 H2 (4) The water-gas shift reaction occurs in the second and third re- actors, as described in Eq. (5): CO þ H2O#CO2 þ H2 (5) It is recommended to perform this reaction in two steps because it is exothermic. So, the first reactor, which is called the high- temperature shift reactor (HTS), operates at 648 K and atmo- spheric pressure. The second reactor, which is called the low- temperature shift reactor (LTS), operates at 498 K and at the same pressure, in order to shift equilibrium in the products direction [52]. A Fe2O3 catalyst was used for the HTS reactor. This catalyst has a low thermal deactivation rate and allows to promote the kinetics by increasing temperaure. For the LTS reactor, a high-activity CuO catalyst was used, which increases the products formation rate under lower temperatures [52]. The last reactor effluent consists mainly of H2, CO2, CO traces and water vapor, which is condensed, separated and recycled. The surplus stream enters a Pressure Swing Adsorption (PSA) separa- tion system, which produces hydrogen at a purity of 99.99%. The reforming reactor and the two shift reactors are of the fixed- bed PFR type. Thus, for design, it is first necessary to evaluate the amount of catalyst to be employed and then calculate the reactor volume. For the first reactor, the raw material must be heated to the operational temperature of 973 K and the effluent must be cooled down to 648 K, at which the HTS reactor operates, and then finally cooled to 498 K for the LTS [55]. Due to the lack of information on all reactions in a same study and under the same conditions, the conversion values and opera- tional conditions adopted by Villaça [52] were used. The conversion values obtained for methane generation and carbon formation are much lower than that of the main by-products, so it was decided to neglect these by-products. The carbon formed will accumulate at the active catalyst sites over time, thereby decreasing the efficiency of the process. This issue was considered in the economic evalua- tion. Methane molar fraction decreases with increasing tempera- ture, and for the temperature adopted in this work (973 K), no methane is detected in the reactor output stream. It can also be observed that the amount of H2 produced almost remains constant with increasing temperature, which is in agreement with experi- ments [41]. As the temperature increases, the molar fraction of carbon monoxide increases. This is because the shift reaction (which transforms CO to CO2) is exothermic and its yield is higher at low temperatures. To model the two shift reactors (HTS and LTS), the Equilibrium Reactor block from Aspen HYSYS® was used to calculate the output streams’ compositions. This can be justified by the fact that this reaction indeed tends to equilibrium in the reactor. This stage of the simulation is different from that presented by Villaça [52], but is in agreement with the considerations exposed earlier. Aspen HYSYS® contains in its database the equilibrium constant value of this reaction as a function of temperature. Excess water is added to the process. Thus, after the LTS reactor, water must be condensed in a flash vessel and recirculated. Aspen HYSYS® does not allow the simulation of adsorption processes with conventional blocks. Thus, the PSA unit is simplified to a Splitter block, in which a process stream is separated into several streams with defined compositions. In the present work, purity and separation data reported in the work of Villaça [52] and a pressure of 0.7 MPa were considered. For sizing, two adsorption towers were considered. The columns operate in cycles, thus ensuring that the process operates contin- uously. The adsorption columns are filled with a type 5A alumi- nosilicate molecular sieve, indicated to separate carbon dioxide from hydrogen-rich streams [30]. The quantity of adsorbent can be calculated considering an average adsorption capacity value of 0.255 kg of adsorbent per kg of adsorbate. With the bulk density, the bed volume is obtained [30,56]. 2.3. Economic analysis The total investment cost was estimated by means of the Lang method. The FOB price of each equipment was estimated from correlations available in the literature [33]. These correlations were calculated for a specific date, so updating is required. It can be done Table 3 Kinetic parameters of power-law models for glycerol steam reforming [49]. Catalyst Temperature (K) Glycerol Order Steam Order Activation Energy kJ mol Ru= Al2O3 623e773 1 e 21.2 Pt= C 623e673 1 e e Co= Al2O3 723e823 0.1 0.4 67.2 Ni= Al2O3 723e823 0.48 0.34 60.0 Co À Ni= Al2O3 773e823 0.25 0.36 63.3 Ni= CeO2 873e923 0.233 e 103.4 Ni À ZrO2= CeO2 973 0.3 e 43.4 R.X. Jimenez et al. / Renewable Energy 158 (2020) 181e191 185
