Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Renewable Energy 158 (2020) 181e191

Contents lists available at ScienceDirect

Renewable Energy
journal homepage: www.elsevier.com/locate/renene

Propylene glycol from glycerol: Process evaluation and break-even


price determination
 F. Young b, *, Heloisa L.S. Fernandes a
nez a, Andre
Roberto X. Jime
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 Pa tria, 156, Nitero
i, Brazil

a r t i c l e i n f o a b s t r a c t

Article history: Propylene glycol can be produced from the hydrogenolysis of glycerol. If hydrogen comes from a
Received 22 November 2019 renewable source, the produced propylene glycol can be considered completely renewable. The objective
Received in revised form of this work is to perform a technical and economic evaluation of a propylene glycol production process
27 April 2020
which uses glycerol as the main raw material. Two possibilities regarding the hydrogen source are
Accepted 23 May 2020
Available online 31 May 2020
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
Keywords:
Glycerol
produced via steam reforming is not competitive with its price if produced by other sources. Conse-
Process simulation quently, the fully renewable propylene glycol production process yields a lower profit, but it is still viable
Propylene glycol and depends on the market.
Hydrogen © 2020 Elsevier Ltd. All rights reserved.
Reforming

1. Introduction number of patents or patent applications on PG production from


glycerol was larger than those on 1,3-propanediol or hydrogen
Glycerol is the main by-product of biodiesel production [1e3]. production until 2013.
Along with the increased supply of biodiesel, glycerol production Propylene glycol can be used as a surfactant, a moisturizer, an
has also increased. In 2017, there were produced about 374.5 antifreeze agent, a solvent, a preservative and in detergent for-
thousand cubic meters of glycerol, which meant a 9.5% increase in mulations [6,8]. It is usually obtained by the hydrolysis of propylene
comparison with 2016 [4]. oxide [8,9]. Propylene oxide is produced from propylene hydro-
The increasing glycerol surplus has motivated the development chloride, which comes from the petrochemical industry.
of chemical processes which employ it as the main raw material. As In Brazil, the only registered producer of propylene glycol is
a consequence, the number of published scientific articles on the Dow®. It has a chemical plant located in the industrial complex of
use of glycerol as the main raw material in chemical processes has Aratu, in the state of Bahia. The company also produces propylene
increased. For example, articles related to glycerol in the Scopus oxide, so propylene glycol is still produced by the conventional
database have increased by almost 7 times per year between 2008 route from petrochemicals. It is noteworthy that this is the only
and 2016 [3]. propylene glycol production plant in Latin America [10]. A simpli-
Among the products that can be obtained from glycerol there fied flowsheet of this process technology is shown in Fig. 1.
are propanediols (1,2- and 1,3-propanediol), propene, epichloro- Alternatively, PG can be produced by a more ecologically sound
hydrin, methanol, ethanol, lactic acid, citric acid, propanoic acid route, which employs glycerol as a carbon and hydrogen source,
and hydrogen [5e7]. Monteiro [3] and Freitas [6] identified that 1,3- and which relies less on fossil sources. Some scientific articles about
propanediol, hydrogen and 1,2-propanediol or propylene glycol this route were published, but there are only a few patents (granted
(PG) are the products formed from glycerol on which researchers or under application) on it. Currently, if USPTO and Lens databases
have published more articles. According to these authors, the 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
* Corresponding author.
vapor-phase reactor and one with a two-phase reactor (reactions
E-mail address: ayoung@id.uff.br (A.F. Young).

https://doi.org/10.1016/j.renene.2020.05.126
0960-1481/© 2020 Elsevier Ltd. All rights reserved.
182 R.X. Jimenez et al. / Renewable Energy 158 (2020) 181e191

Fig. 1. Simplified propylene oxide hydration process flowsheet.

start in the liquid phase and generate acetol, which is transferred to Table 1
the vapor phase to react). It is interesting to note that, although the 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
search was not limited by date, all the results date to the last 8 type 2 active site [12].
years.
These results indicate that this technology is new and may Stage Kinetic Adsorption/Desorption

require more studies until consolidation. Considering this, the H2 þ 2Q1 #2H Q1 e bH
objective of this work is to perform a technical and economic G þ Q2 #GQ2 e bG
GQ2 /AQ2 þ H2 O k1 e
assessment of the partially renewable and fully renewable PG
2HQ1 þ AQ2 /PGQ2 þ 2Q1 k2 e
production processes from glycerol, to find out if these processes AQ2 #A þ Q2 e b1
A
are economically viable in the current Brazilian economic scenario. PGQ2 #P þ Q2 e b1
PG
Equipment design and process simulations were performed with
the aid of the Aspen HYSYS® v8.8 process simulator.

2. Process simulation and equipment design propylene glycol from glycerol: according to the dehydration-
hydrogenation mechanism, glycerol is initially dehydrated at
In order to proceed to the economic analysis, the reaction ki- catalyst acid sites to acetol, which is hydrogenated to 1,2-
netics and vapor-liquid equilibria must be correctly reproduced in propanediol [5,13,21]. A three-phase dehydrogenation-dehydra-
the simulations so that the equipment can be properly designed. tion-hydrogenation mechanism is also suggested [15,17,22,23].
The properties of all components involved in the system are The literature recommends a Langmuir-Hinshelwood-type
available in the Aspen HYSYS® simulator database. 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
2.1. Propylene glycol production from glycerol
dehydration-hydrogenation mechanism [12,19,24,25]. Two types of
catalyst active sites are considered, as observed in Table 1 [12,26].
Glycerol (or glycerin) obtained from biodiesel is available in
Zhou et al. [12] employed a CueZnOeAl2O3 catalyst with a
different purity levels. Crude glycerin is defined as the glycerol-rich
molar ratio of 1:1:0.5 of Cu/Zn/Al, with a particle diameter of
mixture (around 75% glycerol) obtained after the separation of
0.17 mm (80e100 mesh), and reported a conversion of 81.5% of the
biodiesel. After neutralization, glycerin is usually called blonde; its
glycerol and selectivity of 93.4% in terms of propylene glycol. Also,
weight percent composition is 85.0% glycerol, 11.7% water, 3.2%
the authors proposed the rate expressions of the two reaction steps
methanol, and 0.2% biodiesel. After distillation, USP grade glycerol
presented in Table 1, which can be found in Eqs. (1) and (2).
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, k1 bG CG
r1 ¼ (1)
bidistilled glycerol is regarded as the process raw material in this ð1 þ bG CG þ bA CA þ bP CP Þ
work, since kinetic data for the hydrogenolysis and reforming re-
actions refer to glycerol-water mixtures only [12].
k2 bG CG
Furthermore, according to Resolution no. 30 published by ANP r2 ¼  pffiffiffiffiffiffiffiffiffiffiffiffi 2 (2)
(Brazilian National Agency of Petroleum, Natural Gas and Biofuels), ð1 þ bG CG þ bA CA þ bP CP Þ 1 þ bH PH
which regulates biodiesel production activities in Brazil [4],
including the construction stage, the glycerol purification stage In Eqs. (1) and (2), Ci is the molar concentration of component i,
must be present in every biodiesel production unit. Thus, in all PH is the partial pressure of hydrogen, ki is the rate constant of
plants currently operating in Brazil, a stream of composition close reaction i and bi is the absorption rate constant of component i [12].
to that of bidistilled glycerol is available, or at least its composition The numerical values of these constants are described in Table 2. A
can be adjusted for this purpose. preliminary simulation of the reaction in laboratory conditions was
In literature, a wide variety of metal catalysts has been used to carried out in Aspen HYSYS®; simulated and experimental data
promote the CeO bond cleavage reaction of the glycerol molecule. [12] were compared and relative errors of 3.2% in the glycerol
Cu-based catalysts are widely studied as they offer a high degree of conversion and 7.0% in the acetol to propylene glycol selectivity
conversion and selectivity of PG [13e20]. ratio were found. Since average relative errors between experi-
Two mechanisms have been proposed for the production of mental and theoretical glycerol and propylene glycol flowrates at
R.X. Jimenez et al. / Renewable Energy 158 (2020) 181e191 183

