Vous êtes sur la page 1sur 24

Energy Conversion and Management 165 (2018) 696–719

Contents lists available at ScienceDirect

Energy Conversion and Management


journal homepage: www.elsevier.com/locate/enconman

Review

Evaluation of thermochemical routes for hydrogen production from T


biomass: A review

Aitor Arregi, Maider Amutio, Gartzen Lopez , Javier Bilbao, Martin Olazar
Department of Chemical Engineering, University of the Basque Country UPV/EHU, P.O. Box 644, E48080 Bilbao, Spain

A R T I C LE I N FO A B S T R A C T

Keywords: H2 is regarded as one of the cleanest future energy carrier that can be generated from renewable sources and will
Biomass give rise to a reduction of CO2 emissions and environmental problems related to the use of petroleum based
Bio-oil feedstock. Thus, the thermochemical routes from biomass for sustainable H2 production compared to other
Gasification biomass treatment routes have a great potential for its industrial implementation. Gasification of biomass and
Pyrolysis
reforming of the bio-oil produced by biomass pyrolysis are the most researched pathways, although some studies
Reforming
dealing with supercritical water gasification and bio-oil gasification can also be found in the literature.
Hydrogen
Nevertheless, pyrolysis and in-line catalytic steam reforming strategy is gaining great attention due to its ad-
vantages compared to gasification and bio-oil reforming, especially those related to the optimization of each step
(pyrolysis and catalytic steps) and bio-oil feeding. This review deals with the different reactor configurations,
operating conditions and catalysts used in each process and compares the different alternatives in terms of H2
production, with emphasis placing on the advantages of the two-step strategy.

1. Introduction olefins and H2.


The global hydrogen production accounts for approximately 7.7 EJ/
The urgent need to reduce the current energy dependence on fossil year, which may rise to 10 EJ/year by 2050 [7]. The main applications
fuels has promoted a large number of studies that focus on the devel- are related to ammonia production (51%), oil refining (31%), methanol
opment of existing and new processes that use biomass based materials production (10%), and other uses (8%) [7] (Fig. 2). Furthermore, H2
as feedstock. Biomass derived fuels and chemicals can play a major role market is expected to increase in the near future in a 5–10% per year,
in reducing CO2 emissions as well as become a strategic source in order basically due to its consumption in refineries for treating heavy oil
to guarantee energy competitiveness and sustainability [1]. However, fractions and because of the projected demand in the transportation
the establishment of a sustainable energy system should be based on sector or as energy vector [8].
biomass exploitation policies that take into account land usage in order The 96% of the H2 production technologies are based on non-re-
to avoid competition with human and animal food and soil exhaustion newable sources, with the most used processes being natural gas (48%)
[2]. Accordingly, lignocellulosic biomass wastes and crops are regarded and oil (30%) reforming, followed by coal gasification (18%) [9]. Only
as the most suitable alternative feedstocks [3]. 4% of the H2 produced is obtained by means of water electrolysis [9]
Lignocellulosic biomass can be treated using several thermo- (Fig. 3). Consequently, in order to meet the fossil fuel consumption and
chemical or biochemical processes in order to produce energy, bio-fuels CO2 release reduction targets, new sustainable processes derived from
and bio-chemicals [4]. Thermochemical processes, such as gasification renewable sources must be developed, such as the thermochemical ones
and pyrolysis, are characterized by their scalability to industrial units, that use biomass as feedstock shown in Fig. 1.
where the syngas and bio-oil produced as intermediate products can be Although each biomass conversion method has its own advantages
subsequently converted into valuable fuels and chemicals [5]. These and disadvantages, it has been reported that the H2 production cost for
processes have the advantage of being similar to the ones already im- gasification and pyrolysis is similar, around $1.2–2.4/kg (slightly
plemented in oil refineries, although need further development in order higher for the former), which is actually between two and three times
to be cost effective compared to fossil fuels [6]. Fig. 1 displays the main higher than the cost for CH4 steam reforming [10]. Therefore, the
processes involved in a lignocellulosic thermochemical bio-refinery for choice of the most adequate one needs a thorough assessment of the
the production of valuable products such as automotive fuels, light economic aspects in the area where it has to be implemented, as well as


Corresponding author.
E-mail address: gartzen.lopez@ehu.eus (G. Lopez).

https://doi.org/10.1016/j.enconman.2018.03.089
Received 24 January 2018; Received in revised form 27 March 2018; Accepted 28 March 2018
Available online 06 April 2018
0196-8904/ © 2018 Elsevier Ltd. All rights reserved.
A. Arregi et al. Energy Conversion and Management 165 (2018) 696–719

Nomenclature LHSV liquid hourly space velocity


RHC rice husk char
BCC brown coal char S/B, S/C, S/CH4 steam to biomass ratio, steam to carbon ratio,
BTX benzene, toluene, xylene steam to methane ratio
CFB circulating fluidized bed T, TP, TG, TR temperature, pyrolysis temperature, gasification tem-
CNT carbon nanotube perature, reforming temperature
DFB dual fluidized bed WGS Water Gas Shift
DME dimethyl ether WHSV, WBHSV weight hourly space velocity and bio-oil weight
GHSV, GC1HSV gas hourly space velocity and gas hourly space ve- hourly space velocity
locity in equivalent CH4 units

Fig. 1. Schematic representation of the main processes involved in a lignocellulosic thermochemical bio-refinery.

the availability of biomass resources, and the existence of large cata- main thermochemical biomass conversion processes for H2 production.
lytic conversion units that can treat the intermediate products. These
aspects allow making the choice between centralized processes, in 2. Biomass steam gasification
which the final product is produced at the same unit where the biomass
is primarily converted, or decentralized processes, in which the inter- Biomass gasification has been widely studied during last decades,
mediate product (such as the bio-oil derived from flash pyrolysis) can which is due to the fact the gaseous product can be directly used as fuel
be easily transported to catalytic conversion units. or as an intermediate product for the large scale production of fuels and
The mostly studied and developed technologies for H2 production chemicals [32–36]. The process characteristics entail establishing ga-
are steam gasification [11–18] and the reforming of the bio-oil pro- sification plants in the regions where biomass is available, because the
duced in biomass flash pyrolysis [19–25]; however, the biomass fast costs for the transportation of the raw material or the formed gaseous
pyrolysis and in-line reforming of the volatiles has recently gained at- products would be excessive [37]. Fig. 4 shows a schematic
tention, with several studies published in the literature over the last
years [26–31]. In this scenario, this one aims at reviewing the pyrolysis
and in-line reforming strategy and comparing this technology with the

Fig. 2. Global consumption of hydrogen [7]. Fig. 3. Current sources of hydrogen [9].

697
A. Arregi et al. Energy Conversion and Management 165 (2018) 696–719

Fig. 4. Schematic representation of the biomass steam gasification process for hydrogen production.

representation of the steam gasification process. blocking of pipes, heat exchangers and particle filters [15,43]. There-
Biomass gasification typically takes place at temperatures between fore, the improvement of tar cracking performance is one of the greatest
700 and 1200 °C, using air, oxygen, steam or their mixtures as gasifying challenges in the gasifier design [44], with the increase in residence
agent, which leads to a gaseous product mainly formed by H2, CO, CO2, time being an objective usually pursued. However, total elimination of
CH4 and other hydrocarbons. The use of steam enhances H2 formation the most stable tertiary tars requires temperatures of around 1250 °C
and produces a high heating value gas with no nitrogen. Although the and residence times above 0.5 s [42].
highly endothermic nature of steam gasification increases energy costs The char formed in the pyrolysis step is gasified by steam gasifica-
compared to air gasification, it avoids the need for a costly oxygen tion and Boudouard reactions. However, the rate of both reactions is
separation process [10]. low, even at typical gasification temperatures. Thus, char gasification
The main steps involved in biomass steam gasification are drying, represents the limiting step in biomass gasification and an adequate
pyrolysis, heterogeneous char gasification and homogeneous reactions reactor design requires a suitable residence time in order to attain high
undergone by pyrolysis volatiles, i.e., reforming, cracking and Water char conversion efficiency [45–47].
Gas Shift (WGS) reactions. These gasification steps have been sum- Therefore, biomass steam gasification products are as follows: a
marized in Fig. 5. gaseous fraction typically composed of 30–50 vol% of H2, 25–40 vol%
In the drying process, the evaporation of the moisture contained in of CO, 8–20 vol% of CO2 and 6–15 vol% of CH4; a heavier fraction (tar),
the biomass takes place, which is usually a fast step and occurs at low which consists of a complex mixture of aromatic hydrocarbons; and a
temperatures, with its influence on the overall gasification process solid fraction derived from the biomass that has not been totally gasi-
being limited. In the pyrolysis process, biomass is thermally degraded to fied. The yields and properties of these products are influenced by a
yield gases, condensable products and char. The liquid product or pri- number of factors, with the most important ones being reactor config-
mary tar formed in biomass pyrolysis is a complex mixture of oxyge- uration (which determines the contact mode, and therefore mass and
nates made up of alcohols, furans, ketones, saccharides, acids, phenols heat transfer rates), biomass initial characteristics, operating conditions
and so on [38–40]. However, these compounds are in general of marked (such as temperature and steam to biomass ratio), and the use of cat-
unstable nature, and under gasification conditions are rapidly cracked alysts [10,36,48]. Table 1 shows a summary of several biomass steam
or reformed, or evolve towards more stable aromatic structures, such as gasification research papers, in which the effect of the aforementioned
secondary and tertiary tars [41,42]. The tar contained in the syngas is a variables is studied. For the sake of clarity, studies using only steam as
serious problem for its valorisation, given that it is polymerized or gasifying agent have been analyzed. The discussion of the results re-
condensed below 300 °C, and so may cause fouling, corrosion and ported is detailed below.

Fig. 5. Biomass steam gasification reactions and steps involved.

698
Table 1
Studies involving different strategies, feeds, catalysts and operating conditions, and main results obtained (maximum H2 production) reported in the literature for biomass steam gasification.
Reactor configuration Kind of study Biomass Operating conditions Catalyst H2 production Gas yield Tar content Ref.
A. Arregi et al.

(wt%)a (Nm3/kgbiomass) (g/Nm3)

Updraft fixed bed Demonstration plant Pallets wood chips T = 750 °C; S/B = 1.4–2.7 None 6.4 1.3 60 [11]
(1.2 tn day−1)
Downdraft fixed bed Bench scale Pine wood T = 850 °C None 6.2 1.4 – [49]
Fixed bed Laboratory scale (5 g min−1) Pine sawdust T = 600–900 °C; S/B = 1.2 Primary: Dolomite 7.3 1.6 2.5 [50]
Fixed bed Laboratory scale (5 g min−1) Pine sawdust T = 600–900 °C; S/ Primary: Dolomite 11.2 2.53 n.d. [51]
B = 0–2.8
Fixed bed + fixed bed reformer Bench scale (1 kg h−1) Sawdust Gasifier: T = 800 °C; S/ Secondary: nano-Ni–La–Fe/γ- 12.1 2.4 0.08 [52]
B = 0.6 Al2O3
Secondary reactor:
T = 750–850 °C
Fluidized bed Bench scale (1–2.5 kg h−1) Waste wood T = 900 °C; S/B = 1 Silica sand 5.9 1.22 12.5 [53]
Entrained flow Pilot plant (0.67 kg h−1) Dealcoholised marc of grape T = 1050 °C; S/ None 1.3 0.6 – [13]
B = 1.02–3.19
Rotary kiln Pilot plant (3–4 kg h−1) Palm shells T = 850 °C; S/B = 0.6–1 None 3.7 0.8 1.2 [14]
Plasma Pilot plant (10 g) Distillers grains T = 600 °C; S = 1–3 ml/min None 53 vol%b – – [54]
Free fall Laboratory scale (4 g min−1) Legume straw and pine sawdust T = 750–850 °C; S/B = 0–1 Primary: Limestone; Olivine; 5.2 1.2 3.7 [55]
Dolomite
External circulating concurrent moving Laboratory scale Pine sawdust T = 650–800 °C; S/ Primary: Calcined olivine 9.5 2 0.7 [56]
bed (0.24 kg h−1) B = 0–1.2
Conical Spouted Bed Bench scale Pinewood sawdust T = 800–900 °C; S/B = 0–2 None 3.3 0.96 142 [57]
(0.75–1.5 g min−1)
Conical Spouted Bed Bench scale (1.5 g min−1) Pinewood sawdust T = 900 °C; S/B = 1 Primary: Olivine; γ-Alumina 4.4 1.15 22.4 [15]
Circulating Spout-Fluid Bed Bench scale (1.5 g min−1) Rice straw T = 860–900 °C; S/B = 2 Primary: Limestone; 5.5 1.25 2.55 [58]
Dolomite; Synthetic 0–20%

699
CaO/Al2O3
Bubbling fluidized bed Pilot plant (4 kg h−1) Pinewood chips T = 750–780 °C; S/ None 6.9 1.4 38 [59]
B = 0.53–1.1
Bubbling fluidized bed Bench scale (8 g min−1) Miscanthus X Giganteus T = 800, 900 °C; S/B = 1 Primary: Olivine; 3.9 wt% Ni/ 7.8 1.7 5.9 [60]
olivine
−1
Bubbling fluidized bed Bench scale (0.3 kg h ) Almond shells T = 700–820 °C; S/ Primary: Dolomite; Olivine 9.4 1.9 0.6 [61]
B = 0.5–1
−1
Bubbling fluidized bed Bench scale (5–8 g min ) Almond shells T = 800–830 °C; S/ Primary: 10 wt% Fe/olivine 6.6 1.4 1.4 [62]
B = 1.15–1.3
Bubbling fluidized bed Bench scale Almond shells T = 820 °C; S/B = 0.72 Primary: Olivine (0.9)/ 7.3 1.58 2.13 [63]
(6.8–11.62 g min−1) dolomite (0.1)
Bubbling fluidized bed + catalytic filter Bench scale (5–8 g min−1) Almonds shells Gasifier: T = 808–813 °C; S/ Primary: Olivine 14.5 2.9 0.15 [64]
B = 1.03–1.28 Secondary: Al2O3 filter; NiO
Filter: T = 740 °C (47 wt%)/MgO-Al2O3 filter
Bubbling fluidized bed + fixed bed Bench scale (9 g min−1) Pine sawdust Gasifier: T = 750–780 °C; S/ Secondary: Calcined – 2.3 0.5 [65]
reformer B = 1.1 dolomite, calcite and
Secondary reactor: magnesite; Ni/MgAl2O4
T = 740–880 °C
Bubbling fluidized bed + fixed bed Bench scale (0.3 kg h−1) Almond shells Gasifier: T = 770 °C; S/B = 1 Primary: Calcined dolomite 11.0 1.98 0.2 [66]
reformer Secondary reactor: Secondary: Calcined
T = 665–830 °C dolomite; Ni catalyst
Bubbling fluidized bed + tubular Bench scale (10 kg h−1) Pelletized mixture of hard and soft Gasifier: T = 615 °C; S/B = 2 Secondary: ∼25 wt% NiO/ 50 vol%b – 0.084 vol%c [67]
cracker + bubbling fluidized bed woods Thermal tubular cracker: K2O/Al2O3
reformer T = 775 °C
Secondary reactor:
T = 775–875 °C; S/C = 4.5
(continued on next page)
Energy Conversion and Management 165 (2018) 696–719
A. Arregi et al. Energy Conversion and Management 165 (2018) 696–719

2.1. Reactor configurations

[68]

[69]

[70]

[71]

[72]
[73]
[74]
[75]
Ref.