  • 6. with the Plant Cost Index, published periodically by Chemical En- gineering Magazine. The value of the Plant Cost Index for the year on which the correlations were based is 394. For 2018 it is 603.1 [57]. For non-corrosive environments with hydrogen, Seider et al. [33] recommend the use of a 1% Cr and 0.5% Mo steel alloy (ASME code SA-387B). This material was applied only in large equipment used in operations involving the presence of hydrogen. For the calculation of production costs, it was considered that the plant operates continuously 24 h a day for 330 days a year. Prices for raw materials, utilities and catalysts are summarized in Table 4. It was not possible to stablish an average price for hydrogen, because several production processes employ a variety of raw materials and energy sources. For instance, for hydrogen produced via water electrolysis the cost may vary from 1.28 to 4.14 US$/kg, these values being the most recent that could be found in the literature [58]. Negro et al. [59] presented the different costs of hydrogen in Brazil if it was produced via electrolysis, fossil fuel reforming, or renewable fuel reforming. However, these values are out of date and may affect the veracity of the economic analysis. For this reason, these costs were updated using the Brazilian National Consumer Price Index [60], that represents the accumulated infla- tion in the country over time. The results can be observed in Table 5. Thus, in the present work, the hydrogen price range is 0.23e4.14 US$/kg, considering the lowest price of hydrogen if it is obtained by natural gas reforming and the highest price if it is obtained by water electrolysis. The price of hydrogen produced by electrolysis using solar energy was disregarded, since its value is very different from others and the price range obtained would be too wide [59]. Rajkhowa et al. [25] quantified the dominant factors in the loss of catalytic activity for a Cu-based catalyst during the glycerol hydrogenolysis reaction, concluding that for this copper catalyst, sintering occurs due to the agglomeration of Cu atoms, decreasing the effective area of the catalytic surface. The authors evaluated the stability of the catalyst for periods of 80e90 h under maximum operability conditions and different raw material compositions. For a feed of pure glycerol, the catalyst showed high stability throughout the experiments. However, for a feed stream containing pollutants such as sulfur, chlorine, and glycerides from biodiesel production, its activity decreases. This is one more reason to consider bidistilled glycerol as the raw material. Thus, the loss of catalytic activity was neglected. In the hydrogen production process, Levalley et al. [61] indicate that copper and iron catalysts for the shift reaction keep a good performance from 2 to 4 years. An average time of 3 years was assumed for the economic analysis. Berman and Epstein [62] proved that RueAl2O3 catalysts suffer significant thermal decom- position above 1373 K. As the reforming reactor operates at 973 K, it was considered that thermal deactivation is not critical, and the catalyst may be replaced together with the other catalysts. The annual consumption of raw materials and utilities was ob- tained directly from Aspen HYSYS®. For the calculation of manpower, the Brazilian average wage for industrial operators was considered [63]. The number of workers per shift was estimated from the number of equipment units, considering five shifts as recommended by Seider et al. [33]. The remaining production costs were estimated from correlations available in the literature. 