You are on page 1of 17

Chemical Papers (2019) 73:2619–2635

https://doi.org/10.1007/s11696-019-00840-8

REVIEW

A review of steam reforming of glycerol


Adewale George Adeniyi1   · Joshua O. Ighalo1

Received: 26 February 2019 / Accepted: 29 May 2019 / Published online: 4 June 2019
© Institute of Chemistry, Slovak Academy of Sciences 2019

Abstract
The increase in glycerol in the global market (from increased biodiesel production) has led to a glut. Of the different meth-
ods of utilising crude glycerol, steam reforming has been explored as a viable energy generation technique. Over the years,
research has explored process optimisation, process catalysis, thermodynamics analysis, kinetic modelling, exergy and energy
analyses, techno-economic analyses, environmental performance, and computational fluid dynamics (CFD) simulations in this
regard. This paper reviewed and catalogued the current state of knowledge on the production of bio-hydrogen by the steam
reforming of glycerol obtained from transesterification while revealing key knowledge gaps. It was observed that impurities
in crude glycerol have much more negative effect on the overall yield than on the product selectivity. Impurities such as water
and methanol favour the production hydrogen. The catalysts for steam reforming are usually noble and transition metals co-
precipitated together or impregnated on metal oxides and ores. Nickel is considered as the best catalyst for the process with
temperature and pressure optimal pegged around 650–900 °C and 1 atm. Gaping gaps in knowledge abound in the research
area and these are in the domain of crude glycerol use, fluidised bed use, artificial sorbent use, pilot scale studies, cost and
profitability analysis, and investigations on carbon deposition behaviour.

Keywords  Glycerol · Steam reforming · Crude · Catalyst · Hydrogen

List of symbols bl Total number of gram atoms of element l in the


APR Aqueous phase reforming reaction mixture
G Gibbs free energy M Total number of atomic elements
T Temperature n Reaction order
P Pressure E Activation energy
ni Number of moles of species i ri Reaction rate of i
K Total number of chemical species in the reaction R Universal gas constant
mixture Ci Concentration of i
R Gas constant k0 Reaction constant
𝜇i Chemical potential of species i
yi Mole fraction of species i
ΔG0i Standard Gibbs free energy of the formation of Introduction
species i
N Total number of moles of all species in the gas The global energy demand has been on the rise and fossil-
mixture fuel reserves are depleting (Koh and Ghazi 2011). These
ali Number of gram atoms of element l in 1 mol of are the consequences of global population increase and
species i industrialization (Pambudi et al. 2013). There is an unfa-
vourable mismatch in the global demand and supply levels
of fossil-based crude (Ofari-Boateng et al. 2012). Fossil-
fuel use has also exacerbated global warming and climate
* Adewale George Adeniyi changes (Averill et al. 2010). Considering all these, focus
adeniyi.ag@unilorin.edu.ng has shifted to renewable energy sources that utilises cleaner
1
Chemical Engineering Department, Faculty of Engineering
production technologies and can serve as viable alternatives
and Technology, University of Ilorin, P. M. B. 1515, Ilorin, to fossil fuels.
Nigeria

13
Vol.:(0123456789)

2620 Chemical Papers (2019) 73:2619–2635

Biodiesel has gained importance due to the adverse envi- and He 2006; Yang et al. 2012). The crude glycerol usually
ronmental impact of petroleum fuels and because it can has a dark colour and contains variable amounts of soap,
be obtained from a renewable source (Blanco-Marigorta catalyst, alcohol (typically methanol), mono-glycerides, di-
et al. 2013; Tiwari et al. 2007). Biodiesel is defined as the glycerides, glycerol oligomers, polymers, water, unreacted
mono-alkyl esters of long-chain fatty acids derived from the triacylglycerol, and biodiesel (Schwengber et al. 2016).
reaction of vegetable oils or animal fats and alcohol with From Fig. 1, we observe the rise in biodiesel produc-
or without the use of a catalyst (Atabani et al. 2013). Bio- tion from year 2000 and then a steep increase in produc-
diesel is bio-degradable, non-toxic, and environmentally tion beyond 2005 (Bagnato et al. 2017). Europe still has the
friendly (Dou et al. 2014). Biodiesel is obtained via the greatest share in the global output of biodiesel. This increase
transesterification reaction and the by-product of the pro- in glycerol in the market (due to increased biodiesel produc-
cess is glycerol (­ C3H8O3). In converting vegetable oils into tion) has led to a glut in the global market (Johnson and
biodiesel, approximately 10% (w/w) of glycerol is produced Taconi 2007). The glycerol price dropped by two-thirds
as a by-product (Adeniyi et al. 2018; Adhikari et al. 2008a; within the last decade (Adhikari et al. 2007b) because of
Behr et al. 2008). This glycerol is crude as it comes with surplus amounts in the market and is still dropping in the
various impurities usually 10–30% (Dou et al. 2014). The current decade. Currently, there are many applications of
nature and volume of the impurities depends on the nature pure glycerol and these include personal care products, food,
of the feedstock and catalyst used in the process (Thompson oral care, tobacco, and polyurethane production (Adhikari
et al. 2007a). From Fig. 2, we see that cosmetics (37%) and
food (25%) are the major applications of the product (Bag-
nato et al. 2017). Unfortunately, commercial development of
alternative processes for pure glycerol utilization has been
limited, as the price of glycerol purification has made it eco-
nomically unattractive as a feedstock chemical (Johnson and
Taconi 2007; Thompson and He 2006). Most by-product
glycerol is sent to water treatment for digestion. It can also
be purified by distillation, though this is expensive process
and uneconomic (Slinn et al. 2008). Other alternative uses
of glycerol include its conversion to of esters, ethers, acetals,
ketals of glycerol, diols, epoxides, aldehydes, ketones, and
carboxylic acids (Behr et  al. 2008). Reactions involves
hydrogenolysis (Chaminand et al. 2004; Dasari et al. 2005),
selective oxidation (Chaminand et al. 2004; Pagliaro et al.
2007; Zhou et al. 2008), etherification (Pagliaro et al. 2007),
Fig. 1  Biodiesel and glycerol production (Bagnato et al. 2017) dehydration (Pagliaro et al. 2007), pyrolysis and gasification

Fig. 2  Percentage distribution Others


of the main glycerol applica- Poly-urethane
5%
tions found in the open litera- 7%
ture (Bagnato et al. 2017) Pharmaceutical
8% Cosmetics
37%
Tobacco
9%

Alkyl Resin
9%

Food
25%

13
Chemical Papers (2019) 73:2619–2635 2621

(Zhou et al. 2008), and carboxylation (Zhou et al. 2008) on the production of bio-hydrogen by the steam reforming
amongst others. Glycerol can also be utilised in biochemi- of glycerol and to elucidate possible knowledge gaps and
cal conversion processes (such as fermentation) to obtain future perspectives.
useful biochemical like succinic acid, propionic acid, citric
acid, ethanol, dihydroxyacetone, among others (Bagnato
et al. 2017; Da Silva et al. 2009; Ito et al. 2005; Yazdani and Glycerol steam‑reforming theory
Gonzalez 2007). It can also be added in ruminant and non-
ruminant animal diets (Yang et al. 2012). Glycerol ( C3 H8 O3 ) is also known as 1,2,3-propane tri-ol or
One viable and proven possibility is using it as a source glycerine. The chemical structure is presented in Fig. 3.
for the production of hydrogen via the steam-reforming Steam reforming is a high-temperature process whereby
process (Behr et al. 2008). Bio-hydrogen (hydrogen gas glycerol is converted to a hydrogen-rich gaseous product
obtained from bio-sources and bio-processes) are viable known as synthesis gas. The general expressions for the
options for the future due to their theoretically carbon–neu- equation of reaction of the steam reforming of any oxygen-
tral and renewable nature (Adeniyi and Ighalo 2019). Bio- ated organic compound were presented by Czernik et al.
hydrogen poses no environmental problem and also does not (2002):
contribute to atmospheric carbon dioxide emissions (Chat-
m
( )
tanathan et al. 2012). Besides glycerol, other bio-sources Cn Hm Ok + (n − k)H2 O ↔ nCO + n + − k H2 . (1)
2
have been proved to be good feedstock for steam reform-
ing such as bio-oil (Adeniyi et al. 2019; Chen et al. 2016; The chemical reactions involved in the steam reforming
Fu et al. 2014; Remón et al. 2015; Wang et al. 1997; Xie of glycerol into synthesis gas are well understood and have
et al. 2016), vegetable oils (Yenumala and Maity 2011), and been previously outlined in the literature (Dave and Pant
other oxygenated organic compounds (Abdullah et al. 2014; 2011; Pompeo et al. 2010; Slinn et al. 2008; Vaidya and
Hoang et al. 2015; Iwasa et al. 2010; Nabgan et al. 2016; Rodrigues 2009):
Thaicharoensutcharittham et al. 2011; Shurong Wang et al.
2015). At present, almost 95% of the hydrogen produced is ΔH0r = + 250 kJ/mol . (2)
( )
C3 H8 O3 ↔ 3CO + 4H2
obtained from fossil-fuel-based feedstock (Adhikari et al.
2007c) and global hydrogen demand is growing due to the Overall reaction for glycerol steam (Adhikari et  al.
technological advancements in fuel cell industry (Adhikari 2007c, 2008b; Slinn et al. 2008) is presented in the follow-
et al. 2007c). Hydrogen gas possesses the highest energy ing equation:
content per unit of weight (120.7 kJ/g), compared to any C3 H8 O3 + 3H2 O → 7H2 + 3CO2 ΔH0r = + 128 kJ/mol .
( )
of the known fuels (Haryanto et al. 2005). Other reforming
(3)
techniques includes autothermal reforming (ATR), supercrit-
Due to excess steam in the process, the carbon monoxide is
ical water reforming (SCWR), partial oxidation reforming
oxidised to carbon dioxide also leading to the evolution of
(POR), liquid-phase reforming (LPR), and chemical looping
more hydrogen gas. The equation is reversible, and hence,
reforming (CLR) (Schwengber et al. 2016).
both species do exist in the system at dynamic chemical
Over the years, the steam reforming of glycerol has been
equilibrium. This is known as the water–gas shift reaction:
studied by researchers. These studies have come in the
domain of energy and exergy analyses (Hajjaji et al. 2014),
(4)
( 0 )
CO + H2 O ↔ CO2 + H2 ΔHr = − 41 kJ/mol .
kinetic modelling and thermodynamic analysis (Adhikari
et al. 2007a, b; Chen et al. 2009; Da Silva and Müller 2011; At low temperatures, methanation can occur whereby some
Patel et al. 2013; Wang et al. 2008; Wang et al. 2010; Yang of the hydrogen produced combines reversibly with the oxy-
et al. 2011; Yenumala and Maity 2011), sorption enhance- genated compounds in the system to give methane:
ment techniques (Da Silva and Müller 2011; Dou et  al.
2009), catalysis and the minimisation of carbon deposition CO + 3H2 ↔ CH4 + H2 O
( 0 )
ΔHr = − 206 kJ/mol ,
(Adhikari et al. 2007a, c, 2008a, b, 2009; Cheng et al. 2010; (5)
Chiodo et al. 2010; Dave and Pant 2011; Hirai et al. 2005;
Nichele et al. 2012; Pompeo et al. 2010; Profeti et al. 2009;
Zhang et al. 2007), and process simulations (Adeniyi and
Ighalo 2018; Dou et al. 2008; Dou and Song 2010). Bagnato
et al. (2017) presented a critical review of glycerol steam
reforming though with a greater emphasis on the overall pro-
duction and application of the liquid. The aim of this paper
is to review and catalogue the current state of knowledge Fig. 3  Chemical structure of glycerol