Table 2
Parameters for the kinetic model of the glycerol hydrogenolysis reaction [12].

Kinetic and absorption constants Parameters


 
  kJ
mol Activation Energy
Pre-exponential Factor mol
kg s

k1 1.54  107 86.56


k2 7.16  106 57.8
bG 2.22 36.42
bA 8.73 25.94
bPG 5.80 25.77
bH 1.86  105 MPa 36.24

the reactor output are 6.3% and 7.6%, respectively [12], deviations exchanges heat with the glycerol stream to minimize utility ex-
obtained in the preliminary simulation can be considered accept- penses. Then it follows to the separation section, in which excess
able. Thus, the rate expressions in Eqs. (1) and (2) are used in this hydrogen is recovered with a flash vessel, recycled and mixed with
work. pure hydrogen, thereby ensuring a 5:1 M ratio of hydrogen to
The UNIQUAC model was used for phase equilibria calculations glycerol at the process inlet. Finally, two distillation columns adjust
for the system containing the components involved in the pro- the propylene glycol purity to the required value (USP grade,
duction of propylene glycol, since it provides a satisfactory 99.5%).
description of phase equilibria for similar systems [27]. The literature suggests the use of a trickle-bed-reactor for
The propylene glycol process flowsheet can be seen in Fig. 2. The glycerol hydrogenolysis in industrial-scale processes [25,28,29].
stream properties are presented as Supporting Information, Inside the reactor, liquid and gas flow downward through the cat-
Table S3. The raw material is fed into the process and is put under alytic bed. The net flowrate must be low enough so that the liquid
operating conditions before entering the reactor. Excess hydrogen phase forms a thin film over the fixed bed and descends by gravity
is used to facilitate mass transfer in the three-phase system within and gas-phase drag [30]. Thus, the contact among the three phases
the reactor. The global reaction is shown in Eq. (3). is maximized. Trickle-bed-reactors have been widely used in other
processes involving hydrogen, mainly in the hydrodesulfurization
of petroleum fractions [30,31].
C3 H8 O3 þ H2 /C3 H8 O2 þ H2 O
(3) The catalyst density was not mentioned in the articles reviewed
Glycerol PG
but is required for the calculation of the fixed-bed volume. The
The reaction is carried out at T ¼ 493 K, PH ¼ 4 MPa and a fixed-bed volume includes the volume occupied by the catalyst
hydrogen to glycerol molar ratio of 5:1. The catalyst is assumed to solid mass (30%) and the volume of pore voids and voids among the
be CueZnOeAl2O3 and the residence time is adjusted to 85 catalyst particles (70%), which are generally in a proportion of 3:3:4
kgCat :h:kg:mol1
Glycerol , in order to reproduce the results obtained by
[32]. The solid density was calculated from the molar fraction of its
Zhou et al. [12]. Moreover, the authors compared their results at components. Assuming the void fraction in the bed to be equal to
493 K with others available in the literature [21], and showed 0.4, the densities of solid, pellets and bulk can be estimated. For the
consistency in the values that were reported. Zhou et al. [12] proved purpose of reactor sizing, the trickle-bed model was simplified to a
that glycerol conversion increases with higher pressures (testing plug-flow fixed bed model, as available in Aspen HYSYS®.
3 MPa, 4 MPa and 5 MPa), but at 493 K it remains almost the same, The required catalyst mass was obtained by solving the design
independent on pressure. A conservative approach was used to equations for a plug-flow reactor. A plug-flow reactor which
choose a 4 MPa pressure for the reaction simulation. The plant was operates at conditions similar to those in the work by Zhou et al.
designed to operate with a flowrate of 1070 kg/h of glycerol. This [12] requires 992.8 kg of catalyst. Likewise, a conversion of 98.2%
was the average glycerol production in biodiesel plants in Brazil in requires 1489 kg of catalyst, while a conversion of 99.99% requires
2018 [4]. an additional 993 kg of catalyst [12]. Consequently, it is not
The reactor effluent is mainly composed by propylene glycol, convenient to try to achieve such high glycerol conversions. Thus,
hydrogen, unreacted glycerol and by-products. The product stream the plant was designed to operate with an overall glycerol

Fig. 2. Propylene glycol production flowsheet.