3. Results and discussion A brief description of the process equipment size and cost can be found in Table 6, while the relative costs of equipment pieces are shown in Fig. 4. To produce 821 kg/h of propylene glycol, 22 kg/h of hydrogen, 1070 kg/h of glycerol, and 267 kg/h of water are needed. Producing 23 kg/h of hydrogen requires 174 kg/h of glycerol and 127 kg/h of water. It is clear that the distillation columns are the most expensive equipment in the propylene glycol process. This is due to the high purity specification of PG, which leads to larger distillation columns. Due to the high steam demand, the steam generation unit costs represent more than half of the equipment costs for the steam reforming process, as can be seen in Fig. 5. Figs. 6 and 7 show the costs distribution of the utilities consumed in each process. Cooling water and steam are the main utilities used in the PG process because of the distillation columns’ Fig. 3. Glycerol steam reforming flowsheet. Table 4 Prices of raw materials, utilities and products involved in the glycerol hydro- genolysis and glycerol steam reforming processes. Raw Material Unit Price (US$/kg) Reference Glycerol 0.4918 [64] Hydrogen 1.28e4.14 [58] Process Water 0.0008 [65] Utilities 450 psi Vapor 0.0130 [66] a Cooling Water 0.0008 [65] Fuel (US$/m3 ) 0.9768 [67] Electricity (US$/kWh) 0.1032 [68] Effluent Treatment 0.0560 [69] Catalysts Cu= ZnO= Al2O3 1.8000 [70] Ru= Al2O3 21000 [70] Fe2O3 13.900 [70] CuO 6.7000 [70] Products Propylene Glycol 1.4500 [10] a Updated to 2018 with the Plant Cost Index. R.X. Jimenez et al. / Renewable Energy 158 (2020) 181e191186
  • 7. reboilers and condensers. Moreover, there is also a high electricity demand due to the compressor. Considering the production of partially renewable propylene glycol (with hydrogen purchased from an external source) with an average price of 2.185 US$/kg for H2, the break-even price for pro- pylene glycol is 1.17 US$/kg, with a total investment of 4,722,678.07 US$. The break-even price of hydrogen production via glycerol steam reforming is 9.01 US$/kg, with a total investment of 1,441,653.78 US$. The detailed cost statement of the processes is presented as Supporting Information, Table S1. It was found that the break-even price of propylene glycol is 1.36 US$/kg for a fully renewable PG production process (that is, a propylene glycol production process integrated with a glycerol reforming process to produce hydrogen). Therefore, the generation of a 100% green PG resulted in a 16% increase in its final price. Table 5 Prices of hydrogen from different sources in 2003 and 2018 [59]. Process Energy/Raw Material Source 2003 Price (US$/kg) 2018 Price (US$/kg) Water Electrolysis Nuclear Energy 1.98 3.62 Hydraulic Energy 0.79 1.45 Natural Gas 1.58 2.90 Solar 20.76 38.10 Fossil Fuels Reforming Natural Gas 0.12 0.23 Gasoline 1.35 2.48 Methanol 1.19 2.19 Renewable Fuels Reforming Biogas 1.71 3.15 Ethanol 1.50 2.74 Table 6 Propylene glycol and glycerol reforming processes equipment sizing and cost estimates. Equipment Equipment Size Size Variable Cost (US$) Propylene Glycol Process Heat Exchanger 58.81 Total Heat Transfer Area (m2 ) 59,733.88 Reactor 1.20 Volume (m3 ) 26,967.90 Distillation Colum 1 18 Theoretical Stages 16,450.77 Distillation Colum 2 60 Theoretical Stages 137,091.41 Flash Vessel 1.60 Height (m) 405,172.96 Glycerol Steam Reforming Process Vapor Generation Unit e e 91,355.20 Heat Exchanger 1 2.86 Total Heat Transfer Area (m2 ) 4403.80 Heat Exchanger 1 2.86 Total Heat Transfer Area (m2 ) 4140.31 Reforming Reactor 0.0119 Volume (m3 ) 7908.29 HTS Reactor 0.0055 Volume (m3 ) 10,604.89 LTS Reactor 0.0052 Volume (m3 ) 9592.25 Flash Vessel 1.60 Volume (m3 ) 7908.29 PSA 0.1435 Volume (m3 ) 13,404.60 Fig. 4. Propylene glycol process equipment cost distribution. Fig. 5. Steam reforming process equipment cost distribution. R.X. Jimenez et al. / Renewable Energy 158 (2020) 181e191 187
  • 8. However, this additional expense can be offset when the expanding market shares of more sustainable products are taken into account. The costs distributions of the evaluated processes are shown in Table 7. It can be seen that raw material and catalysts correspond to 55e60% of the total production cost. The renewable PG production cost is higher in nearly US$ 1.2 million, due to the price of renew- able hydrogen. The main source of uncertainty in the economic analysis for the propylene glycol production process considering that the hydrogen is purchased from an external source is its purchase price, which depends on the raw material used to