13

2622 Chemical Papers (2019) 73:2619–2635

Table 1  Typical composition of crude glycerol (in wt%)


( 0 )
CO2 + 4H2 ↔ CH4 + 2H2 O ΔHr = − 165 kJ/mol . Components Slinn et al. Veiga and Bussi Veiga
(6) (2008) (2016) et al.
This reaction is exothermic, and, hence, will be favoured at (2017)
low temperatures as opposed to the steam-reforming reac-
Glycerol 33% 64% 64%
tions that are endothermic and favoured by higher tempera-
Fatty matter 40% – –
tures (Adeniyi and Ighalo 2018). Similarly, higher molecular
Methanol 23% < 5% < 5%
weight hydrocarbons can be synthesised during the process
water 3.2%
especially at low temperatures and for heavier feeds. At high
Ash 3.8%
temperatures, coke formation occurs in the system and can
Inorganic salts – 5.7% 5.7%
lead to depositions on the catalyst surface and subsequent
Polyglycerol – 26% 26%
inactivation. Carbon deposition is a very serious challenge
in catalytic steam-reforming systems utilising heterogeneous
catalysts. The deposited carbon on the surface of the solid
catalyst prevents reactants from gaining access to active alongside the negative effect of the carbon source from
sites thereby reducing the catalyst activity. Possible reac- the fatty matter, then the 30% lower value is justified. The
tions leading to carbon deposition in the steam-reforming negative effect of the carbon source from the fatty matter
system (Adhikari et al. 2007b; Schwengber et al. 2016) are is because long-chain fatty acid impurities are harder to
mentioned in the following equations: reform and are more likely to form carbon. It was, however,
recommended in the study that harsher reaction conditions
(7)
( 0
such as longer residence time and higher temperatures and
)
2CO ↔ CO2 + C ΔHr = − 172 kJ/mol ,
S/C ratios should be utilised when crude glycerol is used
as feedstock to overcome the negative effect of the impuri-
(8)
( 0 )
CH4 ↔ 2H2 + C ΔHr = + 75 kJ/mol , ties. Im general, impurities such as water and methanol in
the crude glycerol favour the production hydrogen from the
steam-reforming process. Impurities such as biodiesel cata-
(9)
( 0 )
CO + H2 ↔ H2 O + C ΔHr = − 131 kJ/mol .
lyst and unreacted triglycerides contribute to the increase in
coke formation and catalyst deactivation (Veiga and Bussi
2016).
Composition of crude glycerol

The information in Table 2 (column 2) reveals that most


Process input and performance parameters
studies on the steam reforming of glycerol utilised pure and
analytical grade product for their study. Only three stud-
In steam-reforming studies, several parameters are used to
ies examined glycerol in its crude form (Slinn et al. 2008;
represent the nature of feed input as well as the performance
Veiga and Bussi 2016; Veiga et al. 2017). Adhikari et al.
of the process. A few of these parameters have been item-
(2007c) commented that pure glycerol was used in their
ised in other steam-reforming reviews (Adeniyi et al. 2019)
study instead of crude to avoid the complexity associated
albeit for other feedstock. In this section, we will present
with the interpretation of results. This can be an underlying
the steam-reforming indices as it applies to studies involv-
reason why most other studies followed suit. Typical compo-
ing glycerol feedstock. Key input parameters for the steam
sition of crude glycerol obtained from biodiesel production
reforming of oxygenated compounds are the steam-to-car-
is presented in Table 1. These results are only restricted to
bon (S/C) ratio, temperature, pressure, and catalyst loading:
studies where the glycerol is considered as feedstock for the
steam-reforming process. S/C =
Amount of steam
. (10)
In the study by Slinn et al. (2008), it was observed that Total amount of carbon in the feed (wet basis)
the selectivity of the chemical species in the product stream
The steam-to-carbon ratio (S/C) in mass or moles can
does not change much for pure versus crude glycerol given
be defined as the ratio of the amount of steam to the total
the same set of process conditions. However, the overall
amount of carbon in the feedstock (including the water con-
product yield from the process was observed to be 30%
tent). Several studies on glycerol steam reforming prefer a
lower when crude glycerol is used. The crude glycerol
similar term known as the water–gas feed ratio (WGFR or
composition (see Table 1) reported by Slinn et al. (2008)
WGR) (Adeniyi and Ighalo 2018; Adhikari et al. 2007b).
only contains a 55% alcohol content. If we consider this
The difference is the change in the organic basis from the

13
Chemical Papers (2019) 73:2619–2635 2623

Table 2  Summary of process conditions for glycerol steam reforming


References Nature of glyc- Process parameters Catalysts
erol feed
Temperature Pressure Space time/vel Water–glycerol
molar ratio

Hirai et al. (2005) Analytical 500–600 °C 1 atm 80 L/g h S/C of 3 Ru/Y2O3


grade
Adhikari et al. Pure 327–727 °C 1–5 atm – 1–9 Ni/MgO
(2007b)
Zhang et al. (2007) Pure 250–600 °C 1 atm 11 L/g h 9 Ir/CeO2, Co/CeO2, Ni/CeO2
Adhikari et al. Pure 600–900 °C – – 3–9 Al2O3, Rh/Al2O3, Pt/Al2O3,
(2007c) Pd/Al2O3, Ir/Al2O3, Ru/
Al2O3, Ni/Al2O3, Ce/Al2O3,
Rh/Ce/Al2O3, Pt/Ce/Al2O3,
Pd/Ce/Al2O3, Ir/Ce/Al2O3,
Ru/Ce/Al2O3, Ni/Ce/Al2O3
Slinn et al. (2008) Crude and Ana- 600–900 °C – – S/C of 1.35 Pt/Al2O3
lytical grade
Adhikari et al. Pure 550–650 °C – – 6–12 Ni/MgO, Ni/CeO2, Ni/TiO2
(2008b)
Adhikari et al. Pure 550–650 °C – – 6 Ni/MgO, Ni/CeO2, Ni/TiO2
(2008a)
Dou et al. (2009) Analytical 400–700 °C 1 atm – S/C of 3 Ni/Al2O3
grade
Adhikari et al. Pure – – – – Ni/CeO2
(2009)
Pompeo et al. Pure 250–700 °C 1 atm 0.2–6 min – Pt/SiO2, Pt/γ–Al2O3, Pt/ZrO2,
(2010) Pt/Ce4Zr1α
Cheng et al. (2010) Pure 450–550 °C 1 atm 50 L/g h 30–60 wt% glycerol Co/Al2O3
in water (or S/C
of 1.1–4)
Chiodo et al. (2010) Pure 450–750 °C 1 atm 5–30 L/ml h S/C of 3 Ni/γ-Al2O3, Rh/γ-Al2O3, Ni/
MgO, Ni/CeO2
Dave and Pant Pure 600–700 °C 1 atm 0.96–1.92 kgcat h/ 10–20 wt% glycerol CeO2, ­ZrO2, ­ZrO2/CeO2, Ni/
(2011) Kmolfeed in water CeO2, Ni–ZrO2/CeO2
Nichele et al. Pure 500–650 °C 1 atm – 10 wt% glycerol in Ni/ZrO2, Ni/TiO2, Ni/SBA-15
(2012) water
Ciftci et al. (2014) Pure 130–400 °C (600 rpm, for 13,000/h 50 wt% glycerol in Pt/CeO2, Pt–Re/CeO2, Pt/Ce-
APR) water ZrO2, Pt–Re/Ce-ZrO2, Pt/
ZrO2, Pt–Re/ZrO2, Pt/TiO2,
Pt–Re/TiO2
Pairojpiriyakul Analytical 450–575 °C 25 Mpa 3.15–10.8/h 2.5–10 wt% glyc- Ni/α–Al2O3, Ni/γ–Al2O3, Ni/
et al. (2014) grade erol in water La2O3, Ni/ZrO2, Ni/YSZ
De Rezende et al. Pure 600 °C 1 atm – 9 PtMgxAl-H, PtMgxAl-O
(2015)
Gallegos-Suárez Pure 200 °C 1 atm – 10 wt% glycerol in Ni/CeO2
et al. (2015) water
Veiga and Bussi Crude 650 °C 1 atm – S/C of 3.7 Ni–Mg/AC, Ni–Y/AC,
(2016) Ni–La/AC (AC activated
carbon)
Suffredini et al. Pure 600 °C – – 12 Ni/Al2O3, ­NiAl2O4, ­NiAl2O4/
(2016) Al2O3, ­NiAl2O4 + Al2O3
Pastor-Pérez and Pure 350 °C 1 atm – 10–30 wt% glycerol Pt–Sn/C
Sepúlveda-Escrib- in water
ano (2017)
Ni et al. (2017) Pure 550 °C – 2.28/h S/C of 3 NiO/NiAl2O4, NiO/ZrO2
Bepari et al. (2017) Pure 450–550 °C 1 atm 5.62–10.84 kgcat h/ 8–14 Ni/Fly-ash
Kmolfeed

13

2624 Chemical Papers (2019) 73:2619–2635

Table 2  (continued)
References Nature of glyc- Process parameters Catalysts
erol feed
Temperature Pressure Space time/vel Water–glycerol
molar ratio