184 R.X. Jimenez et al. / Renewable Energy 158 (2020) 181e191

conversion of 98.2%. Finally, the reactor volume can be calculated (light key) and ethylene glycol (heavy key). The columns were
by the ratio of catalyst mass and bulk density. simulated rigorously and its input parameters were estimated
The reaction must occur under isothermal conditions. To meet based on typical methods.
this specification, a cooling jacket was applied, with water at 298 K The minimum number of stages was calculated using the Fenske
as coolant. The heat exchange area was assumed to be approxi- equation. The number of theoretical stages was obtained by the
mately equal to the lateral area of the plug-flow reactor. Gilland-Molokanov correlation. The minimum reflux ratio was
In addition to acetol and propylene glycol, the authors identified estimated from the Underwood equation. Finally, the Kirkbride
the presence of by-products such as ethylene glycol, methanol, equation was used to determine the optimum feed stage [35]. It is
ethanol, propanol and propanoic acid in the product stream, among noteworthy that the calculations of these estimates were per-
which ethylene glycol is the most expressive. Therefore, it is formed in the simulator itself, using the Shortcut Distillation block.
necessary to simulate the formation of these products so that the From the obtained values, it was possible to start the rigorous
simulation is in agreement with the experimental study. design of the distillation columns.
Ethylene glycol and methanol modeling was performed by In a first analysis, it was found that the operation under atmo-
specifying the conversion value found experimentally [12]. Due to spheric pressure demands extreme temperatures in condensers
the lack of information on the proportion of the remaining by- and reboilers. As a consequence, common utilities such as steam
products, their simulation was not feasible. Since the by-products and cooling water could not be used. Also, degradation or poly-
are lower alcohols (except for propanoic acid), they were regar- merization reactions of residual glycerol may occur at elevated
ded as a single representative compound (methanol) in the temperatures (about 473 K) [36].
simulations. Thus, to avoid such problems, it was decided that the equipment
In a preliminary analysis of the separation system, it was found should operate under vacuum. The operating pressure was set as
that for a better recovery of unreacted hydrogen it is necessary to close as possible to atmospheric pressure, but the reboilers tem-
cool the product stream from the reactor. To save resources and peratures were kept at least 15 K below the glycerol degradation
accomplish energy integration, it was decided to implement a temperature. The literature suggests the use of packed columns for
shell-and-tube heat exchanger through which this stream and the vacuum distillation and high viscosity mixtures, as in this case [35].
glycerol and water feed stream are passed. The heat transfer rate, Structured packings offer better separation efficiencies and pres-
the heat exchange area and the global heat transfer coefficient were sure drops, but are more expensive. Thus, the columns were
obtained with the aid of the simulator, from the values of the inlet designed with random packings of 1 in Pall metal rings, due to the
and outlet streams’ temperatures, pressures and flowrates. high number of theoretical stages needed [33].
A minimum temperature approach of 10 K was used to avoid an For the calculation of the packed bed height, the Height Equiv-
excessively large heat exchange area. The initial estimate of pres- alent to a Theoretical Plate (HETP) method was used. For vacuum
sure drop on both sides of the exchanger was 10 kPa [33]. The heat distillation, the required column diameter can be estimated from
exchanger is designed with floating heads to minimize thermal the flooding condition. This condition should be avoided and can be
expansion effects that may occur for a temperature difference estimated from the flooding velocity through the Leva correlation
greater than 30 K between the tube fluid (reactor effluent) and the [33]. The results obtained in the simulation are in agreement with
shell fluid [34]. the work of Gandarias et al. [37].
The methodology described by Seider et al. [33], which is based
on the relative volatility of the compounds, was adopted for the 2.2. Hydrogen production: glycerol steam reforming
synthesis of the separation system. The reactor outlet is a two-
phase stream, which consists of hydrogen, methanol, water, ace- Reforming of oxygenated hydrocarbons consists of the cleavage
tol, propylene glycol, ethylene glycol and glycerol (from the most of the CeC bond to form CO and H2 . Then CO and water may react in
volatile to the least volatile). Three cuts were required to achieve a water-gas shift reaction, releasing more H2 and CO2 [38]. Glycerol
the specified purity: Hydrogen/Methanol, Water/Propylene Glycol reforming needs a catalyst that promotes cleavage of the CeC, OeH
and Propylene Glycol/Ethylene Glycol. and CeH bonds, and if possible also promote the shift reaction
Hydrogen is the first component that must be removed, as it is [38,39]. Reforming can be conducted in both liquid and gaseous
the excess reagent and can be recycled into the process, thereby phases. In the latter case, it is necessary to vaporize glycerol, what
reducing raw materials expenditures. Thus, for the separation of requires higher temperatures [40]. The water-gas shift reaction
hydrogen from the reactor effluent stream, a flash vessel was used. occurs only in the gas phase.
The liquid-gas separator should be such that the vessel diameter Glycerol aqueous phase reforming occurs at higher pressures
must be large enough so that the gas velocity is less than the liquid and lower temperatures, if compared to steam reforming, and of-
droplet separation velocity and that the vessel height is equal to its fers greater possibilities for using crude glycerin as a raw material
diameter, or 1 m if the diameter is less than 1 m. It is recommended [41]. The major disadvantage of this route is its low reaction rates
that hold-up time is about 10 min so that the liquid level inside the [42]. Vapor-phase reforming, in turn, can be conducted at atmo-
equipment is appropriate. Finally, the feed inlet must be at least spheric pressures and has presented energetic advantages and
60 cm above the liquid level [35]. higher yield in terms of H2 [43]. That is why this was the route
To maximize the recovery of pure hydrogen, it is necessary that chosen in this work.
the flash vessel also removes heat from the streams, to promote In literature, there are studies about the activity of different
condensation of the less volatile components. Such heat exchange catalysts and operating conditions that maximize the glycerol
was performed by implementing a coil inside the vessel with reforming and shift reactions, to the detriment of methanation and
cooling water at 298 K. Aspen HYSYS® Adjust block was used so that coke formation. Such reactions decrease the efficiency of the pro-
the cooling rate varies to ensure that the purity of the hydrogen cess in terms of H2 generation and impair the catalytic activity by
obtained is at least 99.5%. carbon deposition at the active catalyst sites [44,45].
The next two cuts were made through distillation columns. The Catalysts involving noble metals, such as Pt, Ru and Pd, and non-
first column cuts between water (light key) and propylene glycol noble metals, such as Ni and Co, have been successfully tested in
(heavy key), and the second column adjusts the purity of the final several studies for both aqueous- and vapor-phase reforming
product to the specified value by cutting between propylene glycol [45e47]. The main disadvantage of nickel-based catalysts is coke
R.X. Jimenez et al. / Renewable Energy 158 (2020) 181e191 185