produce hydrogen, the quantity purchased and the distance between the PG plant and hydrogen plant. For this reason, a sensitivity analysis was performed to assess the impact of the price of H2 on the PG production cost. It was found that a 1% increase in the price of hydrogen implies a 0.045% increase in the PG production cost. It follows that the price of propylene glycol is not greatly influenced by small variations in the price of hydrogen. Considering the limits of the hydrogen price range [59], a price range for propylene glycol of 1.15e1.23 US$/kg is obtained. Thus, it can be said that even for large variations in the price of hydrogen, the propylene glycol production cost is not significantly affected. If a similar sensitivity analysis concerning the glycerol price is performed, it is concluded that for a 1% increase in the price of glycerol the glycerol reforming process cost increases by 0.41% and the propylene glycol production process cost increases by 0.57%. The selling price of PG and the quantity sold is subject to changes due to market offer and demand. Fig. 8 shows the variation of annual net profit with the variation of the PG price. This analysis was performed for a constant propylene glycol production of 824 kg/h, which was the amount obtained by processing all the glycerol generated by an average biodiesel plant in Brazil. A similar analysis is shown in Fig. 9, but in this case the amount of PG produced is varied and its price is kept constant. For the renewable propylene glycol, the minimum quantity to be produced to achieve a zero annual net profit is approximately 776 kg/h. For the partially renewable propylene glycol with hydrogen obtained Fig. 6. Propylene glycol process utilities cost distribution. Fig. 7. Steam reforming process utilities cost distribution. Table 7 Production costs distribution for the partially renewable PG and renewable PG processes. Production Costs Distribution Partially Renewable PG Renewable PG Raw Materials and Catalyst 59.7% 55.7% Direct Costs 7.0% 10.7% Indirect Costs 4.7% 6.8% Utilities 18.7% 16.9% Devaluation 5.0% 5.6% Other Expenses 4.9% 4.3% Fig. 8. Variation in the net annual profit as a function of PG selling price. R.X. Jimenez et al. / Renewable Energy 158 (2020) 181e191188
  • 9. from an external source (and a lower production cost), the mini- mum quantity is 670 kg/h. So, it takes 106 kg/h more PG to make green propylene glycol feasible, which represents a relatively broader market share. A comparison between the conventional process, which uses propylene oxide as the main raw material, and the renewable and partially renewable processes was carried out. Raw material and utility consumption for the conventional process were estimated by simplified mass and energy balances in Aspen HYSYS®, as shown in Fig. 1. The stream properties are presented as Supporting Infor- mation, Table S5. In this technology, the reaction consists of pro- pylene oxide hydration in the presence of an acid catalyst (sulfuric acid), and methanol is used as a solvent. The reaction is normally carried out in a continuous stirred tank [71]. There are some clear disadvantages in the conventional process, such as petroleum-based raw materials, toxic solvents and corro- sion problems. Propylene oxide is commonly obtained from pro- pylene (petroleum based), while methanol is produced from natural gas. Corrosion problems must be taken in consideration, because of the use of H2SO4, which is dissolved in water in 0,1% wt. To produce the same amount of PG, 4000 kg/h of water, 40 kg/h of H2SO4, 643 kg/h of methanol and 675 kg/h of propylene oxide are needed. In comparison, the partially green process uses only 267 kg/ h of water and the fully green process consumes a total of 394 kg/h of process water. Both processes require none of the other com- ponents. Aspen HYSYS® also gives an estimation of the carbon emissions. Its 2237 kg/h for the conventional process, against 467 kg/h for the propylene glycol production stage and 141 kg/h for steam reforming. Fig. 10 (A) shows that the conventional process’s raw material cost is greater in nearly 25% than the partially renewable PG process and 16% than the renewable PG process. This difference is caused by the high demand of solvent, which is used in a propylene oxide- methanol equivolume