Buffoni et al. Pure 450 °C 1 atm 14.7/h S/C of 1.6–15 Pt/SiO2-C


(2017)
Charisiou et al. Analytical 400–750 °C 1 atm 50 L/g h 31 v.v.  % glycerol Ni/Al2O3, Ni/La2O3-Al2O3
(2017) grade
Kousi et al. (2017) Pure 400–600 °C 1 atm 0.45 mg/ml min 20 Ru/Al2O3, Ru/B2O3–Al2O3,
Ru/MgO–Al2O3
Kumar et al. (2017) Pure – – – 5 vol% glycerol in TiO2
water
(Veiga et al. 2017) Crude 500–650 °C 1 atm 20 L/g h S/C of 3.7 Ni–La–Ti
Callison et al. Pure 240 °C (1000 rpm, for – 10 wt% glycerol in Pt/Al2O3
(2018) APR) water
Jiang et al. (2018) Pure 650 °C – 250/h S/C of 3 Ni/ZrO2, Yb–Ni/ZrO2, Pr–Ni/
ZrO2, Ce–Ni/ZrO2, La–Ni/
ZrO2
Karakoc et al. Pure 650 °C 1 atm – 6–15 Ni/CeO2, Ni/Al2O3, Ni/SiO2
(2018)
Seadira et al. Analytical – – – 5 vol% glycerol in TiO2, Ni/TiO2, Cu/TiO2, Co/
(2018) grade water TiO2, Ag/TiO2, Cr/TiO2

carbon content of the glycerol to the entire moles of the Selectivity is always based on the product, while yield is
glycerol itself. This is reasonably easier to compute. It is based on the input. For more specific computations apply-
expressed in Eq. 11 in terms of moles and can be referred to ing to glycerol steam reforming, we apply the following
as water–gas molar ratio (WGMR) (Adhikari et al. 2008b): equations:
Moles of water Moles of H2 produced
WGR =
Moles of glycerol in the feed
. (11) % H2 selectivity =
C atoms produced in the gas phase
1
Temperature increase favours hydrogen selectivity until × × 100%, (14)
RR
a threshold is exceeded. The best pressure condition for the
process is ambient pressure. Higher steam-to-carbon ratio where RR is H ­ 2/CO2 reforming ratio and it is 7/3 in the case
usually increases the hydrogen selectivity too. The fac- of glycerol steam-reforming process:
tors have different ways of interacting as regards the other Moles of H2 produced
chemical species. Detailed factor-interaction studies for each %H2 yield = × 100%, (15)
Maximum moles of H2 (= 7)
chemical species have been conducted (Adeniyi and Ighalo
2018; Adhikari et al. 2007a, b; Wang et al. 2008).
Glycerol in − Glycerol out
The general performance parameters for steam reforming Glycerol conversion = × 100%.
Glycerol in
are hydrogen gas selectivity or hydrogen gas yield. It is the
(16)
percentage (%) molar or mass composition of hydrogen in
the product stream: In some studies, percentage glycerol conversion into gaseous
products was used. This was elucidated as Eq. 19:
Amount of X produced
%X selectivity = × 100%, Gly conv to gaseous products
Total amount of syn-gas (dry basis)
(12) C stoms in gaseous product (17)
= × 100%.
C atoms produced in the gas phase
Amount of X produced
%X Yield = × 100%, (13) The terms explained in this section are all widely used
Total amount of feed
in steam-reforming studies in general and glycerol steam
where X can be hydrogen ( H2 ), carbon monoxide (CO), car- reforming in particular. Even in this review, these param-
bon dioxide ( CO2 ), methane ( CH4 ), or any other chemical eters are referred to in virtually all sections. A proper reader
species present in the product stream. The ‘amount’ can be understanding of these parameters is foundational in obtain-
in moles or grams depending on the basis of calculations. ing a proper appreciation of the current work.

13
Chemical Papers (2019) 73:2619–2635 2625

Process conditions of combinations for use in glycerol steam reforming. The


catalysts studied were A­ l2O3, Rh/Al2O3, Pt/Al2O3, Pd/Al2O3,
Research work on the steam reforming of crude glycerol has Ir/Al2O3, Ru/Al2O3, Ni/Al2O3, Ce/Al2O3, Rh/Ce/Al2O3, Pt/
a vast array of process conditions. These conditions include Ce/Al2O3, Pd/Ce/Al2O3, Ir/Ce/Al2O3, Ru/Ce/Al2O3, and Ni/
the type and nature of catalyst used, process temperature, Ce/Al2O3. The study revealed that Ni/Al2O3 has the best
pressure, steam-to-carbon ratio, space time or velocity, and selectivity for hydrogen compared to the other noble metals.
the presence of a ­CO2 sorbent. In this section, an exami- When the far cheaper cost of Nickel is considered in relation
nation of these conditions and the way which they affect to the noble metals, the extensive interest in nickel-based
the steam-reforming process are presented. Table 2 gives catalyst observed in Table 2 is seemingly justified. Based
a summary of the domain of process conditions examined on their premise, most future studies continued to pay close
in empirical studies and presented in a chronological order. attention to Nickel catalysts in the steam-reforming process.
Haven earlier established the superiority of Nickel (Adhi-
Catalyst kari et al. 2007c) over other metal catalysts, Adhikari et al.
(2008b) proceeded to study the performance of various metal
The catalysts for steam reforming are usually noble and oxides as catalyst support for the Nickel catalyst. MgO,
transition metals co-precipitated together or impregnated ­CeO2, and ­TiO2 were considered. The BET surface areas of
on metal oxides and ores. In most cases reported, incipient the developed catalyst were 67, 50.2, and 64.9 m2/g for Ni/
wetness technique (Adhikari et al. 2007c, 2008a) also known CeO2, Ni/MgO, and ­TiO2, respectively. The best catalyst of
as wet impregnation technique (Adhikari et al. 2008a; Cheng the three was shown to be Ni/CeO2 (Adhikari et al. 2008a,
et al. 2010) was the impregnation techniques used. There b). The maximum ­H2 selectivity was 74.7% with Ni/CeO2 at
have also been reports of the use of anionic and cationic a water–gas molar ratio (WGMR) of 12:1 and temperature of
exchange (Pompeo et al. 2010), micro-emulsion (Gallegos- 600 °C. The conversion of Glycerol was over 99% at these
Suárez et al. 2015), homogenous precipitation (Jiang et al. conditions. In a continuation of the investigations on Nickel,
2018), sol–gel (Buffoni et al. 2017), and the deposition–pre- Chiodo et al. (2010) studied Ni/γ–Al2O3, Rh/γ–Al2O3, Ni/
cipitation method (Zhang et al. 2007) for catalyst prepara- MgO, and Ni/CeO2 for glycerol steam reforming. They
tion. The support is also as important as the metal catalyst observed that Rh/γ–Al2O3 is more stable than the Ni cata-
itself as it exercises a certain level of catalytic activity espe- lysts and coke formation is lesser when the catalyst is used.
cially at lower reforming temperatures (Kousi et al. 2017). Dave and Pant (2011) compared C ­ eO2, ­ZrO2, ­ZrO2/CeO2,
The performance of the catalyst with respect to catalyst Ni/CeO2, and Ni–ZrO2/CeO2 for glycerol steam reform-
deactivation and sintering is also affected by the nature of ing. In the domain of the comparison, Ni–ZrO2/CeO2 was
the support. Adhikari et al. (2007c) was able to show that the observed to be the best in terms of glycerol conversion and
type of catalyst used will affect both the hydrogen electiv- hydrogen yield. In recent times, a plethora of studies have
ity in the product and the over conversion of the feed. This, also studied different catalyst support for Nickel catalyst (see
of course, is one of the fundamental reasons for catalysis. Table 2). In light of these, it can be surmised that nickel is
From Table 2, we can observe that the most studied metal the best metal catalyst for the steam reforming of glycerol.
catalyst for the steam reforming of glycerol is Nickel. The Quite recently, Seadira et al. (2018) considered several metal
study by Adhikari et al. (2007c) was quite foundational to catalysts on T
­ iO2 support for the photocatalytic reforming of
this research trend. This will be examined in the foregoing glycerol. They evaluated ­TiO2, Ni/TiO2, Cu/TiO2, Co/TiO2,
paragraphs. Ag/TiO2, and Cr/TiO2. The ­TiO2 support was in the form of
Adhikari et al. (2007b) utilised Ni/MgO as a catalyst for nano-structured hollow spheres. The researchers made this
the process. The performance of the catalyst in the domain choice, because, in that form, incorporating metals reduces
of hydrogen production was poorer than thermodynamic the band energy gap of the material for visible light harvest-
predictions. Zhang et al. (2007) tested Cobalt, Nickel, and ing and slows down the rate of electron/hole recombina-
Iridium in a ceria oxide matrix support as catalyst for glyc- tion, thereby increasing the photocatalytic activity. Cu/TiO2
erol steam reforming. Though the performance of all three showed highest activity for hydrogen production from the
catalysts was considered excellent for hydrogen production, process.
the Ir/CeO2 was the best of the three. This was ascribed to
the better and more intimate contact between Iridium parti- Temperature
cles and Ceria based on the Ceria-mediated redox process.
The study by Adhikari et al. (2007c) was quite exhaustive The temperatures studied are in wide variations. This is
in the domain of catalysis. They studied the performance because studies focus on different aspects of the process,
of seven different metals on alumina support in a variety such as catalyst performance (selectivity and activity),
carbon deposition or process optimisation, and/or product