formation, whereas for noble metal catalysts the high cost is an volume. For the first reactor, the raw material must be heated to the
important drawback. operational temperature of 973 K and the effluent must be cooled
Kinetic modeling of reforming of oxygenated compounds such down to 648 K, at which the HTS reactor operates, and then finally
as ethanol and methanol has been satisfactorily done through cooled to 498 K for the LTS [55].
power-law expressions [48]. The activation energy depends on the Due to the lack of information on all reactions in a same study
type of catalyst employed in the reactions. For noble metals (Pt and and under the same conditions, the conversion values and opera-
Ru), the reaction rate expression depends linearly on glycerol tional conditions adopted by Villaça [52] were used. The conversion
concentration due to the high activity of such catalysts, while for values obtained for methane generation and carbon formation are
other metals it is a function of a fractional power of the glycerol much lower than that of the main by-products, so it was decided to
concentration. The non-dependence of the rate on water concen- neglect these by-products. The carbon formed will accumulate at
tration is because it is used in excess [49]. Table 3 presents glycerol the active catalyst sites over time, thereby decreasing the efficiency
reforming kinetic parameters for power-law expressions obtained of the process. This issue was considered in the economic evalua-
for different catalysts. tion. Methane molar fraction decreases with increasing tempera-
The Soave-Redlich-Kwong (SRK) equation was chosen to ture, and for the temperature adopted in this work (973 K), no
calculate the thermodynamic properties in this simulation. This methane is detected in the reactor output stream. It can also be
choice is supported by glycerol reforming studies available in the observed that the amount of H2 produced almost remains constant
literature [50e54]. with increasing temperature, which is in agreement with experi-
The glycerol steam reforming process for hydrogen production ments [41].
is designed to meet the hydrogen demand required for hydro- As the temperature increases, the molar fraction of carbon
genolysis and is shown in Fig. 3. The stream properties are pre- monoxide increases. This is because the shift reaction (which
sented as Supporting Information, Table S4. Glycerol and water transforms CO to CO2) is exothermic and its yield is higher at low
enter the process at a 1:9 ratio. The reforming process involves temperatures. To model the two shift reactors (HTS and LTS), the
three reaction steps, carried out in three different reactors. Such Equilibrium Reactor block from Aspen HYSYS® was used to calculate
reactors operate under different conditions and employ different the output streams’ compositions. This can be justified by the fact
catalysts [55]. The first reactor performs reforming itself, as that this reaction indeed tends to equilibrium in the reactor. This
described in Eq. (4): stage of the simulation is different from that presented by Villaça
[52], but is in agreement with the considerations exposed earlier.
C3 H8 O3 #3 CO þ 4 H2 (4) Aspen HYSYS® contains in its database the equilibrium constant
value of this reaction as a function of temperature.
The water-gas shift reaction occurs in the second and third re- Excess water is added to the process. Thus, after the LTS reactor,
actors, as described in Eq. (5): water must be condensed in a flash vessel and recirculated.
Aspen HYSYS® does not allow the simulation of adsorption
CO þ H2 O#CO2 þ H2 (5) processes with conventional blocks. Thus, the PSA unit is simplified
to a Splitter block, in which a process stream is separated into
It is recommended to perform this reaction in two steps because several streams with defined compositions. In the present work,
it is exothermic. So, the first reactor, which is called the high- purity and separation data reported in the work of Villaça [52] and a
temperature shift reactor (HTS), operates at 648 K and atmo- pressure of 0.7 MPa were considered.
spheric pressure. The second reactor, which is called the low- For sizing, two adsorption towers were considered. The columns
temperature shift reactor (LTS), operates at 498 K and at the same operate in cycles, thus ensuring that the process operates contin-
pressure, in order to shift equilibrium in the products direction uously. The adsorption columns are filled with a type 5A alumi-
[52]. nosilicate molecular sieve, indicated to separate carbon dioxide
A Fe2O3 catalyst was used for the HTS reactor. This catalyst has a from hydrogen-rich streams [30]. The quantity of adsorbent can be
low thermal deactivation rate and allows to promote the kinetics by calculated considering an average adsorption capacity value of
increasing temperaure. For the LTS reactor, a high-activity CuO 0.255 kg of adsorbent per kg of adsorbate. With the bulk density,
catalyst was used, which increases the products formation rate the bed volume is obtained [30,56].
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 2.3. Economic analysis
surplus stream enters a Pressure Swing Adsorption (PSA) separa-
tion system, which produces hydrogen at a purity of 99.99%. The total investment cost was estimated by means of the Lang
The reforming reactor and the two shift reactors are of the fixed- method. The FOB price of each equipment was estimated from
bed PFR type. Thus, for design, it is first necessary to evaluate the correlations available in the literature [33]. These correlations were
amount of catalyst to be employed and then calculate the reactor 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].
 
kJ
Catalyst Temperature (K) Glycerol Order Steam Order Activation Energy
mol
Ru= Al2 O3 623e773 1 e 21.2
Pt= C 623e673 1 e e
Co= Al2 O3 723e823 0.1 0.4 67.2
Ni= Al2 O3 723e823 0.48 0.34 60.0
Co  Ni= Al2 O3 773e823 0.25 0.36 63.3
Ni= CeO2 873e923 0.233 e 103.4
Ni  ZrO2 = 973 0.3 e 43.4
CeO2
186 R.X. Jimenez et al. / Renewable Energy 158 (2020) 181e191

Fig. 3. Glycerol steam reforming flowsheet.