mixture. In Fig. 10 (B) the utilities demand of each process can be compared. The propylene oxide process utilities consumption is almost 3.5 times larger than in the renewable PG process. This high consumption is due to the requirement to keep the reactor temperature under 325 K and the separation steps needed to achieve the same purity that has been reached in the other processes. The reactor temperature is a critical variable, because of propylene oxide’s low boiling point. Addi- tionally, it is important to notice that the renewable PG process consumes more utilities than the partially renewable PG process, due to steam reforming. Finally, these results suggest that both green and partially green PG are economically feasible, because both break-even prices of propylene glycol were lower than the reported market price, even though its price depends on the existing demand. The environ- mental advantages of these products seem clear, but, for future work, it would be interesting to perform a formal environmental analysis of the green propylene glycol production process, such as a Life Cycle Assessment (LCA), and to compare it to the conventional PG production process to have more elucidative results in this sense. 4. Conclusion The present work simulated the propylene glycol production Fig. 9. Variation in the net annual profit as a function of PG production. Fig. 10. Comparison of the PG production processes. (A) Raw material costs. (B) Util- ities demands. R.X. Jimenez et al. / Renewable Energy 158 (2020) 181e191 189
  • 10. process using glycerol from biodiesel production as the main raw material, in a Brazilian scenario. A total production cost of 1.17 US$/ kg was obtained when PG is produced using hydrogen from an external source. When it is produced using hydrogen obtained from glycerol steam reforming, its price increases to 1.36 US$/kg. A hydrogen production cost of 9.01 US$/kg was calculated when it is obtained from glycerol steam reforming, which is much higher when compared to the market prices of hydrogen obtained by other processes. It has caused a nearly 70% drop in the annual net profit of propylene glycol production in the base scenario, although the process remains economically viable. Fully renewable PG is approximately 16% more expensive than partially renewable PG, but it is possible, depending on the market. Declaration of competing interest The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper. CRediT authorship contribution statement Roberto X. Jimenez: Conceptualization, Investigation, Software, Writing - original draft, Writing - review editing. Andre F. Young: Conceptualization, Methodology, Supervision, Visualization, Writing - original draft, Writing - review editing. Heloisa L.S. Fernandes: Conceptualization, Supervision, Writing - original draft, Writing - review editing. Appendix A. Supplementary data Supplementary data to this article can be found online at https://doi.org/10.1016/j.renene.2020.05.126. References [1] A.F. Young, F.L.P. Pessoa, E.M. Queiroz, Biodiesel Production Technologies - Supercritical and Enzymatic Production, Novas Ediç~oes Acad^emicas, Saar- brücken, Germany, 2015 (in Portuguese). [2] M.C.S. de Mello, H.G.D. Villardi, A.F. Young, F.L.P. Pessoa, A.M. Salgado, Life cycle assessment of biodiesel produced by the methylic-alkaline and ethylic- enzymatic routes, Fuel 208 (2017) 329e336, https://doi.org/10.1016/ j.fuel.2017.07.014. [3] M.R. Monteiro, C.L. Kugelmeier, R.S. Pinheiro, M.O. Batalha, A. da Silva Cesar, Glycerol from biodiesel production: technological paths for sustainability, Renew. Sustain. Energy Rev. 88 (2018) 109e122, https://doi.org/10.1016/ j.rser.2018.02.019. [4] ANP - Brazilian National Agency of Petroleum, Natural Gas and Biofuels. Biodisel, 2018 accessed 10.1.19, http://www.anp.gov.br/producao-de- biocombustiveis/biodiesel/simp-biodisel (in Portuguese). [5] C.J.A. Mota, Pinto, Catalytic transformations of glycerol for innovation in the chemical industry, Rev. Virtual Quim. 9 (2017) 135e149, https://doi.org/ 10.21577/1984-6835.20170011. [6] Z. Freitas, Glycerin as raw material for chemical industry: Glicerina como materia-prima para indústria química: evaluation of research efforts and commercial initiatives, Federal University of Rio de Janeiro, Rio de Janeiro, Brazil, 2013 (in Portuguese). [7] S. Veluturla, N. Archna, D. Subba Rao, N. Hezil, I.S. Indraja, S. Spoorthi, Catalytic valorization of raw glycerol derived from biodiesel: a review, Biofuels 9 (2018) 305e314, https://doi.org/10.1080/17597269.2016.1266234. [8] H. Mitta, P.K. Seelam, S. Ojala, R.L. Keiski, P. Balla, Tuning Y-zeolite based catalyst with copper for enhanced activity and selectivity in vapor phase hydrogenolysis of glycerol to 1,2-propanediol, Appl. Catal. Gen. 550 (2018) 308e319, https://doi.org/10.1016/j.apcata.2017.10.019. [9] Y. Nakagawa, K. Tomishige, Heterogeneous catalysis of the glycerol hydro- genolysis, Catal. Sci. Technol. 