13

2626 Chemical Papers (2019) 73:2619–2635

selectivity. In a study of process factors, Adeniyi and Ighalo 2012) and molar ratio (Adhikari et al. 2007a, b, c, 2008a,
(2018) observed that hydrogen yield will increase with tem- b). However, some researchers still use steam-to-carbon
perature. And beyond an optimal temperature, the hydrogen ratio (S/C) (Chiodo et al. 2010; Dou et al. 2009; Hirai et al.
production gradually drops will temperature. This is as a 2005; Slinn et al. 2008) which is more popular in bio-oil and
result of other competing reaction such as high-temperature methane reforming papers. Adhikari et al. (2007c) showed
thermal cracking of oxygenates, reverse WGS, and some that the higher the S/C or STGR, the higher the conver-
of the coke deposition reactions. Temperature is the most sion and hydrogen yield from the system. This is because a
studied factor in steam-reforming systems, and its effect higher amount of steam in the system will invariably favour
on product yield and selectivity is already well understood. the water–gas shift reaction. Pairojpiriyakul et al. (2014)
Adhikari et al. (2007b) examine the effect of temperature expressed the S/C ratio in terms of glycerol concentration
on the selectivity of each of the constituents in the product in the feed. It should be noted that this parameter is the
stream albeit in a comparative study between experiments opposite of S/C ratio (which is a measure of water concen-
and thermodynamic results. Though the same trend as that tration in the feed). It was observed from their study that
of Adeniyi and Ighalo (2018) was observed for tempera- increasing the glycerol feed concentration reduced hydrogen
ture, the study was able to pinpoint that Ni/MgO catalyst is yield and increased methane yield albeit for supercritical
not effective enough as its hydrogen yield is comparatively water reforming. This also holds for the conventional steam
poorer than thermodynamic predictions. Slinn et al. (2008) reforming (Bepari et al. 2017).
observed that high yields and hydrogen selectivity can be
obtained at higher temperatures not only for pure glycerol Spacetime and space velocity
but for crude glycerol too. From their study, Chiodo et al.
(2010) concluded that it is more convenient to carry out The spacetime will always show a reverse relationship to any
steam reforming of glycerol at temperature not higher than space velocity effects. Higher velocities of the feed indicate
650 °C despite favourable thermodynamics due to coking lesser space time and vice versa. That being said, increas-
and thermal cracking of species. Encapsulating carbon is ing the contact time of feedstock with catalyst (decrease the
deposited on the surface of the catalysts at higher tempera- weight hourly space velocity or increasing the space time)
tures which subsequently reduces the catalyst stability. will lead to a higher yield of hydrogen. Dave and Pant (2011)
elucidated an increase in glycerol conversion and hydrogen
yield with increase in spacetime. The exact same trend was
Pressure
also observed by Pompeo et al. (2010) and Bepari et al.
(2017). Slinn et al. (2008) gave an interesting interpretation
Thermodynamic analysis has revealed that atmospheric pres-
of the effect of spacetime and velocity on the product selec-
sure is optimum for hydrogen production, and hence, most
tivity and yield. It was explained that the reactant molecule
studies stick with this (Adeniyi and Ighalo 2018; Wang et al.
will need to meet several requirements; an available active
2008; Yang et al. 2011). Adeniyi and Ighalo (2018) observed
site, contact with a water molecule for a certain period of
that optimal hydrogen production will be obtained at atmos-
time, and enough energy to overcome the activation energy
pheric pressure, while methane production was favoured at
barrier. If the flow rate per gram of catalyst is too fast, then
higher pressures. The CO and ­CO2 in the system were fairly
one or all of these conditions may not be met. Without these
insensitive to pressure changes. Chemical equilibrium of the
conditions, the reactant molecules may just be pyrolysing on
system will shift in favour of the lighter chemical species at
the surface of the catalyst support.
lower pressures and vice versa (Adeniyi et al. 2019). Since
hydrogen is the most desired chemical species in the product
Process type
stream (and the lightest too), it would only be practical to
keep the system pressure at atmospheric conditions. This
The nature of the reforming process will determine the
is also desirable, considering that it is cheaper to design
process conditions and ultimately the yield of hydro-
systems to operate at this condition. Due to this already well-
gen. Besides the conventional steam reforming of glyc-
established knowledge on steam reforming, pressure effects
erol, studies have examined the sorption-enhanced steam
is scarcely studied in empirical steam-reforming research. In
reforming (Dou et al. 2009), chemical looping reforming
silico studies still consider pressure though.
(Ni et al. 2017), aqueous phase reforming (Ciftci et al.
2014), supercritical steam reforming (Pairojpiriyakul et al.
Steam‑to‑carbon (S/C) ratio 2014), and photocatalytic reforming (Seadira et al. 2018).
Sorption-enhanced processes usually have very high purity
The steam-to-carbon ratio is sometimes as weight percent of hydrogen in the product stream as they are used when
(Cheng et  al. 2010; Dave and Pant 2011; Nichele et  al. the product is required for highly specialised applications

13
Chemical Papers (2019) 73:2619–2635 2627

like for fuel cells. Photocatalytic reforming, for example, Thermodynamic analyses
takes place at near ambient conditions under the influ-
ence of light. Supercritical steam reforming utilises water Before a review of the findings of thermodynamic analy-
at its supercritical state. Aqueous phase reforming is a ses is presented, a brief summary of the calculation tech-
high-pressure process and low-temperature process. Fur- nique utilised will first be elucidated. Thermodynamic
thermore, the use of either a fixed or fluidised bed will analysis of steam-reforming systems is by minimisation
invariably affect the catalyst performance, and, hence, the of Gibbs free energy method. The minimisation of Gibbs
product yield and quality. Table 3 gives a summary of free energy method has been explicitly elucidated in the
the different reactor and process types utilised in steam literature (Adeniyi et al. 2019; Ashraf and Kumar 2018). If
reforming of glycerol. It is quite conspicuous that most we keep the temperature and pressure of our system con-
studies prefer to utilise fixed bed reactors for their experi- stant, then the equilibrium of the system can be expressed
mental studies. For a majority of the reforming process as put in Eq. 19:
reported, very good yield of hydrogen gas was obtained.
Some researchers presented their results in terms of yield, K

percentage selectivity, non-percentage yield or selectivity, dG = 𝜇i ni dni . (19)
or a combination.
i=1

The objective is to find the set of ni values that will


Sorption enhancement minimize the value of G (Adhikari et al. 2007a, b). This
can be done by two approaches:
Sorption usually encompasses adding a heterogeneous
chemical species in the reacting system to preferentially 1. Stoichiometric approach.
combine with C ­ O 2 in situ and reduce its concentration 2. Non-stoichiometric approach.
in the product stream. In most cases, quicklime (CaO) is
used (Wang et al. 2018). This is possible, because it reacts In the first approach, the system is described by a set of
reversibly with ­CO 2 to form calcium tri-oxo-carbonate stoichiometrically independent reactions which are typically
­(CaCO3) according to the following stoichiometry: chosen arbitrarily from a set of possible reactions. The non-
stoichiometric approach involves finding the equilibrium
(18)
( 0 )
CaO + CO2 ↔ CaCO3 ΔHr = − 178 kJ/mol . composition by the direct minimization of the Gibbs free
energy for a given set of species. There are several advan-
In some systems, dolomite is used (Dou et al. 2009).
tages of non-stoichiometric approach over the former: a
Dolomite contains several metal oxides majorly MgO
selection of the possible set of reactions is not required,
and CaO. Once calcined to a very high temperature, the
divergence does not occur during computation, and an accu-
MgO is no longer able to carbonate; hence, only CaO
rate estimation of the initial equilibrium composition is not
remains as a sorbent. Other natural sorbents are huntite
necessary (Adhikari et al. 2007a, b). The non-stoichiometric
and hydrotalcite (Dou et al. 2014). Synthetic sorbents
approach is the more applied technique in the open literature:
include lithium orthosilicate ­(Li4SiO4), lithium zirconate
­( Li 2ZrO 3), and sodium zirconate ­( Na 2ZrO 3) (Dou et al. K
2014). Ni et al. (2017) considered potassium promoted

G= 𝜇i ni . (20)
lithium zirconate ­( Li 2ZrO 3) considered as a ­CO 2 sorb- i=1
ent in the chemical looping reforming of glycerol. The
study concluded that the sorbent was very effective in its To find the value of ni that will minimize the value of G,
intended purpose and the presence of Nickel oxides as then it is important that the value of ni be in mass balance:
oxygen transfer materials was of help in reducing coking K
and carbon deposition. In general, hydrogen from steam

ali ni = bl , l = 1, … , M. (21)
reforming is difficult to use in energy generation via fuel i=1
cells due to its high ­CO2 and CO content; therefore, sorp-
tion technology is applied to remove these chemical spe- The above expression can then be further expressed by
cies from the product stream and obtain hydrogen gas of the following equation:
high purity (Esteban-Díez et al. 2016; Gil et al. 2016; Ni K K K
et al. 2017; Xie et al. 2016).
∑ ∑ ∑
G= ni ΔG0i + RT ni Inyi + RT ni InP. (22)
i=1 i=1 i=1

13

2628 Chemical Papers (2019) 73:2619–2635

Table 3  Summary of hydrogen production levels obtained for different studies


References Reactor Optimal hydrogen yield or selectivity Process type

Hirai et al. (2005) Fixed bed About 90% maximum H ­ 2 yield Steam reforming
Adhikari et al. (2007b) – About 0.65 mol fraction of H ­ 2 (dry basis) Steam reforming
in the feed
Zhang et al. (2007) Fixed bed 94.1% maximum ­H2 selectivity (Ir/CeO2, Steam reforming
550 °C)
Adhikari et al. (2007c) Fixed bed 80% maximum ­H2 selectivity Steam reforming
Slinn et al. (2008) – About 62% maximum H ­ 2 selectivity Steam reforming
Adhikari et al. (2008b) Fixed bed 74.7% maximum ­H2 selectivity (Ni/CeO2, Steam reforming
600 °C)
Adhikari et al. (2008a) – 66.69% maximum ­H2 selectivity (Ni/CeO2, Steam reforming
550 °C), 56.51% maximum ­H2 yield (Ni/
MgO, 650 °C)
Dou et al. (2009) Fixed bed About 80% maximum H ­ 2 selectivity and Steam reforming (sorption enhanced)
maximum purity of 97% after sorption
Adhikari et al. (2009) Fixed bed – Steam reforming
Pompeo et al. (2010) Fixed bed About 65% maximum H ­ 2 selectivity Steam reforming
Cheng et al. (2010) Fixed bed 65% maximum ­H2 selectivity Steam reforming
Chiodo et al. (2010) Fixed bed 4.5 mol ­H2 per mol glycerol Steam reforming
Dave and Pant (2011) Fixed bed 82.6% maximum ­H2 selectivity Steam reforming
Dave and Pant (2011); Nichele et al. (2012) Fixed bed About 90% H ­ 2 yield (Ni/ZrO2, 5 h, 500 °C) Steam reforming
Ciftci et al. (2014) Fixed bed About 35% glycerol conversion Aqueous phase reforming
Pairojpiriyakul et al. (2014) – 3.8 mol of ­H2 per mol glycerol Supercritical steam reforming
De Rezende et al. (2015) Fixed bed 68.2% maximum ­H2 selectivity Steam reforming
Gallegos-Suárez et al. (2015) Fixed bed – Steam reforming
Veiga and Bussi (2016) Fixed bed Up to 90% ­H2 yield Steam reforming
Suffredini et al. (2016) Fixed bed 70% maximum ­H2 selectivity (600 °C, Steam reforming
­NiAl2O4/Al2O3)
Pastor-Pérez and Sepúlveda-Escribano Fixed bed 45% maximum ­H2 selectivity Low-temperature Steam reforming
(2017)
Ni et al. (2017) Fixed bed 87.9% maximum ­H2 selectivity (NiO/ Chemical looping reforming (sorption
NiAl2O4) enhanced)
Bepari et al. (2017) Fixed bed 5.8 mol of ­H2 per mol glycerol (550 °C) Steam reforming
Buffoni et al. (2017) Fixed bed 5.8 mol of ­H2 per mol glycerol Steam reforming
Charisiou et al. (2017) Fixed bed Up to 85% maximum ­H2 selectivity Steam reforming
(700 °C, Ni/LaAl) and 55% ­H2 yield
(750 °C, Ni/LaAl)
Kousi et al. (2017) Fixed bed 70% ­H2 yield Steam reforming
Kumar et al. (2017) Fixed bed – Photocatalytic reforming
Veiga et al. (2017) Fixed bed Up to 88% ­H2 yield Steam reforming
Callison et al. (2018) – 34% glycerol conversion and 5.5% H ­ 2 yield Aqueous phase reforming
Jiang et al. (2018) Fixed bed Up to 90% maximum ­H2 selectivity (Ce–Ni/ Chemical looping reforming
ZrO2)
Karakoc et al. (2018) Fixed bed 4.82 mol of ­H2 per mol glycerol (Ni/CeO2, Steam reforming
650 °C)
Seadira et al. (2018) – – Photocatalytic reforming