Biomass gasification is an extensively studied process, with a rela-


Tar content

tively high technological development [76]. Therefore, a wide range of


(g/Nm3)

reactor configurations have been used, such as fixed bed (updraft and

1.25

10.3

0.02
downdraft), fluidized bed, entrained flow, spouted beds, rotary kiln or
n.d.

122
10

44
5
plasma reactors [34–36,77–79].
(Nm3/kgbiomass)

The assessment of the effect of reactor configuration on the results


shown is not straightforward, due to different scale, uneven biomass
Gas yield

composition, operating conditions and gas and tar analysis procedures

0.82
0.85
used. However, analyzing the non-catalytic gasification experiments
2.2
1


carried out in updraft fixed bed [11], entrained flow [12,13], rotary
kiln [14], plasma [54], free fall [55], fluidized bed [59,70] and spouted
H2 production

bed [57,80,81] reactors, and although all reactor configurations lead to


53.7 vol%b

48.7 vol%b

H2 productions between 1.3 and 8.5 wt% (with the average value being
51 vol%b

45 vol%b
(wt%)a

4.6 wt%), with tar concentrations in the 4–140 g/Nm3 range, it can be
0.76
3.9

1.8
3.6

concluded that fluidized bed reactors are suitable for the biomass steam
gasification process because high H2 productions (6.9 wt%), gas yields
Primary: Olivine and 3.7 wt%
Primary: Limestone; Olivine;

Secondary: NiO (% 16–40)/

(around 1–1.4 Nm3/kgbiomass), H2 concentrations (> 50 vol%) and low


Secondary: 6.6%NiO/3.4%

MgO/Al2O3 honeycomb.

tar ones (20–80 g/Nm3) are obtained [59,70].


MgO/4.0% K2O/Al2O3

Fe-olivine; Ni-olivine

Ni/olivine mixtures

Furthermore, the scale up of this technology has been satisfactorily


Primary: Olivine

Primary: Olivine

accomplished, solving the heat supply problems by means of the dual


fluidized bed (DFB) configuration, which is a combination of two
Silica sand
Silica sand

bubbling and/or circulating fluidized beds, with one of them acting as


Catalyst

None

gasifier and the other as char combustor [82]. The gasifier is blown
with steam, obtaining a syngas with high H2 content, whereas the char
Secondary reactor: T = 900 °C
Gasifier: T = 700 °C; S/B = 1

is burnt in the combustor in order to provide heat to the system


[18,75,83,84]. However, it has been reported that the heat provided by
Gasifier: T = 850–900 °C
Thermal tubular cracker:

T = 815 °C; S/B = 0.84

T = 820 °C; S/B = 1.5


Operating conditions

char combustion may not be enough, and therefore additional fuel must
T = 820 °C; S/B = 2
T = 650–870 °C; S/

T = 750–850 °C; S/

T = 750–900 °C; S/

Secondary reactor:

be used in order to meet heat specifications [85]. The pioneering DFB


T = 850–900 °C

biomass gasifier was developed by Dr. Kunni, in order to gasify muni-


B = 0.5–1.2

B = 0.3–0.9

B = 0.3–0.9
T = 850 °C

cipal solid wastes [85]. Currently, the vast majority of the biomass
steam gasification demonstration or commercial scale plants are based
on this technology [86]. Thus, among the DFB gasification technologies
reviewed by Göransson et al. [83], the ones that must be remarked due
Willow wood chips; Straw; Wood/

to their capacity are SilvaGas™ plant in USA (40 MWth), which consists
of two circulating fluidized beds, 32 MWth GoBiGas plant in Sweden,
Wood pellets and chips; Bark;

15 MWth Blue Tower plant in Germany, Güssing FICFB plant in Austria


(8 MWth) and Charlmers plant in Sweden (2 MWth), the four based on a
bubbling fluidized bed as gasifier and a circulating fluidized bed as
Pine wood pellets

combustor.
straw mixtures

Sewage sludge
Wood pellets

Wood pellets

Wood pellets

Wood chips
White oak
Biomass

2.2. Effect of biomass characteristics

The main biomass characteristics affecting gasification performance


are biomass type, moisture content and particle size [10]. Regarding
Pilot plant (15–22.5 kg h−1)

Full scale (282–300 kg h−1)

Pilot plant (3.6–5.4 kg h−1)

biomass type, special attention has to be paid to the ash content as, on
Pilot plant (15 kg h−1)

Pilot plant (30 kg h−1)

Pilot plant (25 kg h−1)

the one hand, high ash content biomasses lead to high char yields and,
Tar concentration in the gas (vol%). n.d. Not detected.

on the other, ash leaves the gasifier in the form of micron size parti-
Laboratory scale

culate matter, which has to be removed downstream by means of gas


Kind of study

Pilot plant

cleaning processes [87]. Wei et al. [55] compared legume straw and
pine sawdust gasification performance in a free fall reactor, with the
H2 production defined as gH2/100 gbiomass.

former having a higher ash content. Although legume straw lead to a


lower gas yield and higher char and tar ones than pinewood, the H2
H2 concentration in the gas (vol%).

content in the gas was higher, being 50 vol% for legume straw and
bed + fixed bed reformer
cracker + bubbling fluidized bed

44 vol% for pinewood at 850 °C, which correspond to a H2 production


Bubbling fluidized bed + tubular

of 5.2 and 4.5 wt%, respectively.


Gasifiers can handle biomasses having moisture contents below
35 wt%, although most of them are designed for a 10–15 wt%. The
main drawback of raw materials with high moisture contents lies in the
Reactor configuration
Table 1 (continued)

Dual fluidized bed

Dual fluidized bed

Dual fluidized bed

bed
bed
bed

loss of gasification efficiency due to the reduction of operation tem-


perature [86]. Pfeifer et al. [69] observed for a moisture content in the
fluidized
fluidized
fluidized
fluidized
reformer

original biomass between 6 and 40 wt% that H2 concentration in the


gas increased as the moisture content was higher, although the differ-
Dual
Dual
Dual
Dual

ences were not considerable (from 34 vol% for a moisture content of


b
a

6 wt%, to 38 vol% for 40 wt%).

700
A. Arregi et al. Energy Conversion and Management 165 (2018) 696–719

The biomass particle size has a great influence on the gasification [53]. In addition, tar formation and char yield are hindered due to the
performance. Thus, smaller particles provide larger surface areas per enhancement of reforming and gasification reactions. However, the S/B
unit mass, which increases heat and mass transfer, and therefore gasi- ratio has to be optimized for each process because an excess of steam,
fication rates, leading to high gas yields with high H2 concentrations on the one hand, does not entail an improvement in the quality of the
and low tar contents [10,86]. However, reducing biomass particle size product gas and, on the other hand, it considerably increases the energy
below 1 mm increases exponentially energy consumption [86]. There- consumption of the process [86]. Fig. 7 shows some of the studies in
fore, it is essential to develop a versatile gasification technology that which the effect of steam to biomass ratio on H2 production has been
may handle large biomass particles without compromising heat and studied.
mass transfer phenomena. Thus, particle size hardly affects gasification Umeki et al. [11], operating in an updraft fixed bed gasifier with a
when the studies are carried out using reactor configurations with high feed rate of 1.2 t/day, observed that H2 production passed through a
heat transfer rates, such as the one performed by Michel et al. [60] in a maximum value of 6.4 wt% at S/B = 2.15. However, tar content in-
bubbling fluidized bed by feeding wood pellets with a size between 6 creased from 50 g/Nm3 at S/B = 1.4 to 100 g/Nm3 at S/B = 2.7. In the
and 8 mm and Erkiaga et al. [57] in a conical spouted bed with particles experiments carried out by Iovane et al. [14] in a rotary gasifier with S/
in the 0.75–4 mm size range. The latter configuration has been proven B ratios in the 0.6–1 range, it was concluded that gas yield was hardly
to be especially suitable for handling different types of residues without affected by this parameter (remaining at a value of around 0.8 Nm3/
operational problems, such as plastics [88,89] and biomass/plastic kgbiomass), while the H2 produced increased from 2.9 to 3.7 wt% for S/
mixtures [90]. B = 1. Gil et al. [59] and Göransson et al. [70], operating in a fluidized
and dual fluidized bed, respectively, observed that the S/B ratio had a
2.3. Operating conditions positive effect in both H2 concentration and tar content, with the op-
timum value being the highest one studied (1.1 for the former and 0.9
Gasifier temperature and steam to biomass (S/B) ratio are the for the latter). Thus, the H2 concentration and tar contents at the se-
parameters that need to be carefully selected and controlled for im- lected operating conditions were 55 vol% and 38 g/Nm3 [59] and
proving gasification performance [10,86]. Temperature has a great 51 vol% and 10 g/Nm3 [70].
impact on gas and H2 yields, due to the endothermicity of the reactions In the studies where dolomite and olivine were used as in-bed cat-
involved (pyrolysis, steam reforming, Boudouard and char steam gasi- alyst, S/B ratio also had a positive effect on gasification performance.
fication reactions). Furthermore, an increase in temperature hinders the Thus, Luo et al. [51], operating in a free fall reactor at 900 °C, observed
WGS reaction equilibrium. Overall, at higher temperatures a gaseous that even with the lowest S/B ratio of 0.73, tar was totally cracked and
product with higher H2 and lower CO and CH4 concentrations is ob- that the highest H2 production was obtained with a S/B ratio of 2.1
tained, as well as a reduced tar content and char yield. (11.2 wt%).

2.3.1. Temperature
2.4. Catalysts
Most of the authors that studied the influence of temperature on
product yields and compositions agreed that a more severe operation
The syngas that leaves the gasification reactor requires several
led to an improved gasification performance, minimizing the tar con-
complex and costly purification steps in order to meet specifications for
tent, as shown in Fig. 6. Thus, in the non-catalytic study performed by
the downstream catalytic conversion processes, which are designed for
Wei et al. [55], the maximum H2 productions of 5.2 and 4.5 wt% were
removing the particulate matter (micron size char and ash), the N, S
obtained at the highest temperature of 850 °C for legume straw and
and Cl containing gaseous compounds (such as NH3, HCN, H2S and
pinewood, respectively. At this temperature, tar content was mini-
HCl) and specially the tar components present in the gas
mized, with a reduction from 62.8 to 3.7 g/Nm3 for legume straw and
[42,43,97–101]. Many strategies have been developed in order to ob-
from 45.6 to 6 g/Nm3 for pinewood, in the 750–850 °C temperature
tain a tar free (or low tar content) syngas, and they can be classified into
range. Erkiaga et al. [57] operating in a conical spouted bed reactor,
primary processes, which hinder tar formation in the gasifier, or sec-
concluded that gas yield and H2 concentration increased from
ondary processes, which imply a cleaning of the produced syngas.
0.73 Nm3/kgbiomass and 28 vol% at 800 °C to 0.96 and 38 at 900 °C,
respectively (which corresponds to an increase in the H2 production
from 1.8 to 3.3 wt%).
Similarly, the utilization of a catalyst in the gasification bed resulted
in an analogous effect of temperature on gasification operation. Luo
et al. [51] using dolomite as in bed catalyst in a fixed bed observed the
maximum H2 production of 11.2 wt% at the highest temperature stu-
died (900 °C), and no tar was detected under those conditions. Olivine
also leads to a similar trend with temperature, with the best results
being observed at the highest temperatures studied. Thus, Rapagná
et al. [61] and Michel et al. [16], gasifying almond shells and mis-
canthus in a fluidized bed, attained the maximum gas yields of 1.9 and
1.2 Nm3/kgbiomass and H2 productions of 9.4 and 4.9 wt%, at 820 and
880 °C, respectively. Tar contents were also found to be minimum at the
highest temperature with a value of 0.6 g/Nm3 for the study with al-
mond shells [61]. Therefore, gasification temperature will be limited by
the economics of the process because higher temperatures lead to
higher H2 productions and a gas with lower tar content.

2.3.2. Steam to biomass ratio


Steam to biomass ratio (S/B) also plays a crucial role in gas and H2 Fig. 6. Effect of gasification temperature on tar content in the gaseous product
yields. As S/B is increased, the WGS reaction and steam reforming ones stream. Wei et al., 2007 [55]; Goransson et al., 2011 [70]; Rapagnà et al., 2000
are favoured, thus leading to a higher gaseous yield, i.e., H2 and CO2 [61]; Carpenter et al., 2010 [91]; Koppatz et al., 2011 [18]; Erkiaga et al., 2014
formation is promoted and the one of CO and hydrocarbons is reduced [57]; Zhang and Pang, 2017 [73]; Wang et al., 2017 [92].