1 (2011) 179, https://doi.org/10.1039/ c0cy00054j. [10] Abiquim - Brazilian Chemical Industry Association, Annual Report, 2018 (in Portuguse). [11] Y. Zhang, M.A. Dube, D.D. McLean, M. Kates, Biodiesel production from waste cooking oil: 1. Process design and technological assessment, Bioresour. Technol. 89 (2003) 1e16, https://doi.org/10.1016/S0960-8524(03)00040-3. [12] Z. Zhou, X. Li, T. Zeng, W. Hong, Z. Cheng, W. Yuan, Kinetics of hydrogenolysis of glycerol to propylene glycol over Cu-ZnO-Al2O3 catalysts, Chin. J. Chem. Eng. 18 (2010) 384e390, https://doi.org/10.1016/S1004-9541(10)60235-2. [13] M.A. Dasari, P.-P. Kiatsimkul, W.R. Sutterlin, G.J. Suppes, Low-pressure hydrogenolysis of glycerol to propylene glycol, Appl. Catal. Gen. 281 (2005) 225e231, https://doi.org/10.1016/J.APCATA.2004.11.033. [14] M.N. Gatti, M.D. Mizrahi, J.M. Ramallo-Lopez, F. Pompeo, G.F. Santori, N.N. Nichio, Improvement of the catalytic activity of Ni/SiO2-C by the modi- fication of the support and Zn addition: bio-propylene glycol from glycerol, Appl. Catal. Gen. 548 (2017) 24e32, https://doi.org/10.1016/ J.APCATA.2017.08.037. [15] E.P. Maris, R.J. Davis, Hydrogenolysis of glycerol over carbon-supported Ru and Pt catalysts, J. Catal. 249 (2007) 328e337, https://doi.org/10.1016/ J.JCAT.2007.05.008. [16] T. Miyazawa, Y. Kusunoki, K. Kunimori, K. Tomishige, Glycerol conversion in the aqueous solution under hydrogen over Ru/C þ an ion-exchange resin and its reaction mechanism, J. Catal. 240 (2006) 213e221, https://doi.org/10.1016/ J.JCAT.2006.03.023. [17] C. Montassier, J.C. Menezo, L.C. Hoang, C. Renaud, J. Barbier, Aqueous polyol conversions on ruthenium and on sulfur-modified ruthenium, J. Mol. Catal. 70 (1991) 99e110, https://doi.org/10.1016/0304-5102(91)85008-P. [18] S.R. Schmidt, S.K. Tanielyan, N. Marin, G. Alvez, R.L. Augustine, Selective conversion of glycerol to propylene glycol over fixed bed Raney® Cu catalysts, Top. Catal. 53 (2010) 1214e1216, https://doi.org/10.1007/s11244-010-9565- x. [19] R.V. Sharma, P. Kumar, A.K. Dalai, Selective hydrogenolysis of glycerol to propylene glycol by using Cu:Zn:Cr:Zr mixed metal oxides catalyst, Appl. Catal. Gen. 477 (2014) 147e156, https://doi.org/10.1016/ J.APCATA.2014.03.007. [20] E.S. Vasiliadou, A.A. Lemonidou, Glycerol transformation to value added C3 diols: reaction mechanism, kinetic, and engineering aspects, Wiley Inter- discip. Rev. Energy Environ. 4 (2015) 486e520, https://doi.org/10.1002/ wene.159. [21] G.J. Suppes, W.R. Sutterlin, M. Dasari, Method of Producing Lower Alcohols from Glycerol, 2007. US7943805B2. [22] C. Montassier, D. Giraud, J. Barbier, Polyol conversion by liquid phase het- erogeneous catalysis over metals, Stud. Surf. Sci. Catal. 41 (1988) 165e170, https://doi.org/10.1016/S0167-2991(09)60811-9. [23] S. Wang, Y. Zhang, H. Liu, Selective hydrogenolysis of glycerol to propylene glycol on Cu-ZnO composite catalysts: structural requirements and reaction mechanism, Chem. Asian J. 5 (2010) 1100e1111, https://doi.org/10.1002/ asia.200900668. [24] D. Lahr, B. Shanks, Kinetic analysis of the hydrogenolysis of lower polyhydric Alcohols: glycerol to glycols, Ind. Eng. Chem. Res. 42 (2003) 5467e5472, https://doi.org/10.1021/IE030468L. [25] T. Rajkhowa, G.B. Marin, J.W. Thybaut, A comprehensive kinetic model for Cu catalyzed liquid phase glycerol hydrogenolysis, Appl. Catal. B Environ. 