At high temperatures and low pressure, the system is con- Microsoft Excel has been used to solve the set of linear equa-
sidered to be ideal (Adhikari et al. 2007a, b). Equation 22 is tions by the solver function (Da Silva and Müller 2011).
known as the objective function. It can be solved on mathe- Process simulation softwares like ASPEN Plus (Adeniyi and
matical softwares like Mathcad (Adhikari et al. 2007a, b) and Ighalo 2018; Goicoechea et al. 2015; Montero et al. 2015;
Matlab (Wang et al. 2008) by Lagrange Multiplier method. Patcharavorachot et al. 2018; Vagia and Lemonidou 2007,

13
Chemical Papers (2019) 73:2619–2635 2629

2008; Yang et al. 2011) and ASPEN Hysys (Gallegos-Suárez


E [ ]n
( )
et al. 2015) also utilise this objective function in the mini- ri = k0 Exp − Ci . (23)
RT
misation of Gibbs free energy calculation method to obtain
thermodynamically accurate results. It should be noted that The given form of the rate model does not include the
Eq. 10 does not apply when carbon is incorporated in the water concentration, because water was present in excess
study, because the free energy of formation of solid carbon compared to the concentration of glycerol. In kinetic studies,
is zero and it has no vapour pressure (Adhikari et al. 2007b). it is sacrosanct that ideal reactor conditions be maintained.
The key information from thermodynamic analyses is the Mass transfer limitations can be minimized by selecting an
prediction of product composition at different levels of the appropriate particle-size range for the catalyst. Film diffu-
different input parameters. This can be used in comparison sion in the reaction system can be minimized by selecting
with experimental systems to determine efficiency and in suitable flow rates. The reactor flow conditions (plug or
optimisation and factor-interaction studies. mixed) must be carefully monitored to ensure that they are
Adhikari et al. (2007a) showed that high temperatures, according to standards (Levenspiel 1999; Perry and Don
low pressures, and high WGFRs favour the production 1999). The key information obtained from the kinetic mod-
of hydrogen. The study revealed that the optimum condi- elling is the activation energy and the reaction order for the
tions for hydrogen yield are at > 900 K temperature and 9:1 reforming reaction. Table 4 gives the results obtained by
molar ratio of water to glycerol. The formation of carbon is different researchers. The results of Adhikari et al. (2009)
thermodynamically inhibited at these conditions. The find- showed that the activation energy and the reaction order for
ings of this study were quite similar to those of Wang et al. the glycerol steam-reforming reaction over Ni/CeO2 cata-
(2008). Adeniyi and Ighalo (2018) combined the thermo- lyst were found to be 103.4 kJ/mol and 0.233, respectively,
dynamic analyses with response surface optimisation. The based on a power model. These results were estimated with
study showed a 900 °C optimum temperature, 15.75 mol/mol nonlinear regression analysis using the Marquardt algorithm
steam-to-glycerol ratio (STGR), and 1 atm as the optimum using SAS 9.1 software and proposed to be in line with those
conditions. At these conditions, hydrogen yield was 66.72% of other researchers. Dave and Pant (2011) utilised curve-
and there was no methanation in the system. fitting tool box from MATLAB for their study and obtained
The thermodynamic analysis of Ashraf and Kumar (2018) an activation energy with the reaction order of 43.4 kJ/mol
was for a hydrazine-assisted reforming process. Their stud- and 0.3, respectively, based on a power model.
ies showed that reforming with increased hydrazine in the
system did not lead to zero coke formation or incomplete
conversion and an increase in hydrogen production was also Exergy and energy analyses
observed. The analysis by Chen et al. (2009) involved the
enhancement of the process by the sorption of carbon diox- Studies have been conducted on the energy and exergy anal-
ide. Apart from the improvement of the product purity, it yses of steam reforming (Hajjaji et al. 2014; Rabbani and
was observed that the sorption enhancement also increases Dincer 2015). Hajjaji et al. (2014) investigated the chemi-
the glycerol conversion to hydrogen. The study by Yang cal reactions, design, and simulation of the steam-reforming
et al. (2011) was for an oxidative steam-reforming process. It process on ASPEN Plus, and used it to study the energetic
was revealed that the process preferred moderate S/C ratios and exergetic performances and perform parametric analyses
to compensate between hydrogen yield and energy con- (using intuitive and design of experiment-based methods).
sumption. The comparison of glycerol steam reforming and In studying and optimizing the chemical reaction of steam
ethanol SR showed that ethanol will give a higher hydrogen glycerol reforming and its associated input parameters, their
yield under the same basis because of the more oxygenated study revealed the optimal conditions for maximizing hydro-
composition of glycerol. gen production as a temperature and a water-to-glycerol feed
ratio (WGFR) of 677 °C and 9, respectively. The thermal and
exergetic efficiencies of the resulting process are 66.6% and
Kinetic modelling 59.9%, respectively. These findings by Hajjaji et al. (2014)
were lower than those of other reformates (methane, etha-
The kinetics of glycerol steam-reforming reactions have nol, and methanol) but higher butanol (Da Silva and Mül-
been studied (Adhikari et al. 2009; Bepari et al. 2017; Dave ler 2011). The parametric investigation indicates that the
and Pant 2011; Yang et al. 2018). Both the power law model performance of the process (energetic and exergetic) could
and the Langmuir–Hinshelwood–Howgen–Watson (LHHW) be increased using a judiciously selected combination of
model has been employed to study the kinetics of glycerol the reactor conditions. Based on the investigations, WGFR
steam reforming, though the former is more used. The power of 6 and 827 °C appear to be the exergetic optimal for the
law model is in the following form: entire glycerol-to-hydrogen process. For this recommend

13

2630 Chemical Papers (2019) 73:2619–2635

Table 4  Kinetic parameters for References Model Catalyst Activation energy Reaction order
the catalytic steam reforming of (kJ/mol)
glycerol
Adhikari et al. (2009) Power law Ni/CeO2 103.4 0.233
Cheng et al. (2010) Power law Co/Al2O3 67.20 –
Dave and Pant (2011) Power law Ni-ZrO2/CeO2 43.4 0.3
Bepari et al. (2017) LHHW Ni/Fly-ash 29.0 –
Bepari et al. (2017) Power law Ni/Fly-ash 30.0 –

configuration, the thermal and exergetic efficiencies are The economic analysis underlines also the slightly better per-
78.1% and 66.1%, respectively. formance of conventional steam reforming than autothermal
Rabbani and Dincer (2015) studied a glycerol steam- reforming in terms of specific capital investment costs, opera-
reforming system coupled with a combined power plant for tional and maintenance costs, and hydrogen and power pro-
co-generation of hydrogen and electricity. Parametric and duction costs. The slightly higher carbon capture rate of auto-
optimisation studies were conducted to investigate the effects thermal reforming than the conventional reforming is reflected
of varying operating conditions and state properties on energy ­ O2 capture cost. Hydrogen and power co-generation
in lower C
and exergy efficiencies of the system. Their results showed is surmised as a promising option to further increase the over-
that an increasing steam-to-glycerol ratio increases the hydro- all energy efficiency and to improve the plant flexibility.
gen production, but, due to an increase in energy input, it
decreases the efficiency of the overall system. Increasing the
combustion temperature also increases the efficiency, while an Computer simulations
increasing reformer pressure reduces the hydrogen production
and decreases the overall efficiency of the system. Process simulation studies involving glycerol steam reform-
ing have been reported and most are in the domain of com-
putational fluid dynamics (Dou et al. 2008; Dou and Song
Techno‑economic performance 2010; Wang et al. 2018), energy and exergy analyses (Hajjaji
et al. 2014), thermodynamic analysis (Gallegos-Suárez et al.
Cormos and Cormos (2017) evaluated from the conceptual 2015; Patcharavorachot et al. 2018; Yang et al. 2011), process
design, thermal integration, and techno-economic and envi- comparisons (Patcharavorachot et al. 2018), factor-interaction
ronmental performances point-of-view hydrogen and power studies (Adeniyi and Ighalo 2018), and sorption enhancement
generation using glycerol (as a biodiesel by-product) on a (Wang et al. 2018). Dou et al. (2008) prepared a computa-
large scale with and without carbon capture. Conventional tional fluid dynamics simulation of gas–solid flow in a fluid-
steam reforming and oxygen autothermal reforming were the ized bed reactor to investigate the steam reforming of glycerol
processes considered. The evaluated hydrogen plant was to using a three-step reaction scheme. The Eulerian–Eulerian
conceptually produce 100,000 Nm3/h hydrogen. The power two-fluid approach was adopted to simulate hydrodynam-
plant was to conceptually generate about 500 MW net power ics of fluidization, and chemical reactions were modelled by
output. Similar designs without carbon capture were devel- the laminar finite-rate model. Their results revealed that the
oped to quantify the energy and cost penalties for carbon gas–solid system exhibited a more heterogeneous structure.
capture. The various glycerol-reforming cases were modelled Clusters were observed to fall and stack together along the
and simulated to produce the mass and energy balances for wall, and the process of wall slug formation was very evident
quantification of key plant performance indicators (e.g., fuel in the represented visualisations of gases and solids velocities
consumption, energy efficiency, ancillary energy consump- with vector plots. Figure 4a and b gives a better representa-
tion, specific ­CO2 emissions, capital and operational costs, tion of the kinetic findings of Dou et al. (2008) in the regards
production costs, cash flow analysis, etc.). The designs were of both solids and gases. The heterogeneity of the systems
energy optimized by heat and energy integration analysis. is quite apparent. They also observed that the solid and gas
The evaluations show that glycerol reforming is promising velocities in the core region are much higher than those in the
concept for high energy efficiency processes with low ­CO2 annulus region, whereas solid and gas velocities near the wall
emissions. The technical and environmental results showed are decreasing and downward. This observation is an expla-
that the conventional steam reforming is slightly more energy nation for any potential back-mixing and internal circulation
efficient than the autothermal reforming with or without car- behaviour in the system.
bon capture. On the other hand, the autothermal reforming has Dou and Song (2010) also prepared a simulation to
higher carbon capture rates than the conventional reforming. study the relationship of hydrodynamics with hydrogen