Table 4 electrolysis. The price of hydrogen produced by electrolysis using


Prices of raw materials, utilities and products involved in the glycerol hydro- solar energy was disregarded, since its value is very different from
genolysis and glycerol steam reforming processes.
others and the price range obtained would be too wide [59].
Raw Material Unit Price (US$/kg) Reference Rajkhowa et al. [25] quantified the dominant factors in the loss
Glycerol 0.4918 [64] of catalytic activity for a Cu-based catalyst during the glycerol
Hydrogen 1.28e4.14 [58] hydrogenolysis reaction, concluding that for this copper catalyst,
Process Water 0.0008 [65] sintering occurs due to the agglomeration of Cu atoms, decreasing
Utilities
a the effective area of the catalytic surface. The authors evaluated the
450 psi Vapor 0.0130 [66]
Cooling Water 0.0008 [65] stability of the catalyst for periods of 80e90 h under maximum
Fuel (US$/m3) 0.9768 [67] operability conditions and different raw material compositions. For
Electricity (US$/kWh) 0.1032 [68] a feed of pure glycerol, the catalyst showed high stability
Effluent Treatment 0.0560 [69]
throughout the experiments. However, for a feed stream containing
Catalysts
Cu= ZnO= Al2 O3 1.8000 [70]
pollutants such as sulfur, chlorine, and glycerides from biodiesel
Ru= Al2 O3 21000 [70] production, its activity decreases. This is one more reason to
Fe2 O3 13.900 [70] consider bidistilled glycerol as the raw material. Thus, the loss of
CuO 6.7000 [70] catalytic activity was neglected.
Products
In the hydrogen production process, Levalley et al. [61] indicate
Propylene Glycol 1.4500 [10]
that copper and iron catalysts for the shift reaction keep a good
a
Updated to 2018 with the Plant Cost Index. 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-
with the Plant Cost Index, published periodically by Chemical En-
position above 1373 K. As the reforming reactor operates at 973 K, it
gineering Magazine. The value of the Plant Cost Index for the year
was considered that thermal deactivation is not critical, and the
on which the correlations were based is 394. For 2018 it is 603.1
catalyst may be replaced together with the other catalysts.
[57].
The annual consumption of raw materials and utilities was ob-
For non-corrosive environments with hydrogen, Seider et al.
tained directly from Aspen HYSYS®. For the calculation of
[33] recommend the use of a 1% Cr and 0.5% Mo steel alloy (ASME
manpower, the Brazilian average wage for industrial operators was
code SA-387B). This material was applied only in large equipment
considered [63]. The number of workers per shift was estimated
used in operations involving the presence of hydrogen.
from the number of equipment units, considering five shifts as
For the calculation of production costs, it was considered that
recommended by Seider et al. [33]. The remaining production costs
the plant operates continuously 24 h a day for 330 days a year.
were estimated from correlations available in the literature.
Prices for raw materials, utilities and catalysts are summarized in
Table 4.
It was not possible to stablish an average price for hydrogen, 3. Results and discussion
because several production processes employ a variety of raw
materials and energy sources. For instance, for hydrogen produced A brief description of the process equipment size and cost can be
via water electrolysis the cost may vary from 1.28 to 4.14 US$/kg, found in Table 6, while the relative costs of equipment pieces are
these values being the most recent that could be found in the shown in Fig. 4. To produce 821 kg/h of propylene glycol, 22 kg/h of
literature [58]. Negro et al. [59] presented the different costs of hydrogen, 1070 kg/h of glycerol, and 267 kg/h of water are needed.
hydrogen in Brazil if it was produced via electrolysis, fossil fuel Producing 23 kg/h of hydrogen requires 174 kg/h of glycerol and
reforming, or renewable fuel reforming. However, these values are 127 kg/h of water. It is clear that the distillation columns are the
out of date and may affect the veracity of the economic analysis. For most expensive equipment in the propylene glycol process. This is
this reason, these costs were updated using the Brazilian National due to the high purity specification of PG, which leads to larger
Consumer Price Index [60], that represents the accumulated infla- distillation columns. Due to the high steam demand, the steam
tion in the country over time. The results can be observed in Table 5. generation unit costs represent more than half of the equipment
Thus, in the present work, the hydrogen price range is 0.23e4.14 costs for the steam reforming process, as can be seen in Fig. 5.
US$/kg, considering the lowest price of hydrogen if it is obtained by Figs. 6 and 7 show the costs distribution of the utilities
natural gas reforming and the highest price if it is obtained by water consumed in each process. Cooling water and steam are the main
utilities used in the PG process because of the distillation columns’
R.X. Jimenez et al. / Renewable Energy 158 (2020) 181e191 187

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 Nuclear Energy 1.98 3.62


Electrolysis 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.

reboilers and condensers. Moreover, there is also a high electricity


demand due to the compressor. 1,441,653.78 US$. The detailed cost statement of the processes is
Considering the production of partially renewable propylene presented as Supporting Information, Table S1.
glycol (with hydrogen purchased from an external source) with an It was found that the break-even price of propylene glycol is 1.36
average price of 2.185 US$/kg for H2, the break-even price for pro- US$/kg for a fully renewable PG production process (that is, a
pylene glycol is 1.17 US$/kg, with a total investment of 4,722,678.07 propylene glycol production process integrated with a glycerol
US$. The break-even price of hydrogen production via glycerol reforming process to produce hydrogen). Therefore, the generation
steam reforming is 9.01 US$/kg, with a total investment of of a 100% green PG resulted in a 16% increase in its final price.
188 R.X. Jimenez et al. / Renewable Energy 158 (2020) 181e191

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. 6. Propylene glycol process utilities cost distribution.

Fig. 8. Variation in the net annual profit as a function of PG selling price.

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$/kg is obtained.
Thus, it can be said that even for large variations in the price of
hydrogen, the propylene glycol production cost is not significantly
affected.
If a similar sensitivity analysis concerning the glycerol price is
performed, it is concluded that for a 1% increase in the price of
Fig. 7. Steam reforming process utilities cost distribution.
glycerol the glycerol reforming process cost increases by 0.41% and
the propylene glycol production process cost increases by 0.57%.
The selling price of PG and the quantity sold is subject to
However, this additional expense can be offset when the expanding changes due to market offer and demand. Fig. 8 shows the variation
market shares of more sustainable products are taken into account. of annual net profit with the variation of the PG price. This analysis
The costs distributions of the evaluated processes are shown in was performed for a constant propylene glycol production of
Table 7. It can be seen that raw material and catalysts correspond to 824 kg/h, which was the amount obtained by processing all the
55e60% of the total production cost. The renewable PG production glycerol generated by an average biodiesel plant in Brazil.
cost is higher in nearly US$ 1.2 million, due to the price of renew- A similar analysis is shown in Fig. 9, but in this case the amount
able hydrogen. of PG produced is varied and its price is kept constant. For the
The main source of uncertainty in the economic analysis for the renewable propylene glycol, the minimum quantity to be produced
propylene glycol production process considering that the hydrogen to achieve a zero annual net profit is approximately 776 kg/h. For
is purchased from an external source is its purchase price, which the partially renewable propylene glycol with hydrogen obtained
R.X. Jimenez et al. / Renewable Energy 158 (2020) 181e191 189

Fig. 9. Variation in the net annual profit as a function of PG production.