205 (2017) 469e480, https://doi.org/10.1016/J.APCATB.2016.12.042. [26] C.K. Cheng, S.Y. Foo, A.A. Adesina, Glycerol steam reforming over bimetallic Co-Ni/Al2O3, Ind. Eng. Chem. Res. 49 (2010) 10804e10817, https://doi.org/ 10.1021/ie100462t. [27] A. Lancia, D. Musmarra, F. Pepe, Vapor-liquid equilibria for mixtures of ethylene glycol, propylene glycol, and water between 98º and 122ºC, J. Chem. Eng. Jpn. 29 (1996) 449e455, https://doi.org/10.1252/jcej.29.449. [28] E.S. Vasiliadou, A.A. Lemonidou, Kinetic study of liquid-phase glycerol hydrogenolysis over Cu/SiO2 catalyst, Chem. Eng. J. 231 (2013) 103e112, https://doi.org/10.1016/j.cej.2013.06.096. [29] Y. Xi, J.E. Holladay, J.G. Frye, A.A. Oberg, J.E. Jackson, D.J. Miller, A kinetic and mass transfer model for glycerol hydrogenolysis in a trickle-bed reactor, Org. Process Res. Dev. 14 (2010) 1304e1312, https://doi.org/10.1021/op900336a. [30] R. Perry, D. Green, Perry’s Chemical Engineers’ Handbook, McGraw-Hil, New York, United States, 1997. [31] C.N. Satterfield, Trickle-bed reactors, AIChE J. 21 (1975) 209e228, https:// doi.org/10.1002/aic.690210202. [32] R.J. Farrauto, C.H. Bartholomew, Fundamentals of Industrial Catalytic Pro- cesses, Blackie Academic Professional, London, United Kingdom, 1997. [33] W.D. Seider, J.D. Seader, D.R. Lewin, Product and Process Design Principles - Synthesis, Analysis, and Evaluation, Second, John Wiley and Sons, Inc, Inter- national Edition, 2003. [34] D. Woods, Rules of Thumb in Engineering Practice, Wiley, Darmstadt, Ger- many, 2007. [35] R. Sinnott, G. Towler, Chemical Engineering Design, Elsevier Inc., United States, 2008. [36] C. Ferreira, Process Flowsheet Proposal for Purification of Glycerol from Bio- diesel Production, Federal University of Rio de Janeiro, Rio de Janeiro, Brazil, 2018 (in Portuguese). [37] I. Gandarias, P.L. Arias, I. Agirrezabal-Telleria, Economic assessment for the production of 1,2-Propanediol from bioglycerol hydrogenolysis using molec- ular hydrogen or hydrogen donor molecules, Environ. Prog. Sustain. Energy 35 (2016) 447e454, https://doi.org/10.1002/ep.12232. [38] R.R. Davda, J.W. Shabaker, G.W. Huber, R.D. Cortright, J.A. Dumesic, A review of catalytic issues and process conditions for renewable hydrogen and alkanes by aqueous-phase reforming of oxygenated hydrocarbons over supported metal catalysts, Appl. Catal. B Environ. 56 (2005) 171e186, https://doi.org/ 10.1016/J.APCATB.2004.04.027. [39] A. Iriondo, V.L. Barrio, J.F. Cambra, P.L. Arias, M.B. Güemez, R.M. Navarro, M.C. Sanchez-Sanchez, J.L.G. Fierro, Hydrogen production from glycerol over R.X. Jimenez et al. / Renewable Energy 158 (2020) 181e191190
  • 11. nickel catalysts supported on Al2O3 modified by Mg, Zr, Ce or La, Top. Catal. 49 (2008) 46e58, https://doi.org/10.1007/s11244-008-9060-9. [40] P.D. Vaidya, A.E. Rodrigues, Glycerol reforming for hydrogen production: a review, Chem. Eng. Technol. 32 (2009) 1463e1469, https://doi.org/10.1002/ ceat.200900120. [41] R.L. Manfro, N.F.P. Ribeiro, M.M.V.M. Souza, Production of hydrogen from steam reforming of glycerol using nickel catalysts supported on Al2O3, CeO2 and ZrO2, Catal. Sustain. Energy. 1 (2013) 60e70, https://doi.org/10.2478/cse- 2013-0001. [42] I. Iliuta, H.R. Radfarnia, M.C. Iliuta, Hydrogen production by sorption- enhanced steam glycerol reforming: sorption kinetics and reactor simula- tion, AIChE J. 59 (2013) 2105e2118, https://doi.org/10.1002/aic.13979. [43] R.L. Manfro, M.M.V.M. Souza, Catalyst for hydrogen production through glycerol reforming review, Curr. Top. Catal. 11 (2014) 37e54. [44] S. Adhikari, S. Fernando, S.R. Gwaltney, S.D. Filip To, R. Mark Bricka, P.H. Steele, A. Haryanto, A thermodynamic analysis of hydrogen production by steam reforming of glycerol, Int. J. Hydrogen Energy 32 (2007) 2875e2880, https://doi.org/10.1016/J.IJHYDENE.2007.03.023. [45] B. Zhang, X. Tang, Y. Li, Y. Xu, W. Shen, Hydrogen production from steam reforming of ethanol and glycerol over ceria-supported metal catalysts, Int. J. Hydrogen Energy 32 (2007) 2367e2373, https://doi.org/10.1016/ J.IJHYDENE.2006.11.003. [46] A.O. Menezes, M.T. Rodrigues, A. Zimmaro, L.E.P. Borges, M.A. Fraga, Produc- tion of renewable hydrogen from aqueous-phase reforming of glycerol over Pt catalysts supported on different oxides, Renew. Energy 36 (2011) 595e599, https://doi.org/10.1016/J.RENENE.2010.08.004. [47] K.S. Avasthi, R.N. Reddy, S. Patel, Challenges in the production of hydrogen from glycerol e a biodiesel byproduct via steam reforming process, Procedia Eng. 51 (2013) 423e429, https://doi.org/10.1016/J.PROENG.2013.01.059. [48] S. Adhikari, S.D. Fernando, A. Haryanto, Kinetics and reactor modeling of hydrogen production from glycerol via steam reforming process over Ni/CeO2 catalysts, Chem. Eng. Technol. 