13
Chemical Papers (2019) 73:2619–2635 2631

production. They used a similar fluidisation and reaction transesterification process itself as most researchers prefer
model to that of Dou et al. (2008). Their results showed to use pure glycerol. Beyond the fact that an ease of inter-
that the core-annulus structure of gas–solid flow led to pretation is afforded, the impurities in the crude glycerol
back-mixing and internal circulation behaviour, and, thus, possess a significant effect on the process (in a diversity of
resulted in a poor velocity distribution. The conversion ways) which, at this current time, has not been extensively
of glycerol and the subsequent production of hydrogen explored. Most empirical studies have been done on a labo-
decreased with increasing inlet gas velocity. H­ 2 concen- ratory scale. Pilot studies have not been examined in this
trations in the bed were shown to uneven and increased regard. The large-scale evaluations have only been done on
downstream with higher concentrations found on walls. in silico platforms. CFD simulations have predicted possible
Their model effectively demonstrated a relationship internal circulations and back-mixing in pilot systems. The
between hydrodynamics and hydrogen production. The dynamics of mass and heat transfer in the system will be a
computer simulations on glycerol steam reforming have very interesting study. Design, budgeting, cost, and energy
helped to reveal helpful data needed to design and operate considerations can be very different and more realistic in
a bench-scale catalytic fluidized bed reactor. pilot studies. This is also a great perspective for future
research.
For biomass bio-oil, for example, Adeniyi et al. (2019)
observed that most researchers have now moved focus to
Knowledge gaps and future perspectives the use of fluidised bed in steam reforming, especially in
recent times. This has not yet been observed in glycerol-
Based on the review of glycerol steam-reforming research, reforming studies. Majorly, fixed bed reactors have been
several knowledge gaps have been identified which can utilised and there are little reports on fluidised bed reac-
open up more opportunities for future perspectives in those tors. Due to a more rigorous and effective contacting, the
areas. Very few studies have utilised crude glycerol from the surface area of the catalyst is harnessed to the fullest. It

Fig. 4  Solids (a) and gases; (b)


velocity vector plots; velocity
coloured by magnitude, m/s
(Dou et al. 2008)

13

2632 Chemical Papers (2019) 73:2619–2635

would be interesting to study how crude glycerol will per- References


form under fluidised catalyst conditions and in what ways
the impurities will affect the catalyst. There is a lack of in- Abdullah T, Amran T, Nabgan W, Kamaruddin MJ, Mat R, Johari A,
depth stand-alone studies on carbon deposition behaviour in Ahmad A (2014) Hydrogen production from acetic acid steam
reforming over bimetallic Ni–Co on L ­ a2O3 catalyst-effect of the
glycerol steam reforming as regards the numerous catalysts catalyst dilution. In: Paper presented at the Applied Mechanics
already employed. Sorption enhancements involving some and Materials
more novel and efficient artificial sorbents have not been Adeniyi AG, Ighalo JO (2018) Study of process factor effects and
examined for glycerol reforming. This is another excellent interactions in synthesis gas production via a simulated model
for glycerol steam reforming. Chem Product Process Model.
perspective for future studies. Some other novel-reforming https​://doi.org/10.1515/cppm-2018-0034
techniques such as partial oxidation and autothermal reform- Adeniyi AG, Ighalo JO (2019) Hydrogen production by the steam
ing have been scarcely examined (at least within the domain reforming of waste lubricating oil. Indian Chem Engineer. https​
of this review). Besides the recent techno-economic analysis ://doi.org/10.1080/00194​506.2019.16058​47
Adeniyi AG, Ighalo JO, Eletta AAO (2018) Process integration
by Cormos and Cormos (2017), few researchers have delved and feedstock optimisation of a two-step biodiesel production
into the cost and profitability studies of hydrogen production process from Jatropha curcas using aspen plus. Chem Product
from crude glycerol as it compares to other feedstock. Process Model. https​://doi.org/10.1515/cppm-2018-0055
Adeniyi AG, Otoikhian KS, Ighalo JO (2019) Steam reforming of bio-
mass pyrolysis oil: a review. Int J Chem Reactor Eng. https​://doi.
org/10.1515/ijcre​-2018-0328
Conclusion Adhikari S, Fernando S, Gwaltney SR, To SF, Bricka RM, Steele
PH, Haryanto A (2007a) A thermodynamic analysis of hydro-
This paper reviewed and catalogued the current state of gen production by steam reforming of glycerol. Int J Hydro-
gen Energy 32(14):2875–2880. https​://doi.org/10.1016/j.ijhyd​
knowledge on the production of bio-hydrogen by the steam ene.2007.03.023
reforming of glycerol obtained from transesterification. Adhikari S, Fernando S, Haryanto A (2007b) A comparative thermody-
Key domains of research over the years explored process namic and experimental analysis on hydrogen production by steam
optimisation, catalysis, thermodynamics analysis, kinetic reforming of glycerin. Energy Fuels 21(4):2306–2310. https:​ //doi.
org/10.1021/ef070​035l
modelling, exergy and energy analyses, and computational Adhikari S, Fernando S, Haryanto A (2007c) Production of hydrogen
fluid dynamics (CFD) simulations. From the review, it was by steam reforming of glycerin over alumina-supported metal cat-
observed that the selectivity of the chemical species in alysts. Catal Today 129(3–4):355–364. https​://doi.org/10.1016/j.
the product stream does not change much for pure versus catto​d.2006.09.038
Adhikari S, Fernando SD, Haryanto A (2008a) Hydrogen produc-
crude glycerol reforming given the same set of process tion from glycerin by steam reforming over nickel catalysts.
conditions. However, the overall product yield from the Renew Energy 33(5):1097–1100. https​://doi.org/10.1016/j.renen​
process is lower when crude glycerol is used. Impurities e.2007.09.005
such as water and methanol in the crude glycerol favour Adhikari S, Fernando SD, To SF, Bricka RM, Steele PH, Haryanto A
(2008b) Conversion of glycerol to hydrogen via a steam reform-
the production hydrogen from the steam-reforming pro- ing process over nickel catalysts. Energy Fuels 22(2):1220–1226.
cess. Impurities such as biodiesel catalyst and unreacted https​://doi.org/10.1021/ef700​520f
triglycerides contribute to the increase in coke forma- Adhikari S, Fernando SD, Haryanto A (2009) Kinetics and reac-
tion and catalyst deactivation. Nickel is probably the best tor modeling of hydrogen production from glycerol via steam
reforming process over Ni/CeO2 catalysts. Chem Eng Technol
catalyst for the steam reforming of glycerol based on the 32(4):541–547. https​://doi.org/10.1002/ceat.20080​0462
agreement of many studies. Temperature optimal is always Ashraf J, Kumar A (2018) Thermodynamic evaluation of hydrazine
somewhere around 650–900 °C, while pressure is always assisted glycerol reforming for syngas production and coke inhi-
kept at atmospheric condition. 827 °C is considered as the bition. Int J Hydrogen Energy. https​://doi.org/10.1016/j.ijhyd​
ene.2018.05.074
exergetic optimal of the entire process, while 677 °C is Atabani AE, Silitonga AS, Ong HC, Mahlia TMI, Masjuki HH,
the exergetic optimal for hydrogen selectivity alone. Gap- Badruddin IA, Fayaz H (2013) Non-edible vegetable oils: a criti-
ing gaps in knowledge abound in the research area and cal evaluation of oil extraction, fatty acid compositions, biodiesel
these are in the domain of crude glycerol use, fluidised production, characteristics, engine performance and emissions
production. Renew Sustain Energy Rev 18:211–245. https​://doi.
bed use, artificial sorbent use, pilot scale studies, cost and org/10.1016/j.rser.2012.10.013
profitability analysis, and studies on carbon deposition Averill D, Nabea W, Piard C (2010) A comparison of biodiesel pro-
behaviour. It can be surmised that glycerol is a very good cesses for the conversion of Jatropha curcas. (B.Eng), Worcester
feedstock for the steam-reforming process and a good yield Polytechnic Institute
Bagnato G, Iulianelli A, Sanna A, Basile A (2017) Glycerol production
of hydrogen can be obtained. In view of the impending and transformation: a critical review with particular emphasis on
global energy crisis, commercialising this technology will glycerol reforming reaction for producing hydrogen in conven-
be invaluable for energy sustainability in years to come. tional and membrane reactors. Membranes 7(2):17. https​://doi.
org/10.3390/membr​anes7​02001​7