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, Fig. 10. Comparison of the PG production processes. (A) Raw material costs. (B) Util-
because of the use of H2SO4, which is dissolved in water in 0,1% wt. ities demands.
To produce the same amount of PG, 4000 kg/h of water, 40 kg/h of
reached in the other processes. The reactor temperature is a critical
H2SO4, 643 kg/h of methanol and 675 kg/h of propylene oxide are
variable, because of propylene oxide’s low boiling point. Addi-
needed. In comparison, the partially green process uses only 267 kg/
tionally, it is important to notice that the renewable PG process
h of water and the fully green process consumes a total of 394 kg/h
consumes more utilities than the partially renewable PG process,
of process water. Both processes require none of the other com-
due to steam reforming.
ponents. Aspen HYSYS® also gives an estimation of the carbon
Finally, these results suggest that both green and partially green
emissions. Its 2237 kg/h for the conventional process, against
PG are economically feasible, because both break-even prices of
467 kg/h for the propylene glycol production stage and 141 kg/h for
propylene glycol were lower than the reported market price, even
steam reforming.
though its price depends on the existing demand. The environ-
Fig. 10 (A) shows that the conventional process’s raw material
mental advantages of these products seem clear, but, for future
cost is greater in nearly 25% than the partially renewable PG process
work, it would be interesting to perform a formal environmental
and 16% than the renewable PG process. This difference is caused by
analysis of the green propylene glycol production process, such as a
the high demand of solvent, which is used in a propylene oxide-
Life Cycle Assessment (LCA), and to compare it to the conventional
methanol equivolume mixture. In Fig. 10 (B) the utilities demand
PG production process to have more elucidative results in this
of each process can be compared. The propylene oxide process
sense.
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 4. Conclusion
separation steps needed to achieve the same purity that has been
The present work simulated the propylene glycol production
190 R.X. Jimenez et al. / Renewable Energy 158 (2020) 181e191

process using glycerol from biodiesel production as the main raw 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
material, in a Brazilian scenario. A total production cost of 1.17 US$/
hydrogenolysis of glycerol to propylene glycol, Appl. Catal. Gen. 281 (2005)
kg was obtained when PG is produced using hydrogen from an 225e231, https://doi.org/10.1016/J.APCATA.2004.11.033.
external source. When it is produced using hydrogen obtained from [14] M.N. Gatti, M.D. Mizrahi, J.M. Ramallo-Lopez, F. Pompeo, G.F. Santori,
glycerol steam reforming, its price increases to 1.36 US$/kg. A 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,
hydrogen production cost of 9.01 US$/kg was calculated when it is Appl. Catal. Gen. 548 (2017) 24e32, https://doi.org/10.1016/
obtained from glycerol steam reforming, which is much higher J.APCATA.2017.08.037.
when compared to the market prices of hydrogen obtained by other [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/
processes. It has caused a nearly 70% drop in the annual net profit of J.JCAT.2007.05.008.
propylene glycol production in the base scenario, although the [16] T. Miyazawa, Y. Kusunoki, K. Kunimori, K. Tomishige, Glycerol conversion in
process remains economically viable. Fully renewable PG is 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/
approximately 16% more expensive than partially renewable PG, J.JCAT.2006.03.023.
but it is possible, depending on the market. [17] C. Montassier, J.C. Me nezo, 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.
Declaration of competing interest [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,
The authors declare that they have no known competing Top. Catal. 53 (2010) 1214e1216, https://doi.org/10.1007/s11244-010-9565-
x.
financial interests or personal relationships that could have
[19] R.V. Sharma, P. Kumar, A.K. Dalai, Selective hydrogenolysis of glycerol to
appeared to influence the work reported in this paper. 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.
CRediT authorship contribution statement
[20] E.S. Vasiliadou, A.A. Lemonidou, Glycerol transformation to value added C3
diols: reaction mechanism, kinetic, and engineering aspects, Wiley Inter-
Roberto X. Jime nez: Conceptualization, Investigation, Software, discip. Rev. Energy Environ. 4 (2015) 486e520, https://doi.org/10.1002/
 F. Young:
Writing - original draft, Writing - review & editing. Andre wene.159.
[21] G.J. Suppes, W.R. Sutterlin, M. Dasari, Method of Producing Lower Alcohols
Conceptualization, Methodology, Supervision, Visualization, from Glycerol, 2007. US7943805B2.
Writing - original draft, Writing - review & editing. Heloisa L.S. [22] C. Montassier, D. Giraud, J. Barbier, Polyol conversion by liquid phase het-
Fernandes: Conceptualization, Supervision, Writing - original draft, erogeneous catalysis over metals, Stud. Surf. Sci. Catal. 41 (1988) 165e170,
https://doi.org/10.1016/S0167-2991(09)60811-9.
Writing - review & editing. [23] S. Wang, Y. Zhang, H. Liu, Selective hydrogenolysis of glycerol to propylene
glycol on Cu-ZnO composite catalysts: structural requirements and reaction