701
A. Arregi et al. Energy Conversion and Management 165 (2018) 696–719

Amongst the primary methods for tar elimination the design of the deactivation caused by coke deposition was observed after 2 h of op-
gasifier is usually the first measure adopted. This is the case of fluidized eration.
bed gasifiers whose flexible design allows for several modifications Accordingly, efforts are being directed to the development of active
aimed at improving tar cracking efficiency, which are as follows: in- and coke resistant Ni based catalysts. As observed in Table 1, most of
crease in residence time in the freeboard region [59], location of the the secondary catalysts are based on Ni as active phase, Al2O3 as sup-
feeding zone [102,103] or secondary air injections in the case of air port and a promoter such as Mg, Ca, K or La. Pfeifer and Hofbauer [75]
gasification [104]. operated in a dual fluidized bed and secondary downstream fixed bed
Catalytic gas cleaning methods for tar removal have the advantage reactor using 16–40 wt% NiO/Al2O3 monolithic catalysts, doped with
of an additional increase in gas and H2 yields, due to the enhancement MgO or K2O. Tar content decreased from 8.5 to 0.02 g/Nm3 with the
of tar cracking and steam reforming reactions. The most widely used highest Ni loading in the catalyst, and a gas with a H2 content of 45 vol
catalysts in steam gasification studies are natural minerals, such as % was produced at 900 °C. Regarding catalyst deactivation, a tem-
dolomite or olivine, and Ni based catalysts [105–108]. The former have perature increase from 850 to 900 °C reduced coke content by 25%.
the advantage of their availability and low cost, whereas Ni based The catalytic tar elimination performance of these catalysts can be
catalysts have a high activity for tar cracking and reforming [109]. further enhanced by the utilization of a primary catalyst in the gasifier
They can be used as primary catalysts, directly in the gasifier, or as bed. For example, Di Carlo et al. [17] gasified hazelnut shell in a flui-
secondary catalysts in downstream catalytic processes. Fig. 8 shows the dized bed of olivine particles and reformed the produced gas in a fixed
tar contents obtained with different catalysts by several authors. bed made of a Ni-mayenite (Ca12Al14O33) catalyst. The gas leaving the
gasifier had an approximate tar content of 26 g/Nm3, and the secondary
2.4.1. Primary catalysts treatment reduced this content in an 84% at 800 °C, increasing H2
Dolomite has been widely used as primary catalyst due to its per- concentration in the gas from 48 vol% to 55 vol%. The authors observed
formance and availability [41]. Thus, Wei et al. [55], tested in a free fall no catalyst deactivation under these conditions for 12 h operation.
reactor the performance of different naturally occurring catalysts, such Finally, although the majority of the secondary catalytic treatment
as limestone, olivine and dolomite, finding that dolomite was the one processes have been carried out in a fixed bed reactor, Bain et al. [67]
that led to the lowest tar yield, accounting for a tar content in the gas of and Yung et al. [68] studied the secondary tar cracking and reforming
11.2 g/Nm3 for legume straw and 9.4 g/Nm3 for pinewood at 800 °C. step in a fluidized bed, which was placed downstream a fluidized bed
However, dolomite undergoes a severe attrition, especially in fluidized gasifier, followed by a thermal cracking tubular reactor. Operation in a
beds, which has boosted the utilization of other catalysts with higher fluidized bed reactor implies developing fluidizable Ni based catalysts,
mechanical resistance. with high attrition resistance. Yung et al. [68] used olivine as primary
Olivine has gained attention as primary catalyst in fluidized or high catalyst in the gasifier and developed a fluidizable 6.6%NiO/3.4%
solid circulation beds due to its high mechanical strength, together with MgO/4.0%K2O/Al2O3 reforming catalyst for the reforming step. H2
its good tar cracking activity, as mentioned previously. Koppatz et al. concentration in the gas increased from 32.8 vol% for gasification to
[18] tested the performance of olivine as primary catalyst in a 20 kg/h 53.7 vol% including the reforming step, whereas the tar, whose content
dual fluidized bed reactor, observing that H2 production and gas yield was 13.5 g/Nm3 without the reforming step, practically disappeared.
were increased from 3.1 to 4.2 wt% and 0.99 to 1.13 Nm3/kgbiomass, The authors performed 10 reaction-regeneration cycles, in which H2
respectively, when olivine was used instead of silica sand and that the concentration was reduced to 52.4 vol%, and although the catalyst was
tar content in the gas was reduced from 18.9 to 11 g/Nm3. Erkiaga et al. still active for tar reforming, CH4 concentration increased from 0.17 to
[15] compared the influence of olivine and γ-alumina addition in a 2.7 vol%, which was attributed to H2S deactivation.
conical spouted bed reactor, and observed that both catalysts were
active for tar cracking, i.e., tar was reduced from 142 with sand to 30.1
and 22.4 g/Nm3 with olivine and γ-alumina, respectively.
In order to improve the tar cracking activity of olivine, Rapagná
et al. studied the effect of Fe [62] and Ni [60] addition in a fluidized
bed. An addition of a 10 wt% of Fe increased H2 production from 3.5 to
6.6 wt% and gas yield from 1 to 1.4 Nm3/kgbiomass and reduced tar
content from 3.6 to 1.4 g/Nm3, compared to olivine. Furthermore, these
authors did not observed any coke deposition during the runs [62]. The
addition of 3.9 wt% Ni to the olivine catalyst improved gas production,
from 1 to 1.7 Nm3/kgbiomass and also H2 production, from 3.4 to 7.8 wt
% at 800 °C, with olivine and Ni/olivine, respectively. Tar production
was also reduced by three times, leading to a content of 5.9 g/Nm3 with
the Ni/olivine catalyst [60].

2.4.2. Secondary catalysts


Regarding secondary gas treatments, Delgado et al. [65] corrobo-
rated that Ni based catalysts are more active for tar elimination than
natural minerals. Thus, in the experiments performed in a fluidized bed
gasifier followed by a fixed bed downstream reactor, although both the
dolomite and the commercial Ni catalyst used reduced tar content from
9 g/Nm3 to less than 0.5 g/Nm3, the Ni based catalyst required a lower
operation temperature (740 vs. 800 °C), and lead to a higher gas yield
(2.3 vs. 1.8 Nm3/kgbiomass) and H2/CO ratio (2.55 vs. 2.2) than dolo-
mite. Rapagná et al. [66] also compared dolomite and Ni based catalyst Fig. 7. Effect of steam to biomass ratio on hydrogen production in the gasifi-
for tar removal in a similar reactor configuration, concluding that the cation process. Gil et al., 1999 [59]; Wei et al., 2006 [56]; Umeki et al., 2010
latter was more active because it lead to an increase in gas yield from 1 [11]; Rapagnà et al., 2000 [61]; Franco et al., 2003 [93]; Karmakar et al., 2011
to 1.98 Nm3/kgbiomass and reduction in tar content from 100 to 0.2 g/ [94]; Erkiaga et al., 2014 [57]; Luo et al., 2009 [51]; Gao et al., 2009 [95]; Song
Nm3, obtaining a H2 production of 11 wt% at 830 °C. However, a slight et al., 2012 [96]; Fremaux et al., 2015 [53].

702
A. Arregi et al. Energy Conversion and Management 165 (2018) 696–719

which the influence of biomass type was analyzed, Safari et al. [117]
observed that biomasses with a high cellulose content and low lignin
one lead to higher H2 productions, with the maximum value of 1.5 wt%
being obtained for wheat straw at 440 °C and 25 MPa, with a biomass
loading of 1 wt% in a batch micro reactor system. However, Yanik et al.
[118] concluded that other organic materials apart from cellulose and
lignin are responsible for the differences in gasification yields. The H2
production values reported by these authors in a batch autoclave at
500 °C ranged between 0.4 and 0.9 wt%.
Regarding the influence of catalyst on gasification performance, the
studies performed reveal that although alkali, transition metal and ac-
tivated carbon based catalysts improve H2 production, there is no
consensus regarding which type of catalyst is the most adequate one.
Ding et al. [122] compared the performance of Ni/CeO2/Al2O3, KOH
and calcined dolomite, and concluded that KOH had the best activity in
pinewood supercritical gasification, with H2 production being increased
from 0.2 to 1.1 wt% at 550 °C. Contrarily, Jin et al. [123] compared the
catalytic behaviour of Ca(OH)2, Na2CO3, K2CO3, NaOH, KOH, LiOH,
and ZnCl2, three kinds of Raney-Ni, dolomite, and olivine, in peanut
Fig. 8. Effect of different catalysts on tar content in the syngas. Rapagnà et al., shell gasification at 400 °C and 24 MPa, observing that the Raney-Ni
2000 [61]; Erkiaga et al., 2013 [15]; Miccio et al., 2009 [110]; Rapagnà et al., catalyst with the highest Fe content was the most active one, obtaining
2011 [62]; Koppatz et al., 2011 [18]; Berdugo Vilches et al., 2016 [72]. a H2 production of 2.1 wt%. Furthermore, Ru based catalysts have also
proven to have a high activity, as confirmed by Elif and Nezihe [124],
3. Supercritical water gasification who improved H2 production from 3.2 to 11 wt% in the gasification of
pulp fruit at 600 °C, using a Ru/C catalyst, and by Osada et al. [125],
Supercritical water gasification has been proposed as alternative to who also found that H2 production was enhanced, from 0.1 to 1 wt%,
steam gasification for high moisture content biomass (typically > 30 wt when a Ru/TiO2 catalyst was used in sugarcane bagasse gasification at
%), such as food derivatives [111], algae [112] or sewage sludge, as no 400 °C.
drying pretreatment is required [113]. The process takes place at
temperatures higher than 374 °C and pressures above 22.1 MPa 4. Bio-oil reforming
[114,115], with H2 production being enhanced at temperatures above
600 °C (lower ones have been reported to increase CH4 concentration). Bio-oil is the liquid fraction obtained from biomass pyrolysis, which
However, the process presents serious drawbacks for large scale is the thermal decomposition of biomass in absence of oxygen, and one
feasibility. Pumping of the feedstock is a technological challenge, as of the most energetically efficient and low cost process to obtain liquid
biomass needs to be converted into a pumpable slurry or solution, fuels from a renewable source [131–133]. Thus, under fast pyrolysis
which limits the dry biomass content in the slurry to 20 wt%, depending conditions (high heating rates and short residence times), bio-oil yields
on the biomass type. In addition, plugging by char, tar or alkaline between 60 and 75 wt% are achieved [38,134–137], which can be va-
catalysts may occur in long time runs and the severe operating condi- lorised by the catalytic steam reforming process to produce H2 as main
tions may lead to material corrosion problems [116]. The high energy product.
consumption of the process is one of the most important factors, as Although bio-oil properties change depending on pyrolysis condi-
water has to be maintained at supercritical conditions. tions, reactor configuration and the feedstock used, its direct applica-
These drawbacks have hindered the scalability of the process, and tion involves several drawbacks related to its high viscosity, low
therefore laboratory scale reactors have been used for biomass super- heating value, high oxygen and water content and high corrosiveness,
critical gasification studies, with most of the runs being carried out in which make it chemically unstable and cause ageing problems in its
batch reactors [117–125] and, to a lesser extent, in continuous flow storage [136]. In spite of the disadvantages of the bio-oil compared to a
tubular [126,127] or fluidized bed reactors [128]. The largest super- conventional fuel, it is a good alternative for H2 production from a
critical water gasification facility implemented is the VERENA pilot renewable source [106,138–140]. Thus, the aforementioned problems
plant at Karlsruhe Institute of Technology (100 kg/h), which has been lead to consider different alternatives for its upgrading, such as mix it
designed for 700 °C, 35 MPa and a maximum solids content of 20 wt%, with other fuels or physical, thermal and catalytic treatments (catalytic
and includes a down-flow reactor and a biomass and water preheating cracking, esterification, hydrodeoxygenation, steam reforming, etc.)
step with the reactor outlet stream [113,129]. [136,138,140]. The catalytic steam reforming of bio-oil (Fig. 9) is one
In order to reduce the operating temperature and increase H2 se- of these routes that has been considerably researched, using different
lectivity, several catalysts have been developed, which can be classified reactor configurations, catalysts and operating conditions. The high
into alkali based (such as Na2CO3, K2CO3 or NaOH), transition metal water content of the bio-oil is not a problem for this valorisation route.
based (Ni and Ru) and activated carbon catalysts. Although all of them The aforementioned difficulties related to bio-oil properties and its
have proven to be suitable for enhancing gasification and improving H2 feeding problems boosted the use of bio-oil model compounds as feed,
yield, alkali catalysts lead to corrosion and plugging problems, whereas which helps understanding the mechanisms of reforming reactions
transition metal and activated carbon based catalysts undergo a severe [141], with acetic acid [142–158], acetone [159–166] and ethanol
deactivation [115,130]. [167–170] being among others the most studied ones. Nevertheless,
Several studies have been carried out analyzing the supercritical although some works have been published concerning reforming of
gasification of food and agricultural residues, in which the H2 pro- phenols [150,151,165,166,171–173], and especially m-cresol
ductions reported vary in a wide range, between 0.1 and 11 wt% (de- [174–179], furans [176,177,180], saccharides [162,174] and aromatic
fined as gH2/100 gbiomass) which depend on the biomass type and compounds, such as toluene [181–183] or m-xylene [162], cyclic
loading, operating conditions and catalysts used. Thus, in the non-cat- oxygenated compounds (phenols, saccharides and furans) are less in-
alytic supercritical gasification studies carried out in batch reactors in vestigated, but they are supposed to be responsible for the catalyst
deactivation by coke deposition [180].