13
Chemical Papers (2019) 73:2619–2635 2633

Behr A, Eilting J, Irawadi K, Leschinski J, Lindner F (2008) Improved J Hydrogen Energy 36(3):2057–2075. https​://doi.org/10.1016/j.
utilisation of renewable resources: new important derivatives ijhyd​ene.2010.11.051
of glycerol. Green Chem 10(1):13–30. https​://doi.org/10.1039/ Da Silva GP, Mack M, Contiero J (2009) Glycerol: a promising
B7105​61D and abundant carbon source for industrial microbiology. Bio-
Bepari S, Pradhan NC, Dalai AK (2017) Selective production of hydro- technol Adv 27(1):30–39. https​://doi.org/10.1016/j.biote​chadv​
gen by steam reforming of glycerol over Ni/Fly ash catalyst. Catal .2008.07.006
Today 291:36–46. https​://doi.org/10.1016/j.catto​d.2017.01.015 Dasari MA, Kiatsimkul P-P, Sutterlin WR, Suppes GJ (2005) Low-
Blanco-Marigorta A, Suárez-Medina J, Vera-Castellano A (2013) Exer- pressure hydrogenolysis of glycerol to propylene glycol. Appl
getic analysis of a biodiesel production process from Jatropha Catal A 281(1–2):225–231. https​ : //doi.org/10.1016/j.apcat​
curcas. Appl Energy 101:218–225. https:​ //doi.org/10.1016/j.apene​ a.2004.11.033
rgy.2012.05.037 Dave CD, Pant K (2011) Renewable hydrogen generation by steam
Buffoni IN, Gatti MN, Santori GF, Pompeo F, Nichio NN (2017) reforming of glycerol over zirconia promoted ceria supported cat-
Hydrogen from glycerol steam reforming with a platinum alyst. Renew Energy 36(11):3195–3202. https:​ //doi.org/10.1016/j.
catalyst supported on a S ­ iO 2–C composite. Int J Hydrogen renen​e.2011.03.013
Energy 42(18):12967–12977. https​://doi.org/10.1016/j.ijhyd​ De Rezende SM, Franchini CA, Dieuzeide ML, de Farias AMD, Ama-
ene.2017.04.047 deo N, Fraga MA (2015) Glycerol steam reforming over layered
Callison J, Subramanian N, Rogers S, Chutia A, Gianolio D, Catlow double hydroxide-supported Pt catalysts. Chem Eng J 272:108–
CRA, Dimitratos N (2018) Directed aqueous-phase reforming of 118. https​://doi.org/10.1016/j.cej.2015.03.033
glycerol through tailored platinum nanoparticles. Appl Catal B Dou B, Song Y (2010) A CFD approach on simulation of hydrogen
238:618–628. https​://doi.org/10.1016/j.apcat​b.2018.07.008 production from steam reforming of glycerol in a fluidized bed
Chaminand J, aurent Djakovitch L, Gallezot P, Marion P, Pinel C, reactor. Int J Hydrogen Energy 35(19):10271–10284. https​://doi.
Rosier C (2004) Glycerol hydrogenolysis on heterogeneous org/10.1016/j.ijhyd​ene.2010.07.165
catalysts. Green Chem 6(8):359–361. https​://doi.org/10.1039/ Dou B, Dupont V, Williams PT (2008) Computational fluid dynamics
b4073​78a simulation of gas−solid flow during steam reforming of glycerol
Charisiou N, Siakavelas G, Papageridis K, Baklavaridis A, Tzounis in a fluidized bed reactor. Energy Fuels 22(6):4102–4108. https​
L, Polychronopoulou K, Goula M (2017) Hydrogen production ://doi.org/10.1021/ef800​2679
via the glycerol steam reforming reaction over nickel supported Dou B, Dupont V, Rickett G, Blakeman N, Williams PT, Chen H,
on alumina and lanthana-alumina catalysts. Int J Hydrogen Ghadiri M (2009) Hydrogen production by sorption-enhanced
Energy 42(18):13039–13060. https​://doi.org/10.1016/j.ijhyd​ steam reforming of glycerol. Biores Technol 100(14):3540–3547.
ene.2017.04.048 https​://doi.org/10.1016/j.biort​ech.2009.02.036
Chattanathan SA, Adhikari S, Abdoulmoumine N (2012) A review Dou B, Song Y, Wang C, Chen H, Xu Y (2014) Hydrogen production
on current status of hydrogen production from bio-oil. Renew from catalytic steam reforming of biodiesel byproduct glycerol:
Sustain Energy Rev 16(5):2366–2372. https​://doi.org/10.1016/j. issues and challenges. Renew Sustain Energy Rev 30:950–960.
rser.2012.01.051 https​://doi.org/10.1016/j.rser.2013.11.029
Chen H, Zhang T, Dou B, Dupont V, Williams P, Ghadiri M, Ding Y Esteban-Díez G, Gil MV, Pevida C, Chen D, Rubiera F (2016) Effect
(2009) Thermodynamic analyses of adsorption-enhanced steam of operating conditions on the sorption enhanced steam reform-
reforming of glycerol for hydrogen production. Int J Hydro- ing of blends of acetic acid and acetone as bio-oil model com-
gen Energy 34(17):7208–7222. https​://doi.org/10.1016/j.ijhyd​ pounds. Appl Energy 177:579–590. https​://doi.org/10.1016/j.
ene.2009.06.070 apene​rgy.2016.05.149
Chen G, Yao J, Liu J, Yan B, Shan R (2016) Biomass to hydrogen-rich Fu P, Yi W, Li Z, Bai X, Zhang A, Li Y, Li Z (2014) Investigation
syngas via catalytic steam reforming of bio-oil. Renewable Energy on hydrogen production by catalytic steam reforming of maize
91:315–322. https​://doi.org/10.1016/j.renen​e.2016.01.073 stalk fast pyrolysis bio-oil. Int J Hydrogen Energy 39(26):13962–
Cheng CK, Foo SY, Adesina AA (2010) ­H2-rich synthesis gas pro- 13971. https​://doi.org/10.1016/j.ijhyd​ene.2014.06.165
duction over Co/Al2O3 catalyst via glycerol steam reforming. Gallegos-Suárez E, Guerrero-Ruiz A, Fernández-García M, Rodríguez-
Catal Commun 12(4):292–298. https​://doi.org/10.1016/j.catco​ Ramos I, Kubacka A (2015) Efficient and stable Ni–Ce glycerol
m.2010.09.018 reforming catalysts: chemical imaging using X ray electron and
Chiodo V, Freni S, Galvagno A, Mondello N, Frusteri F (2010) Cat- scanning transmission microscopy. Appl Catal B 165:139–148.
alytic features of Rh and Ni supported catalysts in the steam https​://doi.org/10.1016/j.apcat​b.2014.10.007
reforming of glycerol to produce hydrogen. Appl Catal A 381(1– Gil MV, Fermoso J, Pevida C, Chen D, Rubiera F (2016) Production
2):1–7. https​://doi.org/10.1016/j.apcat​a.2010.03.039 of fuel-cell grade H ­ 2 by sorption enhanced steam reforming of
Ciftci A, Eren S, Ligthart DM, Hensen EJ (2014) Platinum-rhenium acetic acid as a model compound of biomass-derived bio-oil. Appl
synergy on reducible oxide supports in aqueous-phase glyc- Catal B 184:64–76. https​://doi.org/10.1016/j.apcat​b.2015.11.028
erol reforming. ChemCatChem 6(5):1260–1269. https​://doi. Goicoechea S, Ehrich H, Arias PL, Kockmann N (2015) Thermody-
org/10.1002/cctc.20130​1096 namic analysis of acetic acid steam reforming for hydrogen pro-
Cormos A-M, Cormos C-C (2017) Techno-economic and environmen- duction. J Power Sources 279:312–322. https​://doi.org/10.1016/j.
tal performances of glycerol reforming for hydrogen and power jpows​our.2015.01.012
production with low carbon dioxide emissions. Int J Hydro- Hajjaji N, Chahbani A, Khila Z, Pons M-N (2014) A comprehensive
gen Energy 42(12):7798–7810. https​://doi.org/10.1016/j.ijhyd​ energy–exergy-based assessment and parametric study of a hydro-
ene.2016.11.172 gen production process using steam glycerol reforming. Energy
Czernik S, French R, Feik C, Chornet E (2002) Hydrogen by catalytic 64:473–483. https​://doi.org/10.1016/j.energ​y.2013.10.023
steam reforming of liquid byproducts from biomass thermocon- Haryanto A, Fernando S, Murali N, Adhikari S (2005) Current status
version processes. Ind Eng Chem Res 41(17):4209–4215. https​:// of hydrogen production techniques by steam reforming of ethanol:
doi.org/10.1021/ie020​107q a review. Energy Fuels 19(5):2098–2106. https:​ //doi.org/10.1021/
Da Silva AL, Müller IL (2011) Hydrogen production by sorption ef050​0538
enhanced steam reforming of oxygenated hydrocarbons (ethanol,
glycerol, n-butanol and methanol): thermodynamic modelling. Int