32 (2009) 541e547, https://doi.org/10.1002/ ceat.200800462. [49] R. Sundari, P.D. Vaidya, Reaction kinetics of glycerol steam reforming using a Ru/Al2O3 catalyst, Energy Fuels 26 (2012) 4195e4204, https://doi.org/ 10.1021/ef300658n. [50] H. Chen, Y. Ding, N.T. Cong, B. Dou, V. Dupont, M. Ghadiri, P.T. Williams, A comparative study on hydrogen production from steam-glycerol reforming: thermodynamics and experimental, Renew. Energy 36 (2011) 779e788, https://doi.org/10.1016/j.renene.2010.07.026. [51] Y. Liu, R. Farruto, A. Lawal, Autothermal reforming of glycerol in a dual layer monolith catalyst, Chem. Eng. Sci. 89 (2013) 31e39, https://doi.org/10.1016/ J.CES.2012.11.030. [52] J. Villaça, Proposal and Simulation of the Glycerol Steam Reforming Process for Hydrogen Obtaining, Federal University of Rio de Janeiro, Rio de Janeiro, Brazil, 2018 (in Portuguese). [53] X. Wang, S. Li, H. Wang, B. Liu, X. Ma, Thermodynamic analysis of glycerin steam reforming, Energy Fuels 22 (2008) 4285e4291, https://doi.org/10.1021/ ef800487r. [54] G. Yang, H. Yu, F. Peng, H. Wang, J. Yang, D. Xie, Thermodynamic analysis of hydrogen generation via oxidative steam reforming of glycerol, Renew. En- ergy 36 (2011) 2120e2127, https://doi.org/10.1016/J.RENENE.2011.01.022. [55] M.M.V.M. Souza, Hydrogen Technology, Synergia, Rio de Janeiro, Brazil, 2009 (in Portuguese). [56] A. Hines, R. Maddox, Mass Transfer Fundamentals and Applications, Prentice- Hall, Michigan, United States, 1985. [57] S. Jenkins, Economic Indicators, Chem. Eng, 2019. https://www. chemengonline.com/cepci-updates-january-2018-prelim-and-december- 2017-final/?printmode¼1. accessed 10.3.18. [58] L. Bonfim-Rocha, M.L. Gimenes, S.H. Bernardo de Faria, R.O. Silva, L.J. Esteller, Multi-objective design of a new sustainable scenario for bio-methanol pro- duction in Brazil, J. Clean. Prod. 187 (2018) 1043e1056, https://doi.org/ 10.1016/J.JCLEPRO.2018.03.267. [59] M.L.M. Negro, R.C. Garner, M. Linardi, Mass production of hydrogen in Brazil, world clim, Energy Event 3 (2003) 293e298. [60] IBGE - Brazilian Institute of Geography and Statistics, National Consumer Price Index, 2019. https://www.ibge.gov.br/estatisticas/economicas/precos-e- custos/9256-indice-nacional-de-precos-ao-consumidor-amplo.html? =t=series-historicas. accessed 9.29.18. [61] T.L. LeValley, A.R. Richard, M. Fan, The progress in water gas shift and steam reforming hydrogen production technologies e a review, Int. J. Hydrogen Energy 39 (2014) 16983e17000, https://doi.org/10.1016/ j.ijhydene.2014.08.041. [62] A. Berman, M. Epstein, Ruthenium catalysts for high temperature solar reforming of methane, in: Hydrogen Power: Theoretical and Engineering Solutions, Springer, Dordrecht, Netherlands, 1998, pp. 213e218, https:// doi.org/10.1007/978-94-015-9054-9_26. [63] Glassdoor, Industrial Operator Salaries, 2018 accessed 9.20.18, https://www. lovemondays.com.br/salarios/cargo/salario-operador-industrial (in Portuguese). [64] Comex Stat, 2018. http://comexstat.mdic.gov.br/pt/home. accessed 9.20.18. [65] SNIS - National System of Information about Sanitation, Diagnosis of the Water and Sewage Services - 2017, Brasilia, Brazil, 2019. [66] C.A.G. Perlingeiro, Process Systems Engineering, Blucher, S~ao Paulo, Brazil, 2005 (in Portuguese). [67] Comgas - S~ao Paulo Gas Company, Pipe Natural Gas Tariffs, 2018 accessed 10.1.18, https://www.comgas.com.br/tarifas/comercial/ (in Portuguse). [68] ANEEL - National Electricity Regulatory Agency, Energy consumption report by sector, accessed 9.20.18, http://relatorios.aneel.gov.br/_layouts/xlviewer. aspx?id=/RelatoriosSAS/RelSampRegCC.xlsxSource=http://relatorios.aneel. gov.br/RelatoriosSAS/Forms/AllItems.aspxDefaultItemOpen=1, 2018 (in Portuguese). [69] R. Turton, R. Bailie, W. Whiting, J. Shaeiwitz, Analysis, Synthesis and Design of Chemical Processes, Prentice Hall, Boston, United States, 2008. [70] Sigma-Aldrich, Product Directory, 2018. https://www.sigmaaldrich.com/ technical-service-home/product-catalog.html. accessed 10.1.18. [71] H.S. Fogler, Elements of Chemical Engineering, fourth ed., Prentice Hall, New Jersey, United States, 2009. R.X. Jimenez et al. / Renewable Energy 158 (2020) 181e191 191