703
A. Arregi et al. Energy Conversion and Management 165 (2018) 696–719

Consequently, this strategy focuses on the valorisation of the bio-oil 4.2. Operating variables
aqueous fraction and, especially, raw bio-oil, which has recently gained
increasing attention. Therefore, Tables 2 and 3 summarize some of In spite of the fact that a detailed comparison of results is difficult,
these studies, in which reforming of bio-oil aqueous fraction (Table 2) the most relevant ones on the effect operating conditions have on
and raw bio-oil (Table 3) have been performed. Although different re- product yields and, especially, on H2 production, have been summar-
actors and operating conditions have been used, a discussion of the ized in this section.
main results has been detailed below. In this section, the values of H2
production are generally given as g of H2 per 100 g of organic com-
4.2.1. Temperature
pounds in the feed, and otherwise it will be specified in each case.
Temperature is one of the key variables that should be taken into
Moreover, H2 yield is based on the maximum allowable by stoichio-
account in the reforming processes. On the one hand, higher tempera-
metry.
tures enhance reforming kinetics of oxygenated compounds of the bio-
oil and, therefore, higher H2 yield and production are obtained. On the
4.1. Reactor configurations other hand, using high temperatures the coke deposition can be reduced
given that its in situ gasification is promoted [192,199,201,216].
Many studies on the reforming of bio-oil aqueous fraction are found Fig. 10 displays the effect of temperature on H2 yield in several bio-oil
in the literature, which have been carried out mainly in fixed reforming processes studied in the literature.
[23,24,184,186,188–191,196,210] and fluidized bed [21,185,187] re- The effect of this variable on conversion and product distribution
actors. Similarly to the reforming of the bio-oil aqueous fraction, the and, especially, on H2 production has been extensively studied in the
reforming of raw bio-oil has also been performed mainly in fixed bed literature in the reforming of bio-oil aqueous fraction. Thus, Czernik
[25,161,198,200,203–207,210] and fluidized bed [197,202,208] re- et al. [185] reported that coke gasification was enhanced at 850 °C,
actors. Higher conversion and H2 yield were obtained by Remón et al. obtaining a maximum conversion and H2 yield of 95% and 89%, re-
[209] in a fixed bed reactor compared to fluidized bed reactor. spectively, with 1000 h−1 and S/C of 7, which corresponds to a H2
Nevertheless, Lan et al. [201] reported lower carbon deposition in the production of 16.8 wt%. Yan et al. [188] reported a very low H2 yield at
fluidized bed reactor than in the fixed bed reactor, obtaining carbon 450 °C (9.4%), whereas a maximum H2 yield of 69.7% was attained at
contents of around 2.3 wt% and 0.3 wt%, respectively, at 650 °C in the highest temperature studied (800 °C), with water to bio-oil ratio
60 min on stream. being 4.9 and with Ni and Ce contents 12 and 7.5%, respectively. The
Moreover, some authors used a two-stage system for bio-oil re- effect of temperature in the 600–800 °C range was also studied by
forming. The steam reforming of bio-oil aqueous fraction Bimbela et al. [190], obtaining higher conversion and H2 yield and
[192,194,213], raw bio-oil [22,211,212] and bio-oil/bio-ethanol mix- production as temperature was increased, with the maximum values
tures [214,215] was investigated by Remiro et al. and Valle et al. using being 82.62%, 77% and 13.8 wt%, respectively. Nevertheless, the
a two-stage thermal-catalytic process, in which the pyrolytic lignin was highest H2 concentration was obtained at 650 °C (65.48 vol%), which
deposited by repolymerization of certain bio-oil components in the first decreased to 56.87 vol% at 800 °C. Yao et al. [195] studied the influ-
stage and the treated bio-oil together with the gases formed in the first ence of temperature in the 600–900 °C range, with conversion and H2
step were in-line reformed in a fluidized bed reactor. Yao et al. [195] yield being increased from 8.11 to 90.40% and from 5.64 to 57.21%,
used a two-stage fixed bed reactor, where firstly the bio-oil aqueous respectively (maximum H2 production of 10.4 wt%).
fraction was volatilized at 400 °C and subsequently the reforming step Therefore, most studies reveal that high temperatures are needed to
was carried out on modified Ni-Al catalysts. achieve full conversion of bio-oil aqueous fraction. Nevertheless,
Nevertheless, although different reactors have been used for bio-oil Remiro et al. [192] reported full conversion above 550 °C, although
reforming, the main issue to be solved is related to raw bio-oil feeding, catalyst deactivation was lower as temperature was increased in the
which causes several problems in the reforming reactions. Basagiannis 500–650 °C range.
et al. [186] reported coke deposition on the reactor wall due to the Concerning raw bio-oil reforming, similar effect of temperature on
polymerization of unstable compounds in the bio-oil. Although the conversion and product distribution has been reported in the literature,
feeding system was cooled, Kechagiopoulos et al. [20] reported plug- with both increasing at higher reforming temperatures [203,204,207].
ging of the injection nozzle, due to the thermal instability of the aqu- Thus, Wang et al. [198] reported higher carbon conversion and H2 yield
eous fraction of the bio-oil used for the process, which required a new as temperature was increased in the 250–750 °C range, obtaining a
injection-cooling system to feed adequately the bio-oil fraction. More- maximum conversion, H2 yield and H2 production of 96%, 80% and
over, some authors reported aging problems of the bio-oil, which have 20.1 wt%, respectively, at 750 °C and S/C ratio of 9 when C12A7-O−/
been solved by adding different alcohols such as methanol [197] or 18% Mg catalyst was used. As temperature was increased in the
ethanol [215] in order to stabilize the bio-oil and reduce its viscosity. 400–550 °C range by Hou et al. [200], higher conversion and H2 yield

Fig. 9. Schematic representation of the bio-oil reforming process for hydrogen production.

704
A. Arregi et al. Energy Conversion and Management 165 (2018) 696–719

Table 2
Studies involving different strategies, feeds, catalysts and operating conditions, and main results obtained (maximum H2 yield or production) reported in the
literature for reforming of bio-oil aqueous fraction.
Reactor Kind of study Biomass Catalyst Operating conditions Conversion H2 H2 production Ref.
configuration yield (wt%)b
(%)a

Dual fixed bed Laboratory scale Poplar Ni/Al2O3 (reference) T = 825–875 °C 98 88 12.4* [184]
(150 mg min−1) wood Commercial: G-91, C11- GC1HSV = 126,000–80,000 h−1
NK, ICI46-1, ICI46-4 S/C = 4.92–11
Co-promoted Ni and Cr-
promoted Ni supported on
MgO-La2O3-α-Al2O3
Bubbling Bench scale Pine C11-NK T = 800–850 °C 95 89 16.8* [185]
fluidized (120–300 g h−1) sawdust GC1HSV = 700–1000 h−1 3.0d
bed S/C = 7–9
Nozzle-fed fixed Bench scale Beech Ru/MgO/Al2O3 T = 700–800 °C 100 100c – [186]
bed (290–1050 ml min−1) wood GHSV = 4880–16,570 h−1
S/C = 7.2
Fluidized bed – Sawdust Ni/dolomite T = 600–800 °C – 73 – [187]
Ni/modified dolomite WHSV = 1.5 h−1
S/C = 2–10
Laboratory-scale Laboratory scale Rice hull Commercial: Z417 T = 450–800 °C – 69.7 – [188]
fixed bed Ni/CeO2-ZrO2 Water/bio-oil = 3.2–5.8
Ni loading = 5–12 wt%
Ce loading = 5–10 wt%
Fixed bed Laboratory scale Corn G90-LDP T = 500–700 °C 80 65 12.5 [189]
(4 ml h−1) stover WHSV = 0.87 h−1
S/C = 4–18
Bench scale Small bench scale Pine wood Ni/Al coprecipitated T = 600–800 °C 82.62 77 13.8 [190]
fixed bed catalysts Space
Ni loading = 23–33% time = 0–5.01 gcatalyst min gorganic−1
S/C = 5.58
Fixed bed – Pine wood Ni-Mo supported on T = 600–800 °C 98 67.5 – [191]
sawdust modified sepiolite WHSV = 0.9–5.4 h−1
S/C = 2–25
Fixed bed/ Laboratory scale Pine Ni/La2O3-αAl2O3 T1 = 500 °C 100 96 15.6* [192]
fluidized (0.1 ml min−1) sawdust T2 = 600–800 °C 10.1d
bed Space time = 0.10–0.45 gcatalyst h gbio-
−1
oil
S/C = 12
Fixed bed or Laboratory scale Pine Ni/Al-Mg-O modified with T = 650 °C 80 – 13.8 [193]
fluidized (0.12–0.76 ml min−1) sawdust Co or Cu Space time = 4 gcatalyst min gorcanic−1
bed S/C = 7.6
Fixed bed/ Laboratory scale Pine Ni/αAl2O3 T1 = 200 °C 100 96 15.6* [194]
fluidized (0.1 ml min−1) sawdust Ni/La2O3-αAl2O3 T2 = 600–800 °C 10.1d
bed Space time = 0.10–0.45 gcatalyst h gbio-
−1
oil
S/C = 12
Fixed bed/fixed Laboratory scale Corn stalk Ni-Al modified with Ca, T1 = 400 °C (volatilization) 90.4 57.2 10.4* [195]
bed (0.3 g min−1) Ce, Mg, Mn and Zn T2 = 600–900 °C
Space time = 1.67 gcatalyst min gbio-oil−1
S/C = 3.54–9
Fixed bed Laboratory scale Pine/ Ni/Mg-Al modified with T = 650 °C 78.7 – 14.7 [196]
(0.12 ml min−1) poplar Ce GC1HSV = 13,000 h−1
wood

a
H2 yield referred to maximum allowable by stoichiometry.
b
H2 production defined as gH2/100 gorganic.
c
H2 selectivity.
d
H2 production defined as gH2/100 gbio-oil.
* Calculated by H2 yield and bio-oil composition.

and production were achieved, obtaining maximum values of 94.9%, fraction at 650 °C on a Mg modified Ni-Al catalyst, obtaining a sig-
92.5% and 17.5 wt%, respectively, at 550 °C with the highest S/C ratio nificant increase in conversion and H2 concentration from 81.01% to
studied on a 15% Ni-CNT catalyst. Valle et al. [211] obtained a H2 yield 83.26% and from 63.13 vol% to 67.37 vol%, respectively, and
of around 88% above 650 °C, 0.10 gcatalyst h gbio-oil−1 and S/C ratio of 6. achieving a maximum H2 production of 13.3 wt%. Liu et al. [191]
studied the effect of space velocity at 750 °C and S/C ratio of 16, ob-
taining the highest values of conversion (98%) and H2 yield (62.5%)
4.2.2. Space time below 1.8 h−1.
Space time is an important variable whose effect has been studied Similar results were obtained in the literature when raw bio-oil
by several authors in the reforming of bio-oil aqueous fraction. Most reforming was carried out, obtaining higher H2 yields as space velocity
studies reported that higher space times enhance reforming and WGS is decreased [22,25,204,207,208]. Wu et al. [199] reported that lower
reactions, and therefore higher H2 yields are obtained. Thus, Medrano GHSV values lead to higher CH4 conversion and H2 yield, obtaining full
et al. [21] studied two different space time values (4.11 and conversion and a H2 yield of 81.1% at 800 °C and with a GHSV value
8.97 gcat min gorganics−1) in the steam reforming of bio-oil aqueous

705
A. Arregi et al. Energy Conversion and Management 165 (2018) 696–719

Table 3
Studies involving different strategies, feeds, catalysts and operating conditions, and main results obtained (maximum H2 yield or production) reported in the
literature for reforming of raw bio-oil.
Reactor configuration Kind of study Biomass Catalyst Operating conditions Conversion H2 H2 Ref.
yield production
(%)a (wt%)b

Fixed bed Laboratory scale Beech wood Pt, Pd, Rh supported on T = 650–950 °C – 75 14.5d,* [161]
(14 μL min−1) Al2O3 and CeZrO2 GHSV = 3090 h−1
S/C = 5/10.8
Fluidized bed Bench scale (84 g h−1) Hardwood + 10% C11-NK T = 850 °C 95 80 12.9d [197]
methanol Laboratory Ni catalysts GC1HSV = 920 h−1
S/C = 5.8
Fixed bed Laboratory scale Biomass C12A7-O-/M catalysts T = 250–750 °C 96 80 20.1* [198]
M = Mg, K, Ce GHSV = 10,000 h−1 12.4d,*
S/C = 1.5–9
Fixed bed/fixed bed Laboratory scale Sawdust 1 – Dolomite T1 = 700–900 °C 100 81.1 13.2d,* [199]
2 – Ni/MgO T2 = 600–800 °C
(WBHSV)1 = 0.5–5 h−1
(GHSV)2 = 1800–14,400 h−1
(S/C)1 = 1–16
(S/CH4)2 = 1–4
Fixed bed – Sawdust Ni-CNT T = 350–550 °C 94.9 92.5 17.5* [200]
Ni loading = 5, 15, 35% GHSV = 12,000 h−1
S/C = 2.0–6.1
Fixed bed or fluidized bed Bench scale Rice husk Ni/MgO-La2O3-Al2O3 Tfixed = 650–950 °C – 75.88 17.6* [201]
LHSVfixed = 0.8–2.5 h−1
S/Cfixed = 3–14
Tfluidized = 500–800 °C
LHSVfluidized = 0.2–1.5 h−1
S/Cfluidized = 8–20
Fluidized bed Bench scale Rice husk NiO/MgO T = 500–800 °C – 56.3 13.1* [202]
(100 g h−1) WHSV = 0.2–0.8 h−1
S/C = 8–20
Fixed bed Laboratory scale Biomass Ni/Al2O3 T = 750–950 °C 81 85 14.3d,* [203]
(2.13 ml min−1) Ru-Ni/Al2O3 WBHSV = 13 h−1
Ni-MgO/Al2O3 S/C = 5
Ni loading = 0–18%
Ru loading = 0.5%
Fixed bed Laboratory scale Biomass Ni/Al2O3 T = 650–950 °C 79 73 12.3d,* [204]
(2.13 ml min−1) Ni loading = 0–18% WBHSV = 9–26 h−1
S/C = 1–8
Fixed bed Laboratory scale Sawdust NiCo/Ce-Zr-O T = 850 °C – 72.15 13.2* [205]
(12 ml min−1) Ni loading = 2–10% WHSV = 2.62 h−1
Co loading = 2–10%
Fixed bed Laboratory scale Biomass Ni/ZrO2 T = 850 °C – 92 20.1* [206]
(2.13 ml min−1) Ni/Al2O3 WBHSV = 13 h−1 15.5d,*
Ni loading = 0–18% S/C = 5
Fixed bed/fluidized bed Laboratory scale Pine sawdust 1 – Dolomite T1 = 500 °C 100 95 15.4* [22]
(0.1 ml min−1) 2 – Ni/La2O3-αAl2O3 T2 = 600–800 °C 11.7d
GC1HSV = 8000–156,000 h−1
S/C = 1–15
Fixed bed Laboratory scale Maize stalk Ni-Ce/Al2O3 T = 700–900 °C – 71.4 11.1d,* [207]
Ni loading = 5.1–20.4% WBHSV = 8–24 h−1
S/C = 1–9
Fluidized bed Bench scale Pine sawdust Ni/dolomite T = 500–800 °C 80.0 45c – [208]
(100 g h−1) Ni/modified dolomite WHSV = 0.2–1 h−1
S/C = 8–20
Fixed bed or fluidized bed Laboratory scale Pine sawdust Ni/Co-Al-Mg T = 650 °C 95 – 17.0* [209]
(0.3 g min−1) GC1HSV = 13,000 h−1
S/C = 7.6
Fixed bed Laboratory scale Corn cob Ce-Ni/Co-Al2O3 T = 650–850 °C – 85 11.7* [25]
LHSV = 0.08–0.23 h−1 6.6d,*
S/C = 9–15
Fixed bed Laboratory scale (1 g/ Coconut shell Fe/olivine T = 800 °C 97.2 79.3 – [210]
h) WBHSV = 0.5 h−1
S/C = 2
Fixed bed/fluidized bed Laboratory scale Pine sawdust Ni/La2O3-αAl2O3 T1 = 500 °C 100 93 – [211,212]
(0.08–0.1 ml min−1) T2 = 550–700 °C
Space
time = 0.04–0.38 gcat h gbio-oil
S/C = 1.5–6

a
H2 yield referred to maximum allowable by stoichiometry.
b
H2 production defined as gH2/100 gorganic.
c
H2 concentration in the gas (vol%).
d
H2 production defined as gH2/100 gbio-oil.
* Calculated by H2 yield and bio-oil composition.