Appendix A. Supplementary data 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
Supplementary data to this article can be found online at Alcohols: glycerol to glycols, Ind. Eng. Chem. Res. 42 (2003) 5467e5472,
https://doi.org/10.1016/j.renene.2020.05.126. 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
References (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
[1] A.F. Young, F.L.P. Pessoa, E.M. Queiroz, Biodiesel Production Technologies - Co-Ni/Al2O3, Ind. Eng. Chem. Res. 49 (2010) 10804e10817, https://doi.org/
Supercritical and Enzymatic Production, Novas Ediço ~es Acade ^micas, Saar- 10.1021/ie100462t.
brücken, Germany, 2015 (in Portuguese). [27] A. Lancia, D. Musmarra, F. Pepe, Vapor-liquid equilibria for mixtures of
[2] M.C.S. de Mello, H.G.D. Villardi, A.F. Young, F.L.P. Pessoa, A.M. Salgado, Life ethylene glycol, propylene glycol, and water between 98º and 122ºC, J. Chem.
cycle assessment of biodiesel produced by the methylic-alkaline and ethylic- Eng. Jpn. 29 (1996) 449e455, https://doi.org/10.1252/jcej.29.449.
enzymatic routes, Fuel 208 (2017) 329e336, https://doi.org/10.1016/ [28] E.S. Vasiliadou, A.A. Lemonidou, Kinetic study of liquid-phase glycerol
j.fuel.2017.07.014. hydrogenolysis over Cu/SiO2 catalyst, Chem. Eng. J. 231 (2013) 103e112,
[3] M.R. Monteiro, C.L. Kugelmeier, R.S. Pinheiro, M.O. Batalha, A. da Silva Ce sar, https://doi.org/10.1016/j.cej.2013.06.096.
Glycerol from biodiesel production: technological paths for sustainability, [29] Y. Xi, J.E. Holladay, J.G. Frye, A.A. Oberg, J.E. Jackson, D.J. Miller, A kinetic and
Renew. Sustain. Energy Rev. 88 (2018) 109e122, https://doi.org/10.1016/ mass transfer model for glycerol hydrogenolysis in a trickle-bed reactor, Org.
j.rser.2018.02.019. Process Res. Dev. 14 (2010) 1304e1312, https://doi.org/10.1021/op900336a.
[4] ANP - Brazilian National Agency of Petroleum, Natural Gas and Biofuels. [30] R. Perry, D. Green, Perry’s Chemical Engineers’ Handbook, McGraw-Hil, New
Biodisel, 2018 accessed 10.1.19, http://www.anp.gov.br/producao-de- York, United States, 1997.
biocombustiveis/biodiesel/simp-biodisel (in Portuguese). [31] C.N. Satterfield, Trickle-bed reactors, AIChE J. 21 (1975) 209e228, https://
[5] C.J.A. Mota, Pinto, Catalytic transformations of glycerol for innovation in the doi.org/10.1002/aic.690210202.
chemical industry, Rev. Virtual Quim. 9 (2017) 135e149, https://doi.org/ [32] R.J. Farrauto, C.H. Bartholomew, Fundamentals of Industrial Catalytic Pro-
10.21577/1984-6835.20170011. cesses, Blackie Academic & Professional, London, United Kingdom, 1997.
[6] Z. Freitas, Glycerin as raw material for chemical industry: Glicerina como [33] W.D. Seider, J.D. Seader, D.R. Lewin, Product and Process Design Principles -
mate ria-prima para indústria química: evaluation of research efforts and Synthesis, Analysis, and Evaluation, Second, John Wiley and Sons, Inc, Inter-
commercial initiatives, Federal University of Rio de Janeiro, Rio de Janeiro, national Edition, 2003.
Brazil, 2013 (in Portuguese). [34] D. Woods, Rules of Thumb in Engineering Practice, Wiley, Darmstadt, Ger-
[7] S. Veluturla, N. Archna, D. Subba Rao, N. Hezil, I.S. Indraja, S. Spoorthi, Catalytic many, 2007.
valorization of raw glycerol derived from biodiesel: a review, Biofuels 9 (2018) [35] R. Sinnott, G. Towler, Chemical Engineering Design, Elsevier Inc., United
305e314, https://doi.org/10.1080/17597269.2016.1266234. States, 2008.
[8] H. Mitta, P.K. Seelam, S. Ojala, R.L. Keiski, P. Balla, Tuning Y-zeolite based [36] C. Ferreira, Process Flowsheet Proposal for Purification of Glycerol from Bio-
catalyst with copper for enhanced activity and selectivity in vapor phase diesel Production, Federal University of Rio de Janeiro, Rio de Janeiro, Brazil,
hydrogenolysis of glycerol to 1,2-propanediol, Appl. Catal. Gen. 550 (2018) 2018 (in Portuguese).
308e319, https://doi.org/10.1016/j.apcata.2017.10.019. [37] I. Gandarias, P.L. Arias, I. Agirrezabal-Telleria, Economic assessment for the
[9] Y. Nakagawa, K. Tomishige, Heterogeneous catalysis of the glycerol hydro- production of 1,2-Propanediol from bioglycerol hydrogenolysis using molec-
genolysis, Catal. Sci. Technol. 1 (2011) 179, https://doi.org/10.1039/ ular hydrogen or hydrogen donor molecules, Environ. Prog. Sustain. Energy 35
c0cy00054j. (2016) 447e454, https://doi.org/10.1002/ep.12232.
[10] Abiquim - Brazilian Chemical Industry Association, Annual Report, 2018 (in [38] R.R. Davda, J.W. Shabaker, G.W. Huber, R.D. Cortright, J.A. Dumesic, A review
Portuguse). of catalytic issues and process conditions for renewable hydrogen and alkanes
[11] Y. Zhang, M.A. Dube , D.D. McLean, M. Kates, Biodiesel production from waste by aqueous-phase reforming of oxygenated hydrocarbons over supported
cooking oil: 1. Process design and technological assessment, Bioresour. metal catalysts, Appl. Catal. B Environ. 56 (2005) 171e186, https://doi.org/
Technol. 89 (2003) 1e16, https://doi.org/10.1016/S0960-8524(03)00040-3. 10.1016/J.APCATB.2004.04.027.
[12] Z. Zhou, X. Li, T. Zeng, W. Hong, Z. Cheng, W. Yuan, Kinetics of hydrogenolysis [39] A. Iriondo, V.L. Barrio, J.F. Cambra, P.L. Arias, M.B. Güemez, R.M. Navarro,
of glycerol to propylene glycol over Cu-ZnO-Al2O3 catalysts, Chin. J. Chem. M.C. S anchez-S anchez, J.L.G. Fierro, Hydrogen production from glycerol over
R.X. Jimenez et al. / Renewable Energy 158 (2020) 181e191 191