13

2634 Chemical Papers (2019) 73:2619–2635

Hirai T, Ikenaga N-O, Miyake T, Suzuki T (2005) Production of hydro- Pagliaro M, Ciriminna R, Kimura H, Rossi M, Della Pina C (2007)
gen by steam reforming of glycerin on ruthenium catalyst. Energy From glycerol to value-added products. Angew Chem Int Ed
Fuels 19(4):1761–1762. https​://doi.org/10.1021/ef050​121q 46(24):4434–4440. https​://doi.org/10.1002/anie.20060​4694
Hoang TMC, Geerdink B, Sturm JM, Lefferts L, Seshan K (2015) Pairojpiriyakul T, Croiset E, Kiatkittipong K, Kiatkittipong W,
Steam reforming of acetic acid—a major component in the vola- Arpornwichanop A, Assabumrungrat S (2014) Catalytic reform-
tiles formed during gasification of humin. Appl Catal B Environ ing of glycerol in supercritical water with nickel-based cata-
163:74–82. https​://doi.org/10.1016/j.apcat​b.2014.07.046 lysts. Int J Hydrogen Energy 39(27):14739–14750. https​://doi.
Ito T, Nakashimada Y, Senba K, Matsui T, Nishio N (2005) Hydro- org/10.1016/j.ijhyd​ene.2014.07.079
gen and ethanol production from glycerol-containing wastes dis- Pambudi NA, Laukkanen T, Fogelholm C-J, Kohl T, Jarvinen M
charged after biodiesel manufacturing process. J Biosci Bioeng (2013) Prediction of gas composition of Jatropha curcas Linn
100(3):260–265. https​://doi.org/10.1263/jbb.100.260 oil cake in entrained flow reactors using ASPEN PLUS simu-
Iwasa N, Yamane T, Takei M, Ozaki J-I, Arai M (2010) Hydrogen lation software. Int J Sustain Eng 6(2):142–150. https​: //doi.
production by steam reforming of acetic acid: comparison of org/10.1080/19397​038.2012.65771​5
conventional supported metal catalysts and metal-incorporated Pastor-Pérez L, Sepúlveda-Escribano A (2017) Low temperature
mesoporous smectite-like catalysts. Int J Hydrogen Energy glycerol steam reforming on bimetallic PtSn/C catalysts: on
35(1):110–117. https​://doi.org/10.1016/j.ijhyd​ene.2009.10.053 the effect of the Sn content. Fuel 194:222–228. https​: //doi.
Jiang B, Li L, Bian Z, Li Z, Sun Y, Sun Z, Goula MA et al (2018) org/10.1016/j.fuel.2017.01.023
Chemical looping glycerol reforming for hydrogen production by Patcharavorachot Y, Chatrattanawet N, Arpornwichanop A, Ass-
Ni@ZrO2 nanocomposite oxygen carriers. Int J Hydrogen Energy abumrungrat S (2018) Optimization of hydrogen production
1:1. https​://doi.org/10.1016/j.ijhyd​ene.2018.05.065 from three reforming approaches of glycerol via using supercrit-
Johnson DT, Taconi KA (2007) The glycerin glut: options for the ical water with in situ C­ O2 separation. Int J Hydrogen Energy.
value-added conversion of crude glycerol resulting from bio- https​://doi.org/10.1016/j.ijhyd​ene.2018.07.053
diesel production. Environ Prog 26(4):338–348. https​: //doi. Patel M, Jindal TK, Pant KK (2013) Kinetic study of steam reforming
org/10.1002/ep.10225​ of ethanol on Ni-based ceria–zirconia catalyst. Ind Eng Chem Res
Karakoc OP, Kibar M, Akin A, Yildiz M (2018) Nickel-based cata- 52(45):15763–15771. https​://doi.org/10.1021/ie401​570s
lysts for hydrogen production by steam reforming of glycerol. Perry RH, Don GW (1999) Perry’s chemical engineers’ handbook, 7th
Int J Environ Sci Technol 2018:1–8. https​://doi.org/10.1007/ edn. McGraw-Hill companies Inc, New York
s1376​2-018-1875-8 Pompeo F, Santori G, Nichio NN (2010) Hydrogen and/or syngas from
Koh MY, Ghazi TIM (2011) A review of biodiesel production from steam reforming of glycerol. Study of platinum catalysts. Int J
Jatropha curcas L. oil. Renew Sustain Energy Rev 15(5):2240– Hydrogen Energy 35(17):8912–8920. https​://doi.org/10.1016/j.
2251. https​://doi.org/10.1016/j.rser.2011.02.013 ijhyd​ene.2010.06.011
Kousi K, Kondarides D, Verykios X, Papadopoulou C (2017) Profeti LP, Ticianelli EA, Assaf EM (2009) Production of hydrogen
Glycerol steam reforming over modified Ru/Al 2O 3 catalysts. via steam reforming of biofuels on Ni/CeO2–Al2O3 catalysts pro-
Appl Catal A 542:201–211. https​: //doi.org/10.1016/j.apcat​ moted by noble metals. Int J Hydrogen Energy 34(12):5049–5060.
a.2017.05.027 https​://doi.org/10.1016/j.ijhyd​ene.2009.03.050
Kumar DP, Kumari VD, Karthik M, Sathish M, Shankar M (2017) Rabbani M, Dincer I (2015) Energetic and exergetic assessments of
Shape dependence structural, optical and photocatalytic proper- glycerol steam reforming in a combined power plant for hydrogen
ties of TiO2 nanocrystals for enhanced hydrogen production via production. Int J Hydrogen Energy 40(34):11125–11132. https​://
glycerol reforming. Sol Energy Mater Sol Cells 163:113–119. doi.org/10.1016/j.ijhyd​ene.2015.04.012
https​://doi.org/10.1016/j.solma​t.2017.01.007 Remón J, Broust F, Volle G, García L, Arauzo J (2015) Hydrogen pro-
Levenspiel O (1999) Chemical reaction engineering, 3rd edn. Wiley, duction from pine and poplar bio-oils by catalytic steam reform-
Hoboken ing. Influence of the bio-oil composition on the process. Int J
Montero C, Oar-Arteta L, Remiro A, Arandia A, Bilbao J, Gayubo Hydrogen Energy 40(16):5593–5608. https​://doi.org/10.1016/j.
AG (2015) Thermodynamic comparison between bio-oil and ijhyd​ene.2015.02.117
ethanol steam reforming. Int J Hydrogen Energy 40(46):15963– Schwengber CA, Alves HJ, Schaffner RA, da Silva FA, Sequinel R,
15971. https​://doi.org/10.1016/j.ijhyd​ene.2015.09.125 Bach VR, Ferracin RJ (2016) Overview of glycerol reforming for
Nabgan W, Abdullah TAT, Mat R, Nabgan B, Gambo Y, Moghada- hydrogen production. Renew Sustain Energy Rev 58:259–266.
mian K (2016) Acetic acid-phenol steam reforming for hydro- https​://doi.org/10.1016/j.rser.2015.12.279
gen production: effect of different composition of L ­ a2O3–Al2O3 Seadira TW, Sadanandam G, Ntho T, Masuku CM, Scurrell MS (2018)
support for bimetallic Ni–Co catalyst. J Environ Chem Eng Preparation and characterization of metals supported on nano-
4(3):2765–2773. https​://doi.org/10.1016/j.jece.2016.05.030 structured ­TiO2 hollow spheres for production of hydrogen via
Ni Y, Wang C, Chen Y, Cai X, Dou B, Chen H, Wang K (2017) High photocatalytic reforming of glycerol. Appl Catal B 222:133–145.
purity hydrogen production from sorption enhanced chemical https​://doi.org/10.1016/j.apcat​b.2017.09.072
looping glycerol reforming: application of NiO-based oxygen Slinn M, Kendall K, Mallon C, Andrews J (2008) Steam reforming of
transfer materials and potassium promoted ­Li2ZrO3 as ­CO2 sorb- biodiesel by-product to make renewable hydrogen. Biores Technol
ent. Appl Therm Eng 124:454–465. https​://doi.org/10.1016/j. 99(13):5851–5858. https:​ //doi.org/10.1016/j.biorte​ ch.2007.10.003
applt​herma​leng.2017.06.00 Suffredini DF, Thyssen VV, de Almeida PM, Gomes RS, Borges MC,
Nichele V, Signoretto M, Menegazzo F, Gallo A, Dal Santo V, Cru- de Farias AMD, Brandão ST et  al (2016) Renewable hydro-
ciani G, Cerrato G (2012) Glycerol steam reforming for hydro- gen from glycerol reforming over nickel aluminate-based cata-
gen production: design of Ni supported catalysts. Appl Catal lysts. Catal Today 289:96–104. https​://doi.org/10.1016/j.catto​
B 111:225–232. https​://doi.org/10.1016/j.apcat​b.2011.10.003 d.2016.07.027
Ofari-Boateng C, Keat TL, JitKang L (2012) Sustainability assess- Thaicharoensutcharittham S, Meeyoo V, Kitiyanan B, Rangsunvigit P,
ment of microalgal biodiesel production processes: an exergetic Rirksomboon T (2011) Hydrogen production by steam reforming
analysis approach with Aspen Plus. Int J Exergy 10(4):400–416. of acetic acid over Ni-based catalysts. Catal Today 164(1):257–
https​://doi.org/10.1504/IJEX.2012.04751​0 261. https​://doi.org/10.1016/j.catto​d.2010.10.054

13
Chemical Papers (2019) 73:2619–2635 2635

Thompson JC, He BB (2006) Characterization of crude glycerol from Wang S, Yang X, Xu S, Zhang K, Li B (2018) Assessment of sorption-
biodiesel production from multiple feedstocks. Appl Eng Agric enhanced crude glycerol steam reforming process via CFD simu-
22(2):261–265. https​://doi.org/10.13031​/2013.20272​ lation. Int J Hydrogen Energy 43(32):14996–15004. https​://doi.
Tiwari AK, Kumar A, Raheman H (2007) Biodiesel production from org/10.1016/j.ijhyd​ene.2018.06.053
jatropha oil (Jatropha curcas) with high free fatty acids: an opti- Xie H, Yu Q, Zuo Z, Han Z, Yao X, Qin Q (2016) Hydrogen production
mized process. Biomass Bioenerg 31(8):569–575. https​://doi. via sorption-enhanced catalytic steam reforming of bio-oil. Int
org/10.1016/j.biomb​ioe.2007.03.003 J Hydrogen Energy 41(4):2345–2353. https​://doi.org/10.1016/j.
Vagia EC, Lemonidou AA (2007) Thermodynamic analysis of hydro- ijhyd​ene.2015.12.156
gen production via steam reforming of selected components of Yang G, Yu H, Peng F, Wang H, Yang J, Xie D (2011) Thermody-
aqueous bio-oil fraction. Int J Hydrogen Energy 32(2):212–223. namic analysis of hydrogen generation via oxidative steam reform-
https​://doi.org/10.1016/j.ijhyd​ene.2006.08.021 ing of glycerol. Renew Energy 36(8):2120–2127. https​://doi.
Vagia EC, Lemonidou AA (2008) Thermodynamic analysis of org/10.1016/j.renen​e.2011.01.022
hydrogen production via autothermal steam reforming of Yang F, Hanna MA, Sun R (2012) Value-added uses for crude glyc-
selected components of aqueous bio-oil fraction. Int J Hydro- erol–a byproduct of biodiesel production. Biotechnol Biofuels
gen Energy 33(10):2489–2500. https​://doi.org/10.1016/j.ijhyd​ 5(1):13. https​://doi.org/10.1186/1754-6834-5-13
ene.2008.02.057 Yang X, Wang S, Li Z, Zhang K, Li B (2018) Enhancement of mem-
Vaidya PD, Rodrigues AE (2009) Glycerol reforming for hydrogen brane hydrogen separation on glycerol steam reforming in a flu-
production: a review. Chem Eng Technol 32(10):1463–1469. https​ idized bed reactor. Int J Hydrogen Energy 43(41):18863–18872.
://doi.org/10.1002/ceat.20090​0120 https​://doi.org/10.1016/j.ijhyd​ene.2018.08.069
Veiga S, Bussi J (2016) Steam reforming of crude glycerol over nickel Yazdani SS, Gonzalez R (2007) Anaerobic fermentation of glyc-
supported on activated carbon. Energy Convers Manage 141:79– erol: a path to economic viability for the biofuels industry. Curr
84. https​://doi.org/10.1016/j.encon​man.2016.04.103 Opin Biotechnol 18(3):213–219. https​://doi.org/10.1016/j.copbi​
Veiga S, Faccio R, Segobia DO, Apesteguıa C, Bussi J (2017) Hydro- o.2007.05.002
gen production by crude glycerol steam reforming over Ni–La– Yenumala SR, Maity SK (2011) Reforming of vegetable oil for pro-
Ti mixed oxide catalysts. Int J Hydrogen Energy. https​://doi. duction of hydrogen: a thermodynamic analysis. Int J Hydrogen
org/10.1016/j.ijhyd​ene.2017.10.118 Energy 36(18):11666–11675. https​://doi.org/10.1016/j.ijhyd​
Wang D, Czernik S, Montane D, Mann M, Chornet E (1997) Biomass ene.2011.06.055
to hydrogen via fast pyrolysis and catalytic steam reforming of the Zhang B, Tang X, Li Y, Xu Y, Shen W (2007) Hydrogen production
pyrolysis oil or its fractions. Industr Eng Chem Res 36(5):1507– from steam reforming of ethanol and glycerol over ceria-supported
1518. https​://doi.org/10.1021/ie960​396g metal catalysts. Int J Hydrogen Energy 32(13):2367–2373. https:​ //
Wang X, Li S, Wang H, Liu B, Ma X (2008) Thermodynamic analysis doi.org/10.1016/j.ijhyd​ene.2006.11.003
of glycerin steam reforming. Energy Fuels 22(6):4285–4291. https​ Zhou C-HC, Beltramini JN, Fan Y-X, Lu GM (2008) Chemoselective
://doi.org/10.1021/ef800​487r catalytic conversion of glycerol as a biorenewable source to valu-
Wang X, Wang N, Li M, Li S, Wang S, Ma X (2010) Hydrogen able commodity chemicals. Chem Soc Rev 37(3):527–549. https​
production by glycerol steam reforming with in situ hydrogen ://doi.org/10.1039/B7073​43G
separation: a thermodynamic investigation. Int J Hydrogen
Energy 35(19):10252–10256. https​://doi.org/10.1016/j.ijhyd​ Publisher’s Note Springer Nature remains neutral with regard to
ene.2010.07.140 jurisdictional claims in published maps and institutional affiliations.
Wang S, Zhang F, Cai Q, Zhu L, Luo Z (2015) Steam reforming of ace-
tic acid over coal ash supported Fe and Ni catalysts. Int J Hydro-
gen Energy 40(35):11406–11413. https​://doi.org/10.1016/j.ijhyd​
ene.2015.03.056

13

You might also like