706
A. Arregi et al. Energy Conversion and Management 165 (2018) 696–719

lower than 3600 h−1. Remiro et al. [22] reported higher conversion and its capacity for water adsorption, and therefore for enhancing the WGS
H2 yield when space velocity was decreased, obtaining full conversion reaction. Moreover, higher stability of the catalyst was also observed
and H2 yield of 95% at a S/C ratio of 9 and for GC1HSV lower than when La2O3 was added, maintaining almost full conversion for up to
65,000 h−1. Moreover, as space velocity was decreased lower coke 20 h on stream, whereas it decreased subsequent to 7 h when the Ni/
content on the catalyst was observed, which decreases from 1.14 to Al2O3 catalyst was used. Different Ni-Al catalysts modified with Ca, Ce,
0.1 wt% in the 156,000–8000 h−1 range. Mg, Mn and Zn were used by Yao et al. [195] in the reforming of bio-oil
Nevertheless, when low space velocity values are used, higher aqueous fraction at 800 °C and a S/C ratio of 3.54. The highest H2 yield
performance of the WGS reaction is observed, given that CO con- of 56.46% was obtained when Ni-Mg-Al catalyst was used, followed by
centration decreases and CO2 concentration increases, at the same time Ni-Ce-Al (55.30%) and Ni-Zn-Al (52.01%) catalysts.
as secondary products such as CH4 and light hydrocarbons produced Similarly, the addition of promoters has also been investigated in
from mainly cracking reactions disappear [21,186,190]. bio-oil reforming. Salehi et al. [203] compared Ru-Ni/Al2O3 and Ni/
Al2O3 catalyst with different Ni contents, obtaining a maximum H2
4.2.3. Steam/C ratio yield of 85% at 950 °C when Ru-Ni/Al2O3 catalyst was used with a Ni
The most important effect of S/C ratio is the enhancement of the content higher than 10.7%.
WGS reaction, which favours the formation of H2 and CO2 as main The effect of metal content on the catalyst has also been analyzed in
products. Thus, most studies reported that CO, CH4 and light hydro- the literature for the reforming of bio-oil aqueous fraction. Yan et al.
carbon yields are decreased as S/C ratio is increased [188] studied different Ni (between 5 and 12 wt%) and Ce (between 5
[22,187,188,198,208]. Fig. 11 shows the effect of steam to C ratio on and 10 wt%) loadings in Ni/CeO2-ZrO2 catalysts and they did not ob-
H2 yield in several bio-oil reforming studies. serve high differences between the different loadings, although a little
When S/C ratio was increased from 2 to 5 by Li et al. [187] in the increase in H2 yield was achieved at the highest Ni loading studied.
reforming of bio-oil aqueous fraction, H2 yield was significantly af- Bimbela et al. [190] investigated the effect of Ni content in the 23–33%
fected, although above this value the influence of S/C ratio was less range on product distribution and the highest conversion (63.35%) and
pronounced. Ortiz-Toral et al. [189] reported that for S/C > 8 the H2 concentration (65.48 vol%) were obtained with the 28% Ni catalyst
effect of this variable on H2 production was negligible. Liu et al. [191] at 650 °C and 1.67 gcatalyst min gorganic−1.
reported that for S/C ratios above 12, the yields become constant, ob- Concerning raw bio-oil reforming, the influence of Ni content on a
taining a H2 yield of 61% with the highest S/C ratio, at 750 °C and Ni/Al2O3 catalyst was studied by Seyedeyn-Azad et al. [204], obtaining
1.8 h−1. the highest conversion (79%) and H2 yield (73%) when a Ni content of
Similarly, studies related to raw bio-oil reforming also report higher 14.1% was used. The same authors reported the effect of the support by
H2 yields and carbon conversions when higher S/C ratios were used comparing Ni/ZrO2 and Ni/Al2O3 catalysts, achieving higher H2 yields
[202]. Thus, Wang et al. [198] observed the enhancement of the WGS when the Ni/ZrO2 catalyst with 14% Ni was used [206]. Nevertheless,
reaction as S/C ratio was increased in the 1.5–9 range, especially be- the amount of coke deposited on the catalyst was much higher when the
tween 1.5 and 4, with the increase being smaller in the 4–9 range. Lan Ni/ZrO2 catalyst was used than when the Ni/Al2O3 catalyst was used.
et al. [201] reported the enhancement of reforming and WGS reactions The effect of Ni and Co content on a NiCo/Ce-Zr-O catalyst was studied
at high S/C values in both fixed bed and fluidized bed reactors, al- by Zhang et al. [205] in the 2–10% range for both metals, obtaining the
though the steam needed in the fluidized bed reactor was higher be- best results in terms of H2 yield (72.15%), H2 production (13.2 wt%)
cause it acted also as a fluidizing agent. Remiro et al. [22] reported and coke deposition with 3Ni9Co/Ce-Zr-O catalyst.
higher H2 yields as S/C ratio was increased in the 1–15 range, although Moreover, increasing attention is being paid to the use of CO2 ad-
above S/C of 5 the increase was less significant, obtaining a maximum sorbents in reforming processes, given that the thermodynamic equili-
H2 yield of 95% at the highest S/C ratio studied. brium of bio-oil reforming reaction is shifted, and therefore higher H2
Moreover, at high S/C ratios water adsorption on the catalyst sur- yields are obtained [22,165,213,218,219]. Thus, Remiro et al. [213]
face is enhanced, favouring the gasification of coke, and therefore im- used dolomite as CO2 adsorbent in the catalytic steam reforming of bio-
proving the overall reforming process [184,198,207]. Thus, when S/C
ratio was increased by Li et al. [187], lower coke yield was obtained,
with the decrease being more pronounced from 2 to 5.

4.3. Catalysts

Different catalysts have been studied in the literature in the re-


forming of bio-oil aqueous fraction and raw bio-oil, although Ni based
are the most used ones [106,139,216,217]. Czernik et al. [197] com-
pared a commercial (C11-NK) and laboratory prepared catalysts in the
reforming of raw bio-oil and reported higher stability of the commercial
catalyst due to the higher Ni content and surface area of that catalyst.
Furthermore, it has high initial conversion (95%) and H2 production
(12.9 wt% by mass unit of bio-oil). Nevertheless, some authors also
used noble metals, which are very active for reforming reactions, al-
though its high cost is a drawback for its industrial implementation.
Thus, Rioche et al. [161] studied Rh and Pt catalysts supported on
Al2O3 and CeZrO2 in the reforming of raw bio-oil, obtaining a max-
imum H2 yield of 75% using 1% Pt/CeZrO2 catalyst, with these values
being noticeably higher compared to Al2O3 supported catalysts.
Furthermore, different promoters have also been studied in the re- Fig. 10. Effect of reforming temperature on the hydrogen yield (percentage of
forming of bio-oil aqueous fraction in order to improve catalyst activity the maximum allowable by stoichiometry) in aqueous and raw bio-oil re-
and stability, with Ce, La, Mg and Ca being the most studied ones. The forming processes. Li et al., 2009 [187]; Liu et al., 2013 [191]; Remiro et al.,
effect of La2O3 addition to Ni/Al2O3 catalyst was studied by Valle et al. 2013 [192]; Yao et al., 2014 [195]; Wang et al., 2007 [198]; Xu et al., 2010
[194] and a good performance was reported, which was explained by [202]; Valle et al., 2018 [211].

707
A. Arregi et al. Energy Conversion and Management 165 (2018) 696–719

stability of 3Ni9Co/Ce-Zr-O catalyst, obtaining a decrease of 12% in H2


yield in 360 min on stream and a value of around 64% when the cat-
alyst was deactivated. Fu et al. [207] reported slow deactivation of the
Ni-Ce/Al2O3 catalyst up to 6 h on stream at 850 °C, S/C ratio of 6 and
WBHSV of 12 h−1, whereas the deactivation was faster subsequent to
6 h until a low H2 yield was obtained in 10 h on stream.

5. Bio-oil steam gasification

Bio-oil gasification has been proposed as an alternative to biomass


gasification for syngas production, with the aim of reducing the ex-
pensive biomass transport costs [221]. Therefore, as in the case of bio-
oil reforming, several decentralized pyrolysis plants coupled with a
centralized bio-oil gasification facility would be required in order to
take advantage of the economies of scale [222]. In addition, the pyr-
olysis char may also be upgraded together with the bio-oil, by feeding it
as bio-slurry, thus taking advantage of the potential of biomass pyr-
olysis char to produce H2 [223,224]. In this way, the biomass that has
an energy density of about 2 GJ/m3 is upgraded to a bio-slurry with a
Fig. 11. Effect of steam to C ratio on H2 yield (percentage of the maximum density of 25 GJ/m3, thereby considerably reducing the transport costs
allowable by stoichiometry) in aqueous and raw bio-oil reforming studies. Li
[36]. It has been reported that bio-oil gasification becomes more at-
et al., 2009 [187]; Liu et al., 2013 [191]; Wang et al., 2007 [198]; Lan et al.,
tractive economically than biomass gasification for large scale plants,
2010 [201]; Seyedeyn-Azad et al., 2011 [204]; Fu et al., 2014 [207]; Valle
et al., 2018 [212].
low energy density biomass and very long transport distances. How-
ever, improvements in biomass pyrolysis efficiency and gasification
plants’ capital costs would make bio-oil gasification cost-competitive
oil aqueous fraction, obtaining full conversion of the bio-oil and H2 also in smaller distances [225].
yields around 99%. Nevertheless, further studies dealing with reaction/ The differences between bio-oil gasification and reforming are
regeneration cycles are required in order to study the loss of activity of vague and sometimes are not well distinguished in the literature. In this
the adsorbents. paper, gasification has been considered as the process that takes place
at higher temperatures than bio-oil reforming (around 800–1400 °C),
4.4. Catalyst deactivation and is carried out non-catalytically or in the presence of primary mi-
neral catalysts, as in biomass gasification. Therefore, bio-oil gasification
Some authors reported high deactivation of the catalyst due to coke produces a syngas with a similar composition than the one obtained in
deposition on its surface. Catalyst deactivation leads to secondary re- biomass gasification, whereas bio-oil reforming yields a gas with higher
actions, such as thermal decomposition of bio-oil and Boudouard re- H2 content. The lower H2 yield and energy efficiency and higher costs of
actions. Although it has not been extensively studied in the literature, bio-oil gasification are responsible for the lower technological devel-
some authors reported that H2 and CO2 yields decrease and CO, CH4 opment of this thermochemical route [226].
and hydrocarbons yields increase with time on stream due to the at- The H2 productions (defined in this section as gH2/100 gbio-oil) ob-
tenuation of oxygenated compound reforming and the enhancement of tained by several authors in the literature vary in a wide range, as a
secondary reactions in the reforming of bio-oil aqueous fraction result of differences in the bio-oils used, reactor configurations and
[21,185,191,192]. Moreover, coke deposition may also be attributed to operating conditions. In the non-catalytic studies analyzed, the average
the dehydration and polymerization of the aromatic compounds in the H2 production can be established at around 4 wt%, although values
bio-oil [220]. This trend on product yields was observed by Garcia et al. between 1.4 and 12.6 wt% have been reported. Thus, Panigrahi et al.
[184], obtaining by-products, such as CH4, methanol, benzene, toluene [227], operating in a fixed bed micro-reactor at 800 °C, obtained the
and naphthalene as the catalyst was deactivating. Czernik et al. [185] maximum H2 production of 4.4 wt% with a steam/carbon ratio of 2. A
reported a conversion of around 100% of the bio-oil aqueous fraction similar value of 4 wt% was attained by Kan et al. [228], also in a fixed
and an initial H2 yield of 85%, although it decreased to 77% after 12 h bed at 800 °C and a S/C of 10.6. Furthermore, Latifi et al. [229], op-
on stream due to coke deposition. Bimbela et al. [190] reported a de- erating in a Jiggle bed reactor at 800 °C, without adding additional
creasing trend of conversion with time on stream due to the low activity steam to the bio-oil, observed a H2 production of 3.5 wt%. Van Rossum
of the catalyst for WGS reaction, at the same time as H2 production et al. [230] obtained a lower H2 production of 1.4 wt% in a fluidized
decreases and those of CO, CH4 and C2 increase. Nevertheless, at the bed at 790 °C and S/C of 1.9. Finally, Chhiti et al. [231] reached the
highest space time studied, H2 yield was maintained constant for up to highest H2 production of 12.6 wt% at 1400 °C and S/C = 8.3 in a la-
120 min on stream. Moreover, Valle et al. [194] reported higher sta- boratory scale high temperature entrained flow reactor.
bility (conversion constant for up to 5 h on stream) and lower coke
deposition with Ni-La2O3/αAl2O3 catalyst than with Ni/Al2O3 catalyst, 6. Pyrolysis and in-line steam reforming
obtaining a coke content of 1.40 and 2.50 wt% at 600 °C and 0 and
0.10 wt% at 800 °C, respectively. Among the different thermochemical routes aforementioned for H2
The same trend was observed when raw bio-oil was valorised by production from biomass, the two-stage pyrolysis-reforming strategy
catalytic steam reforming [22,198]. Thus, Rioche et al. [161] per- has been gaining increasing attention recently, given that it has several
formed a long experiment with 1% Pt/CeZrO2 catalyst, obtaining a slow advantages in relation to the gasification process due to the separation
deactivation of the catalyst after 9 h on stream, with H2 yield being of pyrolysis and reforming steps in different zones. Thus, this process
between 50 and 60%. Hou et al. [200] studied the stability of a 15% Ni- allows for independent temperature optimization in the pyrolysis and
CNT catalyst at 500 °C, GHSV of 12,000 h−1 and S/C ratio of 6.1, re- reforming steps [232]. Moreover, compared to catalytic gasification
gistering a slow decrease on carbon conversion and H2 yield for the first process, the direct contact between the feedstock and its impurities
8 h on stream, whereas above that time H2 yield decreased considerably with the reforming catalyst is avoided [233]. In addition, the reforming
(50% lower) until 15 h on stream. Zhang et al. [205] studied the step is carried out at lower temperature in comparison to catalytic