nickel catalysts supported on Al2O3 modified by Mg, Zr, Ce or La, Top. Catal. 49 [54] G. Yang, H. Yu, F. Peng, H. Wang, J. Yang, D. Xie, Thermodynamic analysis of
(2008) 46e58, https://doi.org/10.1007/s11244-008-9060-9. hydrogen generation via oxidative steam reforming of glycerol, Renew. En-
[40] P.D. Vaidya, A.E. Rodrigues, Glycerol reforming for hydrogen production: a ergy 36 (2011) 2120e2127, https://doi.org/10.1016/J.RENENE.2011.01.022.
review, Chem. Eng. Technol. 32 (2009) 1463e1469, https://doi.org/10.1002/ [55] M.M.V.M. Souza, Hydrogen Technology, Synergia, Rio de Janeiro, Brazil, 2009
ceat.200900120. (in Portuguese).
[41] R.L. Manfro, N.F.P. Ribeiro, M.M.V.M. Souza, Production of hydrogen from [56] A. Hines, R. Maddox, Mass Transfer Fundamentals and Applications, Prentice-
steam reforming of glycerol using nickel catalysts supported on Al2O3, CeO2 Hall, Michigan, United States, 1985.
and ZrO2, Catal. Sustain. Energy. 1 (2013) 60e70, https://doi.org/10.2478/cse- [57] S. Jenkins, Economic Indicators, Chem. Eng, 2019. https://www.
2013-0001. chemengonline.com/cepci-updates-january-2018-prelim-and-december-
[42] I. Iliuta, H.R. Radfarnia, M.C. Iliuta, Hydrogen production by sorption- 2017-final/?printmode¼1. accessed 10.3.18.
enhanced steam glycerol reforming: sorption kinetics and reactor simula- [58] L. Bonfim-Rocha, M.L. Gimenes, S.H. Bernardo de Faria, R.O. Silva, L.J. Esteller,
tion, AIChE J. 59 (2013) 2105e2118, https://doi.org/10.1002/aic.13979. Multi-objective design of a new sustainable scenario for bio-methanol pro-
[43] R.L. Manfro, M.M.V.M. Souza, Catalyst for hydrogen production through duction in Brazil, J. Clean. Prod. 187 (2018) 1043e1056, https://doi.org/
glycerol reforming review, Curr. Top. Catal. 11 (2014) 37e54. 10.1016/J.JCLEPRO.2018.03.267.
[44] S. Adhikari, S. Fernando, S.R. Gwaltney, S.D. Filip To, R. Mark Bricka, [59] M.L.M. Negro, R.C. Garner, M. Linardi, Mass production of hydrogen in Brazil,
P.H. Steele, A. Haryanto, A thermodynamic analysis of hydrogen production by world clim, Energy Event 3 (2003) 293e298.
steam reforming of glycerol, Int. J. Hydrogen Energy 32 (2007) 2875e2880, [60] IBGE - Brazilian Institute of Geography and Statistics, National Consumer Price
https://doi.org/10.1016/J.IJHYDENE.2007.03.023. Index, 2019. https://www.ibge.gov.br/estatisticas/economicas/precos-e-
[45] B. Zhang, X. Tang, Y. Li, Y. Xu, W. Shen, Hydrogen production from steam custos/9256-indice-nacional-de-precos-ao-consumidor-amplo.html?
reforming of ethanol and glycerol over ceria-supported metal catalysts, Int. J. =&t=series-historicas. accessed 9.29.18.
Hydrogen Energy 32 (2007) 2367e2373, https://doi.org/10.1016/ [61] T.L. LeValley, A.R. Richard, M. Fan, The progress in water gas shift and steam
J.IJHYDENE.2006.11.003. reforming hydrogen production technologies e a review, Int. J. Hydrogen
[46] A.O. Menezes, M.T. Rodrigues, A. Zimmaro, L.E.P. Borges, M.A. Fraga, Produc- Energy 39 (2014) 16983e17000, https://doi.org/10.1016/
tion of renewable hydrogen from aqueous-phase reforming of glycerol over Pt j.ijhydene.2014.08.041.
catalysts supported on different oxides, Renew. Energy 36 (2011) 595e599, [62] A. Berman, M. Epstein, Ruthenium catalysts for high temperature solar
https://doi.org/10.1016/J.RENENE.2010.08.004. reforming of methane, in: Hydrogen Power: Theoretical and Engineering
[47] K.S. Avasthi, R.N. Reddy, S. Patel, Challenges in the production of hydrogen Solutions, Springer, Dordrecht, Netherlands, 1998, pp. 213e218, https://
from glycerol e a biodiesel byproduct via steam reforming process, Procedia doi.org/10.1007/978-94-015-9054-9_26.
Eng. 51 (2013) 423e429, https://doi.org/10.1016/J.PROENG.2013.01.059. [63] Glassdoor, Industrial Operator Salaries, 2018 accessed 9.20.18, https://www.
[48] S. Adhikari, S.D. Fernando, A. Haryanto, Kinetics and reactor modeling of lovemondays.com.br/salarios/cargo/salario-operador-industrial (in
hydrogen production from glycerol via steam reforming process over Ni/CeO2 Portuguese).
catalysts, Chem. Eng. Technol. 32 (2009) 541e547, https://doi.org/10.1002/ [64] Comex Stat, 2018. http://comexstat.mdic.gov.br/pt/home. accessed 9.20.18.
ceat.200800462. [65] SNIS - National System of Information about Sanitation, Diagnosis of the
[49] R. Sundari, P.D. Vaidya, Reaction kinetics of glycerol steam reforming using a Water and Sewage Services - 2017, Brasilia, Brazil, 2019.
Ru/Al2O3 catalyst, Energy Fuels 26 (2012) 4195e4204, https://doi.org/ [66] C.A.G. Perlingeiro, Process Systems Engineering, Blucher, Sa ~o Paulo, Brazil,
10.1021/ef300658n. 2005 (in Portuguese).
[50] H. Chen, Y. Ding, N.T. Cong, B. Dou, V. Dupont, M. Ghadiri, P.T. Williams, [67] Comga s - S~
ao Paulo Gas Company, Pipe Natural Gas Tariffs, 2018 accessed
A comparative study on hydrogen production from steam-glycerol reforming: 10.1.18, https://www.comgas.com.br/tarifas/comercial/ (in Portuguse).
thermodynamics and experimental, Renew. Energy 36 (2011) 779e788, [68] ANEEL - National Electricity Regulatory Agency, Energy consumption report
https://doi.org/10.1016/j.renene.2010.07.026. by sector, accessed 9.20.18, http://relatorios.aneel.gov.br/_layouts/xlviewer.
[51] Y. Liu, R. Farruto, A. Lawal, Autothermal reforming of glycerol in a dual layer aspx?id=/RelatoriosSAS/RelSampRegCC.xlsx&Source=http://relatorios.aneel.
monolith catalyst, Chem. Eng. Sci. 89 (2013) 31e39, https://doi.org/10.1016/ gov.br/RelatoriosSAS/Forms/AllItems.aspx&DefaultItemOpen=1, 2018 (in
J.CES.2012.11.030. Portuguese).
[52] J. Villaça, Proposal and Simulation of the Glycerol Steam Reforming Process for [69] R. Turton, R. Bailie, W. Whiting, J. Shaeiwitz, Analysis, Synthesis and Design of
Hydrogen Obtaining, Federal University of Rio de Janeiro, Rio de Janeiro, Chemical Processes, Prentice Hall, Boston, United States, 2008.
Brazil, 2018 (in Portuguese). [70] Sigma-Aldrich, Product Directory, 2018. https://www.sigmaaldrich.com/
[53] X. Wang, S. Li, H. Wang, B. Liu, X. Ma, Thermodynamic analysis of glycerin technical-service-home/product-catalog.html. accessed 10.1.18.
steam reforming, Energy Fuels 22 (2008) 4285e4291, https://doi.org/10.1021/ [71] H.S. Fogler, Elements of Chemical Engineering, fourth ed., Prentice Hall, New
ef800487r. Jersey, United States, 2009.

You might also like