708
A. Arregi et al. Energy Conversion and Management 165 (2018) 696–719

gasification, and therefore catalyst deactivation by sintering can be fluidized bed reactors. Although the use of fixed bed reactors is more
avoided [234,235], especially when Ni is used as active phase, which is extended, fluidized bed reactors avoid operational problems related to
the most widely used catalyst in reforming of oxygenated compounds bed blockage by coke deposition in the reforming step [201,236]. Thus,
and hydrocarbons. Furthermore, the gasification process does not focus Fig. 13 shows the different reactor configurations studied in the lit-
on H2 production, but syngas is the main product. The quality and erature for this two-step process.
concentration of the syngas depends on the gasifying agent used, and The group headed by Prof. Williams carries out the process in dis-
therefore subsequent catalytic processes are needed to enhance H2 continuous mode, using a two stage fixed bed reactor (Fig. 13a), in
production. Moreover, highly active steam reforming catalysts allow for which the biomass is pyrolyzed in the first stage and the volatiles from
attaining complete conversion of biomass derived volatiles, with the gas the pyrolysis circulate through the catalytic reforming step, with wood
product being free of tars, which is a remarkable advantage over con- sawdust and pellets being the feedstock used for most of the studies
ventional gasification [31]. [26,27,239,240], although other materials, such as cellulose [241], rice
Regarding bio-oil reforming, volatiles from biomass are valorised in- husk, sugar cane bagasse and wheat straw [242] or biomass/plastic
line, and therefore condensation and volatilization of the intermediate mixtures [251,252] have also been investigated. This two-stage fixed
compounds from biomass pyrolysis are not required, which avoids the bed reactor strategy has also been used by other authors for the pyr-
difficult handling of these products in addition to possible mass loss and olysis-reforming of different materials, such as pine wood chip, pig
re-polymerization during the raw bio-oil volatilization step. Moreover, manure compost, sewage sludge [28,243], cedar wood [253], bagasse
the main advantage of the pyrolysis-reforming process lies in its ease [254] and corncob [246–248]. The process developed by the research
scaling up with continuous feed. Fig. 12 shows a schematic re- group headed by Prof. Yoshikawa consists also in a two-step pyrolysis-
presentation of the two-step pyrolysis-reforming process. reforming reactor, in which pyrolysis and reforming zones are also se-
Although the routes described in previous sections such as direct parate [244,245,255].
gasification or bio-oil reforming have been extensively studied in the Although most studies in the literature are carried out in laboratory-
literature, the operating problems reported by some authors in these scale batch reactors, Xiao et al. [249] worked in a two stage fluidized-
processes [201,236] have made the two-step process an interesting al- fixed bed configuration with continuous biomass in the feed, with the
ternative for H2 production from biomass. pyrolysis step being carried out in a fluidized bed reactor of silica sand
Therefore, the following section focuses on an in-depth review on and the reforming step in a fixed bed reactor on a commercial Ni/Al2O3
this process, in which the most relevant factors affecting product dis- and Ni/BCC (brown coal char) bed (Fig. 13b). Wang et al. [256] used a
tribution will be discussed. Among others, the type of the reactor for similar configuration, in which continuous pyrolysis of cellulose was
both pyrolysis and reforming steps and their configuration plays an carried out at the lower section of the reactor (fluidized bed) and the
essential role for the process. Therefore, a comparison of the most re- upper section is the reactor for the catalytic reforming step (fixed bed).
levant results in the literature using different feeds, catalyst and oper- Koike et al. [29] and Li et al. [257,258] studied the pyrolysis-reforming
ating conditions for the two-stage pyrolysis-reforming process is sum- process in a continuous feed laboratory scale dual-bed reactor, which
marized in Table 4. The results have been arranged according to the includes a primary bed for biomass pyrolysis and a secondary catalyst
reactor configuration used in each study. bed for the reforming of pyrolysis products. A different reactor con-
figuration was used by Efika et al. [250], which was provided by a
screw-kiln reactor for the continuous biomass pyrolysis and a fixed bed
6.1. Reactor configurations reactor for the reforming of pyrolysis vapours (Fig. 13c).
Ma et al. [30] improved this two step process introducing another
The main reactors used in the literature for biomass pyrolysis are stage between the pyrolysis and reforming reactors for gas-solid si-
bubbling fluid beds, circulating fluid beds and transporter beds, ro- multaneous gasification (Fig. 13d). A different reactor configuration has
tating cone, ablative reactor, vacuum reactor, screw kiln reactor and been used by Arregi et al. [31,259] for the pyrolysis-catalytic steam
spouted bed reactor [134]. The kind of pyrolysis reactor and its features reforming of biomass, in which a spouted bed reactor was used for the
are relevant factors for this process, given that they condition the dis- pyrolysis step and the volatile stream leaving this reactor was in-
tribution of pyrolysis products and their yields. Thus, bubbling fluidized troduced in a fluidized bed reforming reactor (Fig. 13e). The use of this
bed reactors are the most developed technology, using sand as flui- configuration has been successfully applied to the valorisation of plas-
dizing solid because it allows an excellent gas-solid contact, and tics [234,236,260,261] and biomass/plastic mixtures [262]. Moreover,
therefore improves heat transfer [134]. Moreover, it should be taken this technology has many possibilities for industrial implementation
into account that the product stream of the pyrolysis step is the feed of due to the suitability of the spouted bed for the pyrolysis step
the subsequent reforming step, which is usually carried out in fixed or

Fig. 12. Schematic representation of the pyrolysis-reforming of biomass.

709
A. Arregi et al. Energy Conversion and Management 165 (2018) 696–719

Table 4
Studies involving different strategies, feeds, catalysts and operating conditions, and main results obtained (maximum H2 production) reported in the literature for
pyrolysis and in-line reforming of biomass.
Reactor configuration Kind of study Biomass Catalyst Operating conditions H2 concentration H2 production Ref.
(vol%) (wt%)a

Fixed bed/fixed bed Modelling study (1 g) Wood sawdust 10 wt% Ni/Al2O3 TP = 300, 400, 500, 600 °C 38.1 2.2 [26]
(40 °C min−1)
TR = 800 °C
Steam flow rate = 4.74 g h−1 (second
step)
Fixed bed/fixed bed Laboratory scale Wood sawdust Nano Fe-Zn/Al2O3 TP = 500 °C (40 °C min−1) 40.0 1.9 [237]
(0.5 g) Zn/Al ratio: 1:1, 1:2, TR = 800 °C
1:3, 1:4 Steam flow rate = 0.05 g min−1
(second step)
Fixed bed/fixed bed Laboratory scale Wood sawdust Ni/CaAlOx TP = 500 °C (40 °C min−1) 45.0 3.1 [27]
(0.5 g) Ca/Al molar ratio: TR = 800 °C
1:3, 1:2, 1:1, 2:1, 3:1 Steam flow rate = 0.05 g min−1
(second step)
Fixed bed/fixed bed Laboratory scale Wood sawdust Nanosized NiZnAlOx TP = 535 °C (40 °C min−1) 48.1 4.0 [238]
(0.8 g) TR = 800 °C
Steam flow rate = 4.74 ml h−1
Fixed bed/fixed bed Laboratory scale (1 g) Wood sawdust Ca promoted Ni-Mg- TP = 550 °C (40 °C min−1) 53.3 4.3 [239]
AlOx TR = 800 °C
Steam flow rate = 4.74 g h−1
Fixed bed/fixed bed Laboratory scale Wood pellets Acid treated tyre char TP = 500 °C (40 °C min−1) 56 6.1 [240]
TR = 700–900 °C
S/B ratio = 1.82, 3.32, 4.32, 6 g g−1
Fixed bed/fixed bed Laboratory scale Cellulose 15 wt% Ni/Al2O3 TP = – – 5.9 [241]
(1.5 g) TR = 800 °C
Steam flow rate = 0, 0.01, 0.02, 0.1,
0.2 g min−1 (second step)
Fixed bed/fixed bed Laboratory scale (4 g) Rice husk/ Dolomite TP = 950 °C (20 °C min−1) 59.1 5.1 [242]
sugar cane 10 wt% Ni/dolomite TR = 950 °C
bagasse/wheat Steam flow rate = 0.1 g min−1
straw (second step)
Steam/biomass ratio = 1.37
Fixed bed/fixed bed Laboratory scale (1 g) Pine wood Ni-brown coal char TP = 700 °C (10 °C min−1) – 10.0 [28]
chip/pig TR = 450–700 °C
manure
compost
Fixed bed/fixed bed Laboratory scale Sewage sludge Commercial Ni/ TP = 900 °C (10 °C min−1) 70 11.6b [243]
Al2O3 TR = 400–750 °C
Space velocity = 4000–8000 h−1
Steam partial pressure = 30–53 kPa
(second stage)
Fixed bed/fixed bed Bench scale (5 g) Rice husk Raw, char, Fe char, TP = 800 °C 31.5 5.5 [244]
Ni-Fe/char, Ni-Fe TR = 600–900 °C
char and Ni char
Fixed bed/fixed bed Bench scale (20 g) Rice husk RHC Ni TP = 750 °C 24.3 – [245]
TR = 500–900 °C
Fixed bed/fixed bed Laboratory scale (1 g) Corncob Ni – lignite char TP = 900 °C (10 °C min−1) 72.7 12.2b [246]
TR = 500–700 °C
Space velocity = 2400–7200 h−1
Fixed bed/fixed bed Laboratory scale (1 g) Corncob Ni – exchanged resin TP = 900 °C (10 °C min−1) 72.5 12.3b [247]
char TR = 500–700 °C
Space velocity = 2400–7200 h−1
Fixed bed/fixed bed Laboratory scale (1 g) Corncob Limonite TP = 900 °C 70.4 14.8b [248]
TR = 400–750 °C
Space velocity = 2400–7200 h−1
Fluidized bed/fixed Bench scale Pine wood Ni/Al2O3 and Ni- TP = 530–700 °C 60.0 9.3 [249]
bed (1–2 g min−1) chip/pig brown coal char TR = 550–710 °C
manure Space velocity = 9100–11,300 h−1
compost S/C ratio = 0–3 mol/mol (first step)
Screw-kiln/fixed bed Bench scale Wood pellets NiO/Al2O3 TP = 500 °C 44.4 – [250]
(0.24 kg h−1) NiO/CeO2/Al2O3 TR = 760 °C
NiO/SiO2 Biomass feed rate = 0.24 kg h−1
Steam (second step)
Fluidized bed/ Bench scale Timber wood NiO/MgO TP = 600 °C 51.0 7.6 [30]
entrained flow sawdust TG = 700, 750, 800, 850 °C
gasification/fixed TR = 700, 750, 800, 850 °C
bed Steam/biomass ratio = 3 (first step)
Biomass/catalyst ratio = 1
Spouted bed/fluidized Bench scale Pine wood Commercial Ni/ 66.0 11.0 [31]
bed (0.6–1.5 g min−1) sawdust Al2O3 (G90-LDP)
(continued on next page)

710
A. Arregi et al. Energy Conversion and Management 165 (2018) 696–719

Table 4 (continued)

Reactor configuration Kind of study Biomass Catalyst Operating conditions H2 concentration H2 production Ref.
(vol%) (wt%)a

TP = 500 °C
TR = 550, 600, 650, 700 °C
Space time = 2.5, 5, 10, 15, 20,
30 gcat min gvolatiles−1
Steam flow rate = 3 ml min−1 (first
step)
S/C ratio = 3.9, 5.8, 7.7, 9.7

a
H2 production defined as gH2/100 gbiomass.
b
H2 production defined as gH2/100 gbiomass, daf (dry and ash free).

[263–269]. In fact, this reactor has been satisfactorily scaled up for the 600–700 °C range H2 production kept at around 11.0 wt%.
continuous pyrolysis of 25 kg h−1 of biomass [270,271]. The ad- In the same way, by increasing the reforming temperature, the
vantages of the conical spouted bed reactor in relation to other pyr- concentrations of H2 and CO in the gaseous stream increase, which is
olysis technologies has been discussed elsewhere [272]. explained by the endothermic nature of the reforming reactions (for
Thus, some of the technologies currently used for the pyrolysis-re- both oxygenated compounds and CH4). Cao et al. [243] obtained a H2
forming process have some issues to resolve due to the discontinuous concentration of 66.6 vol% at 500 °C on a commercial Ni/Al2O3 cata-
mode of the process. Nevertheless, as described above, great progress lyst, with the maximum value being 70.0 vol% when a temperature of
has been made recently, especially for the implementation of con- 550 °C was used, although the maximum H2 production (11.6 wt%) was
tinuous biomass feed integrated into different kinds of reactors and obtained at a temperature of 650 °C. The importance of the exothermic
configurations, with all these improvements contributing to the suc- WGS reaction should be noted, given that it plays an essential role in
cessful scaling up of the process. this process and is favoured at relatively low temperatures, producing
H2 and CO2 as main products. Thus, Arregi et al. [31] reported similar
6.2. Operating variables H2 concentration (of around 66 vol%) in the 550–700 °C range, al-
though a slight decrease in CO2 and, especially, an increase in CO
The operating conditions play a vital role in pyrolysis and in-line concentration was observed.
reforming of biomass. In this section, the effect of the most significant
operating variables on H2 production (defined as g of H2 per 100 g of 6.2.2. Space time
biomass) will be discussed, i.e., reforming temperature, reforming space The reforming space time is another important factor in order to
time and steam/carbon (S/C) ratio. ensure a tar-free gas product to be used for different applications. When
space time is increased, steam reforming and WGS reactions are en-
6.2.1. Reforming temperature hanced, favouring the formation of H2 and CO2 and hindering that of
The product yields obtained in the reforming step depend especially CO. This conclusion was reported by Xiao et al. [249], who observed a
on the reforming temperature, given that steam reforming of the oxy- slight increase in H2 production from 6.2 to 7.3 wt% and in CO2 from
genated compounds and hydrocarbons (CH4 and light hydrocarbons) 70.0 to 87.1 wt% when space velocity was reduced from 11,000 to
produced from biomass pyrolysis are endothermic reactions, and they 9,200 h−1, whereas CO and CH4 decrease from 23.0 to 19.9 wt% and
are therefore favoured at high temperatures. Nevertheless, the WGS from 9.1 to 2.9 wt%, respectively. Cao et al. [243] also observed that
reaction is an exothermic reaction which is enhanced at low tempera- higher total gas yield and H2 production were obtained and lower coke
tures, and therefore the optimum reforming temperature to enhance H2 was deposited on the catalyst when space time was increased. Higher
production should be found. Consequently, most studies reveal that tar reforming efficiency was reported by Shen et al. [245], which in-
high reforming temperatures enhance steam reforming reactions and creased from 50.1 to 99.8% when catalyst mass was increased from 3 to
lead to high conversion or gas yield and low tar yields [249,258], at the 10 g, using a RHC Ni catalyst and a reforming temperature of 700 °C.
same time as H2 production is also enhanced (Fig. 14). Moreover, this catalyst is highly stable, as evidenced by the almost
Xiao et al. [28] obtained lower tar yields above 600 °C, with the constant tar reforming efficiency (99.6%) after 5 reforming cycles. The
minimum tar content observing at 700 °C, although the highest H2 influence of space time between 2.5 and 30 gcatalyst min gvolatiles−1 using
production of 10.0 wt% was achieved at 650 °C. Higher activity of the a commercial Ni/Al2O3 catalyst was studied by Arregi et al. [31], who
catalyst was also reported by Koike et al. [29] when temperature was reported a high H2 production of 11.7 wt% by displacing the WGS re-
increased from 550 to 650 °C, with a considerable reduction in the action and at the same time reforming CH4 and light hydrocarbons.
amount of coke deposited, which is the main deactivation cause of the
catalyst in the pyrolysis-reforming of biomass [219,242,250] and bio- 6.2.3. Steam to C ratio
oil reforming processes [192,195,201,273,274]. Shen et al. [245] also Amongst the different reforming agents, steam is one of the most
reported a slight increase in tar conversion efficiency from 92.3 to used in the literature and the S/C ratio is a relevant factor in reforming
100 wt% using RHC Ni catalyst when temperature was increased in the processes. Efika et al. [250] studied the reforming process without and
500–900 °C range. with steam, and the enhancement of gas yield with steam addition was
Concerning H2 production, Waheed and Williams [242] used a high reported, increasing from 40.5 to 45.3 wt% without catalyst and to
reforming temperature (950 °C) for the pyrolysis-reforming of rice husk, 54.0 wt% when NiO/SiO2 catalyst was used. Moreover, H2 concentra-
sugar cane bagasse and wheat straw, obtaining a maximum H2 pro- tion increased from 14.0 (sand without steam) to 18.2 vol% (sand with
duction of 5.1 wt% when rice husk and wheat straw were fed, whereas steam) and to 44.4 vol% when NiO/Al2O3 catalyst was used. Cao et al.
4.9 wt% was obtained with sugar cane bagasse on 10 wt% Ni/dolomite [243] also studied the effect of steam partial pressure (SPP) between 30
catalyst, with these values being higher than those obtained with only and 53 kPa, obtaining a higher H2 yield and lower carbon deposition at
dolomite. When pine wood was valorised by Arregi et al. [31] by pyr- high steam partial pressure values.
olysis-reforming strategy, an increase in H2 yield was reported when Moreover, higher H2 and CO2 and lower CO and CH4 yields are
reforming temperature was increased from 550 to 600 °C, whereas in obtained by increasing S/C ratio due especially to the displacement of

711
A. Arregi et al. Energy Conversion and Management 165 (2018) 696–719

Fig. 13. Reactor configurations for pyrolysis and in-line catalytic steam reforming of biomass: (a) two-stage fixed bed reactor; (b) fluidized – fixed bed reactor; (c)
screw kiln – fixed bed reactor; (d) fluidized bed, entrained flow and fixed bed reactor, and; (e) spouted bed – fluidized bed reactor.

90.1%, whereas CO decreased from 14.4 to 8.9% in the same S/C ratio
range. Zou et al. [241] studied the influence of steam addition between
0 and 0.2 g min−1, obtaining a maximum H2 production of 5.9 wt%
between 0.05 and 0.1 g min−1 of steam addition, without any im-
provement above this value.
Thus, when steam to C ratio is increased, higher H2 production in
the gaseous fraction and lower tar content (non-converted product) are
obtained, given that steam enhances both reforming and WGS reac-
tions, although higher S/C ratio requires higher external heat supply
[249]. Therefore, the optimization of this factor is necessary in order to
obtain an economically viable process.

6.3. Catalysts

Although different catalysts can be used for reforming of oxyge-


nated compounds, the most used ones are those based on Ni and Co due
mainly to its low cost in relation to noble metals, such as, Pt, Pd, Ru or
Rh. Efika et al. [250] investigated three different catalysts: NiO/Al2O3,
NiO/CeO2/Al2O3 and NiO/SiO2 (prepared by an incipient wetness
method and by a sol-gel method). The sol-gel NiO/SiO2 catalyst pro-
Fig. 14. Effect of temperature on hydrogen production in pyrolysis-reforming
duced the highest gas yield, whereas NiO/CeO2/Al2O3 was the catalyst
studies. Xiao et al., 2013 [28]; Ma et al., 2014 [30]; Arregi et al., 2016 [31].
with the lowest gas production. Nevertheless, NiO/Al2O3 was the cat-
alyst giving the highest H2 concentration of 44.4 vol%. Koike et al. [29]
the WGS reaction equilibrium. A significant effect was reported by compared three different catalysts and the lower coke formation was
Arregi et al. [31] when they increased S/C ratio between 3.9 and 9.7, observed for Ni + MnOx/Al2O3 followed by Ni + CeO2/Al2O3 and Ni/
i.e., H2 yield increased from 89.2 to 94.2% and CO2 yield from 84.0 to Al2O3 catalysts. Li et al. [258] investigated Ni-Cu/Mg/Al bimetallic

712
A. Arregi et al. Energy Conversion and Management 165 (2018) 696–719

catalysts with different Cu/Ni ratios, obtaining higher activity and

reduce its deactivation, improve its stability


Optimization of conditions in pyrolysis and
lower deactivation by coke when Cu/Ni ratio was 0.25, with the results

Small units in centralized plants in the


Two-step biomass pyrolysis-reforming

Optimize the catalyst composition to


Lower temperature than gasification
being better compared to monometallic Ni/Mg/Al and Cu/Mg/Al cat-

regions where biomass is available


alysts. Chen et al. [237] studied Fe-Zn/Al2O3 nanocatalysts with dif-

Avoids bio-oil feeding problems


ferent Zn/Al ratios (1:1, 1:2, 1:3 and 1:4), and a maximum H2 con-

Tar-free product is obtained

and catalyst regeneration


centration and production of 40 vol% and 1.9 wt%, respectively, were

catalytic reforming steps


Continuous regime: 10 g

Scale up of the process


obtained when 1:1 ratio was used. They have also investigated Ni/
CaAlOx catalysts with different Ca/Al molar ratios (1:3, 1:2, 1:1, 2:1

Batch regime: 5 g
and 3:1) [27] and H2 concentration and production of 45 vol% and
3.1 wt%, respectively, were obtained when a Ca/Al molar ratio of 1:2
was used. Dong et al. [238] synthesized NiZnAlOx catalysts with dif-
ferent Ni molar fractions (5, 10, 15, 25 and 35%), obtaining the max-
imum H2 production (4.0 wt%) when 35% Ni was used.

Lower energy efficiency and higher


Concerning catalyst supports, although Al2O3 is commonly used,

coupled with a centralized bio-oil

Higher energy density of the bio-


different kinds of supports can be found in the literature. Li et al. [257]

Decentralized pyrolysis plant


reported a higher activity for Co catalyst when they are supported on

Technological development
slurry compared to bio-oil
BaAl12O19 than on other supports, such as Al2O3, ZrO2, SiO2, MgO and

Reduce energetic costs


TiO2 due to its high dispersion. Moreover, Waheed et al. [275] used a

Raw bio-oil feeding


Bio-oil gasification

gasification plant
10 wt% Ni-dolomite catalyst, obtaining a maximum H2 production of
6.1 wt% when rice husk was fed. Ye et al. [276] studied size-confined Ni
catalysts on acidic MCM-41 supports and obtained a maximum H2
production of 4.3 wt%. Jin et al. [239] studied the effect calcium ad-

costs
3a g
dition has on Ni-Mg-AlOx catalyst and observed that H2 production
increased from 2.1 to 3.6 wt%. Santamaria et al. [277] compared five

reduce its deactivation, improve its stability


Decentralized pyrolysis plant coupled with
different supports (Al2O3, SiO2, MgO, TiO2 and ZrO2) and Al2O3, MgO

Feeding lost in bio-oil condensation and


and ZrO2 were those leading to the most encouraging results.

Optimize the catalyst composition to


a centralized bio-oil reforming plant

Lower temperature than gasification


Higher energy density of the bio-oil
Furthermore, the use of char is gaining increasing attention recently
and several studies using it as support have been published. Thus, Xiao
et al. [249] compared a commercial Ni/Al2O3 catalyst and Ni/BCC

and catalyst regeneration


Comparison study of different strategies for hydrogen production from biomass and bio-oil by thermochemical processes.

Scale up of the process


(brown coal char), and higher H2, CO and CH4 yields, lower tar content

compared to biomass

Raw bio-oil feeding


volatilization steps
and higher resistance to coke deposition were observed when Ni/BCC
Raw bio-oil: 10a g
Bio-oil reforming

catalyst was used. Yao et al. [278] used biochar as a catalyst in the

(600–800 °C)
pyrolysis-reforming of wheat straw, obtaining a maximum H2 con-
centration and production of 64.02 vol% and 9.21 wt%, respectively.
The group headed by Prof. Yoshikawa also studied Ni-Fe catalysts [244]
and Ni catalysts [245,255] supported on rice husk char (RHC). They
reported higher tar removal efficiency with RHC Ni and RHC Ni-B

Technological development
Treatment of biomass with

High energy consumption


Unfeasible at large scale

catalysts, 96.9 and 98.6%, respectively, compared to Ni-Fe char cata-


Difficulties in feedstock

Reduce energetic costs


high moisture content
Without catalyst: 2 g

lyst, for which a maximum efficiency of 92.3% was achieved. It should


Supercritical water

With catalyst: 4 g

be pointed out that the recovery of the char from the pyrolysis step is of
great interest for the viability of the process.
gasification

pumping

7. Main remarks and conclusions

Table 5 compares different routes for H2 production from biomass


Unsuitable tar content even with secondary

and bio-oil by thermochemical processes. Considering the different re-


Centralized plants in the regions where

Optimize the catalyst composition to

actor configurations, plant size, operating conditions and catalysts used


catalysts for its further utilization in

Reduce energetic and material costs


reduce its deactivation, improve its
stability and catalyst regeneration
High temperatures (700–1200 °C)

in the processes studied in the literature, it is difficult to compare the H2


Calculated considering a bio-oil yield of 75 wt% [134].

production obtained by the different routes. In addition, it should be


Biomass steam gasification

pointed out that the values of H2 production mentioned in the previous


Decrease the tar content

sections have been calculated based using different feeds (based on


Without catalyst: 4 g

biomass is available

synthesis processes

100 g of biomass, bio-oil or organics). Consequently, an average H2


With catalyst: 7 g

production (wt%) based on 100 g of biomass has been calculated for all
the processes using the results obtained in the literature. In the pro-
cesses in which bio-oil is in the feed, H2 production has been calculated
considering a bio-oil yield of 75 wt% in the pyrolysis step [134]; that is,
the bio-oil lost in the condensation and vaporization processes not been
Advantages and disadvantages
Average H2 production (wt%,

considered in the calculation, which would decrease the production of


H2 in these processes. Moreover, biomass heterogeneity is a fact to be
g H2/100 g biomass)

taken into account. Thus, biomass type, particle size, moisture content
and ash composition are different in each study and so the values
showed in Table 5 are average values calculated from the literature. As
observed in Table 5, supercritical water gasification and bio-oil gasifi-
Challenges
Strategy

cation are the routes in which the lowest H2 production have been
Table 5

obtained, 4 g and 3 g of H2 per 100 g of biomass, respectively. It should


a

be pointed out that these routes have been scarcely studied in the

713
A. Arregi et al. Energy Conversion and Management 165 (2018) 696–719

literature, and therefore their technology is less developed than those of (CTQ2016-75535-R (AEI/FEDER, UE) and CTQ2015-69436-R
the other routes. Although supercritical water gasification can treat (MINECO/FEDER, UE)), the Basque Government (IT748-13) and the
biomass with high moisture content, this process has several drawbacks University of the Basque Country (UFI 11/39). Aitor Arregi thanks the
involving biomass feeding and high energy costs, which makes it un- University of the Basque Country for his postgraduate grant (UPV/EHU
feasible at large scale. In the case of bio-oil gasification, the higher 2017).
energy density of the bio-slurry compared to bio-oil is an interesting
advantage, although bio-oil feeding problems, low energy efficiency References
and high energy costs explain the low technological development of this
process. [1] Sanna A. Advanced biofuels from thermochemical processing of sustainable bio-
Biomass steam gasification is one of the main thermochemical mass in Europe. Bioenergy Res 2014;7:36–47.
[2] Popp J, Lakner Z, Harangi-Rákos M, Fári M. The effect of bioenergy expansion:
routes studied in the literature, achieving an average H2 production of food, energy, and environment. Renew Sustain Energy Rev 2014;32:559–78.
4 g (without catalyst) and 7 g (with catalyst) per 100 g of biomass. The [3] Somerville C, Youngs H, Taylor C, Davis SC, Long SP. Feedstocks for lignocellulosic
need of a secondary catalyst in the gasification process is noteworthy, biofuels. Science 2010;329:790–2.
[4] Melero JA, Iglesias J, Garcia A. Biomass as renewable feedstock in standard re-
as evidenced by the low H2 production obtained with a primary cata- finery units. Feasibility, opportunities and challenges. Energy Environ Sci
lyst, which complicates the gasification process and increases energy 2012;5:7393–420.
costs due to the high temperatures needed in both steps. The major [5] Parajuli R, Dalgaard T, Jørgensen U, Adamsen APS, Knudsen MT, Birkved M, et al.
Biorefining in the prevailing energy and materials crisis: a review of sustainable
challenge of biomass gasification is tar reduction, given that tar causes pathways for biorefinery value chains and sustainability assessment methodolo-
serious problems for syngas further use in synthesis, even when sec- gies. Renew Sustain Energy Rev 2015;43:244–63.
ondary catalysts are used. Therefore, optimization of the catalyst is [6] Balat H, Kirtay E. Hydrogen from biomass – present scenario and future prospects.
Int J Hydrogen Energy 2010;35:7416–26.
imperative in order to improve its stability and regeneration.
[7] Hydrogen Council. Hydrogen scaling up. A sustainable pathway for the global
Concerning raw bio-oil steam reforming, which has also been ex- energy transition; 2017.
tensively studied in the literature, an average H2 production of 10 g per [8] Levin DB, Chahine R. Challenges for renewable hydrogen production from bio-
100 g of biomass has been achieved. The higher energy density of the mass. Int J Hydrogen Energy 2010;35:4962–9.
[9] International Energy Agency (IEA). Technology roadmap: hydrogen and fuel cells.
bio-oil compared to biomass and the lower temperatures needed for its Paris; 2015.
reforming compared to biomass gasification make it an interesting [10] Parthasarathy P, Narayanan KS. Hydrogen production from steam gasification of
strategy for bio-oil valorisation, although the scaling up of the process is biomass: influence of process parameters on hydrogen yield – a review. Renewable
Energy 2014;66:570–9.
its major challenge nowadays. Moreover, as mentioned above, bio-oil [11] Umeki K, Yamamoto K, Namioka T, Yoshikawa K. High temperature steam-only
feeding still involves drawbacks because part of the bio-oil is lost in the gasification of woody biomass. Appl Energy 2010;87:791–8.
condensation and volatilization steps. Comparing both strategies, bio- [12] Ogi T, Nakanishi M, Fukuda Y, Matsumoto K. Gasification of oil palm residues
(empty fruit bunch) in an entrained-flow gasifier. Fuel 2013;104:28–35.
mass gasification is carried out in large centralized plants located in the [13] Hernández JJ, Aranda G, Barba J, Mendoza JM. Effect of steam content in the air-
regions where biomass is available, whereas in the bio-oil reforming steam flow on biomass entrained flow gasification. Fuel Process Technol
strategy, biomass pyrolysis is carried out in decentralized small pyr- 2012;99:43–55.
[14] Iovane P, Donatelli A, Molino A. Influence of feeding ratio on steam gasification of
olysis plants and the bio-oil is transported to centralized bio-oil re- palm shells in a rotary kiln pilot plant. Experimental and numerical investigations.
forming plants, thereby reducing the transportation costs due to the Biomass Bioenergy 2013;56:423–31.
higher energy density of the bio-oil compared to biomass. Nevertheless, [15] Erkiaga A, Lopez G, Amutio M, Bilbao J, Olazar M. Steam gasification of biomass in
a conical spouted bed reactor with olivine and γ-alumina as primary catalysts. Fuel
the technology for biomass gasification is more developed than that for
Process Technol 2013;116:292–9.
bio-oil reforming, which is still in a preliminary phase. [16] Michel R, Rapagnà S, Burg P, Mazziotti di Celso G, Courson C, Zimny T, et al.
Furthermore, the results obtained in the two-step pyrolysis-re- Steam gasification of Miscanthus X Giganteus with olivine as catalyst production
forming of biomass are encouraging as an alternative way of biomass of syngas and analysis of tars (IR, NMR and GC/MS). Biomass Bioenergy
2011;35:2650–8.
gasification and bio-oil reforming for H2 production, achieving an [17] Di Carlo A, Borello D, Sisinni M, Savuto E, Venturini P, Bocci E, et al. Reforming of
average H2 production of 10 g per 100 g of biomass. This route allows tar contained in a raw fuel gas from biomass gasification using nickel-mayenite
optimizing both pyrolysis and reforming steps and avoids the problems catalyst. Int J Hydrogen Energy 2015;40:9088–95.
[18] Koppatz S, Pfeifer C, Hofbauer H. Comparison of the performance behaviour of
related to raw bio-oil feeding. The main advantage of this process is that silica sand and olivine in a dual fluidised bed reactor system for steam gasification
a free-tar product is obtained, which is the biggest challenge of biomass of biomass at pilot plant scale. Chem Eng J 2011;175:468–83.
gasification. Moreover, it is a good strategy in the areas in which bio- [19] Basagiannis AC, Verykios XE. Catalytic steam reforming of acetic acid for hy-
drogen production. Int J Hydrogen Energy 2007;32:3343–55.
mass is available, although there are still issues to be resolved. The [20] Kechagiopoulos PN, Voutetakis SS, Lemonidou AA, Vasalos IA. Hydrogen pro-
deactivation of the catalyst and, especially, its stability are essential in duction via reforming of the aqueous phase of bio-oil over Ni/olivine catalysts in a
order to scale up this process, although few studies in the literature deal spouted bed reactor. Ind Eng Chem Res 2009;48:1400–8.
[21] Medrano JA, Oliva M, Ruiz J, García L, Arauzo J. Hydrogen from aqueous fraction
with the deactivation of the catalyst and its influence on product dis-
of biomass pyrolysis liquids by catalytic steam reforming in fluidized bed. Energy
tribution. Moreover, although different kinds of metal phases, supports 2011;36:2215–24.
and promoters have been used, there are no studies in which their re- [22] Remiro A, Valle B, Aguayo AT, Bilbao J, Gayubo AG. Steam reforming of raw bio-
oil in a fluidized bed reactor with prior separation of pyrolytic lignin. Energy Fuels
generation by coke combustion or steam gasification has been carried
2013;27:7549–59.
out, with these aspects being highly relevant for future studies. [23] Remón J, Broust F, Volle G, García L, Arauzo J. Hydrogen production from pine
Therefore, the two-step process is a different route for H2 produc- and poplar bio-oils by catalytic steam reforming. Influence of the bio-oil compo-
tion, which may be a good alternative for gasification and bio-oil re- sition on the process. Int J Hydrogen Energy 2015;40:5593–608.
[24] Seyedeyn-Azad F, Abedi J, Sampouri S. Catalytic steam reforming of aqueous
forming processes. Moreover, easy scaling up is the main advantage of phase of bio-oil over Ni-based alumina-supported catalysts. Ind Eng Chem Res
the pyrolysis and in-line reforming process, which in turn avoids 2014;53:17937–44.
transportation of the bio-oil needed in the bio-oil reforming process. [25] Xie H, Yu Q, Zuo Z, Han Z, Yao X, Qin Q. Hydrogen production via sorption-
enhanced catalytic steam reforming of bio-oil. Int J Hydrogen Energy
Nevertheless, the use of one or another strategy depends on the area in 2016;41:2345–53.
which it will be installed, their technological development and the [26] Olaleye AK, Adedayo KJ, Wu C, Nahil MA, Wang M, Williams PT. Experimental
measures taken in order to resolve the problems mentioned above. study, dynamic modelling, validation and analysis of hydrogen production from
biomass pyrolysis/gasification of biomass in a two-stage fixed bed reaction system.
Fuel 2014;137:364–74.
Acknowledgments [27] Chen F, Wu C, Dong L, Vassallo A, Williams PT, Huang J. Characteristics and
catalytic properties of Ni/CaAlOx catalyst for hydrogen-enriched syngas produc-
tion from pyrolysis-steam reforming of biomass sawdust. Appl Catal B
This work was carried out with financial support from the Ministry 2016;183:168–75.
of Economy and Competitiveness of the Spanish Government

714
A. Arregi et al. Energy Conversion and Management 165 (2018) 696–719

[28] Xiao X, Cao J, Meng X, Le DD, Li L, Ogawa Y, et al. Synthesis gas production from [62] Rapagnà S, Virginie M, Gallucci K, Courson C, Di Marcello M, Kiennemann A, et al.
catalytic gasification of waste biomass using nickel-loaded brown coal char. Fuel Fe/olivine catalyst for biomass steam gasification: preparation, characterization
2013;103:135–40. and testing at real process conditions. Catal Today 2011;176:163–8.
[29] Koike M, Ishikawa C, Li D, Wang L, Nakagawa Y, Tomishige K. Catalytic perfor- [63] Rapagnà S, Gallucci K, Foscolo PU. Olivine, dolomite and ceramic filters in one
mance of manganese-promoted nickel catalysts for the steam reforming of tar from vessel to produce clean gas from biomass. Waste Manage (Oxford)
biomass pyrolysis to synthesis gas. Fuel 2013;103:122–9. 2018;71:792–800.
[30] Ma Z, Zhang S, Xie D, Yan Y. A novel integrated process for hydrogen production [64] Rapagnà S, Gallucci K, Di Marcello M, Foscolo PU, Nacken M, Heidenreich S, et al.
from biomass. Int J Hydrogen Energy 2014;39:1274–9. First Al2O3 based catalytic filter candles operating in the fluidized bed gasifier
[31] Arregi A, Lopez G, Amutio M, Barbarias I, Bilbao J, Olazar M. Hydrogen produc- freeboard. Fuel 2012;97:718–24.
tion from biomass by continuous fast pyrolysis and in-line steam reforming. RSC [65] Delgado J, Aznar MP, Corella J. Calcined dolomite, magnesite, and calcite for
Adv 2016;6:25975–85. cleaning hot gas from a fluidized bed biomass gasifier with steam: life and use-
[32] Zhang W. Automotive fuels from biomass via gasification. Fuel Process Technol fulness. Ind Eng Chem Res 1996;35:3637–43.
2010;91:866–76. [66] Rapagnà S, Jand N, Foscolo PU. Catalytic gasification of biomass to produce hy-
[33] Bulushev DA, Ross JRH. Catalysis for conversion of biomass to fuels via pyrolysis drogen rich gas. Int J Hydrogen Energy 1998;23:551–7.
and gasification: a review. Catal Today 2011;171:1–13. [67] Bain RL, Dayton DC, Carpenter DL, Czernik SR, Feik CJ, French RJ, et al.
[34] Mahinpey N, Gomez A. Review of gasification fundamentals and new findings: Evaluation of catalyst deactivation during catalytic steam reforming of biomass-
reactors, feedstock, and kinetic studies. Chem Eng Sci 2016;148:14–31. derived syngas. Ind Eng Chem Res 2005;44:7945–56.
[35] Heidenreich S, Foscolo PU. New concepts in biomass gasification. Prog Energy [68] Yung MM, Magrini-Bair KA, Parent YO, Carpenter DL, Feik CJ, Gaston KR, et al.
Combust Sci 2015;46:72–95. Demonstration and characterization of Ni/Mg/K/AD90 used for pilot-scale con-
[36] Sikarwar VS, Zhao M, Clough P, Yao J, Zhong X, Memon MZ, et al. An overview of ditioning of biomass-derived syngas. Catal Lett 2010;134:242–9.
advances in biomass gasification. Energy Environ Sci 2016;9:2939–77. [69] Pfeifer C, Koppatz S, Hofbauer H. Steam gasification of various feedstocks at a dual
[37] Miao Z, Shastri Y, Grift TE, Hansen AC, Ting KC. Lignocellulosic biomass feedstock fluidised bed gasifier: impacts of operation conditions and bed materials. Biomass
transportation alternatives, logistics, equipment configurations, and modeling. Convers Biorefin 2011;1:39–53.
Biofuel Bioprod Biorefin 2012;6:351–62. [70] Göransson K, Söderlind U, Zhang W. Experimental test on a novel dual fluidised
[38] Mohan D, Pittman Jr CU, Steele PH. Pyrolysis of wood/biomass for bio-oil: a cri- bed biomass gasifier for synthetic fuel production. Fuel 2011;90:1340–9.
tical review. Energy Fuels 2006;20:848–89. [71] Pfeifer C, Rauch R, Hofbauer H. In-bed catalytic tar reduction in a dual fluidized
[39] Di Blasi C. Modeling chemical and physical processes of wood and biomass pyr- bed biomass steam gasifier. Ind Eng Chem Res 2004;43:1634–40.
olysis. Prog Energy Combust Sci 2008;34:47–90. [72] Berdugo Vilches T, Marinkovic J, Seemann M, Thunman H. Comparing active bed
[40] Kan T, Strezov V, Evans TJ. Lignocellulosic biomass pyrolysis: a review of product materials in a dual fluidized bed biomass gasifier: olivine, bauxite, quartz-sand,
properties and effects of pyrolysis parameters. Renew Sustain Energy Rev and ilmenite. Energy Fuels 2016;30:4848–57.
2016;57:126–1140. [73] Zhang Z, Pang S. Experimental investigation of biomass devolatilization in steam
[41] Devi L, Ptasinski KJ, Janssen FJJG, Van Paasen SVB, Bergman PCA, Kiel JHA. gasification in a dual fluidised bed gasifier. Fuel 2017;188:628–35.
Catalytic decomposition of biomass tars: use of dolomite and untreated olivine. [74] Schweitzer D, Gredinger A, Schmid M, Waizmann G, Beirow M, Spörl R, et al.
Renewable Energy 2005;30:565–87. Steam gasification of wood pellets, sewage sludge and manure: Gasification per-
[42] Font Palma C. Modelling of tar formation and evolution for biomass gasification: a formance and concentration of impurities. Biomass Bioenergy 2018;111:308–19.
review. Appl Energy 2013;111:129–41. [75] Pfeifer C, Hofbauer H. Development of catalytic tar decomposition downstream
[43] Peng WX, Wang LS, Mirzaee M, Ahmadi H, Esfahani MJ, Fremaux S. Hydrogen and from a dual fluidized bed biomass steam gasifier. Powder Technol 2008;180:9–16.
syngas production by catalytic biomass gasification. Energy Convers Manage [76] Sansaniwal SK, Pal K, Rosen MA, Tyagi SK. Recent advances in the development of
2017;135:270–3. biomass gasification technology: a comprehensive review. Renew Sustain Energy
[44] Sansaniwal SK, Rosen MA, Tyagi SK. Global challenges in the sustainable devel- Rev 2017;72:363–84.
opment of biomass gasification: an overview. Renew Sustain Energy Rev [77] Molino A, Chianese S, Musmarra D. Biomass gasification technology: the state of
2017;80:23–43. the art overview. J Energy Chem 2016;25:10–25.
[45] Ollero P, Serrera A, Arjona R, Alcantarilla S. The CO2 gasification kinetics of olive [78] Susastriawan AAP, Saptoadi H, Purnomo. Small-scale downdraft gasifiers for
residue. Biomass Bioenergy 2002;24:151–61. biomass gasification: a review. Renew Sustain Energy Rev 2017;76:989–1003.
[46] Lopez G, Alvarez J, Amutio M, Arregi A, Bilbao J, Olazar M. Assessment of steam [79] Alauddin ZABZ, Lahijani P, Mohammadi M, Mohamed AR. Gasification of lig-
gasification kinetics of the char from lignocellulosic biomass in a conical spouted nocellulosic biomass in fluidized beds for renewable energy development: a re-
bed reactor. Energy 2016;107:493–501. view. Renew Sustain Energy Rev 2010;14:2852–62.
[47] Ahmed II, Gupta AK. Kinetics of woodchips char gasification with steam and [80] Bellouard Q, Abanades S, Rodat S. Biomass gasification in an innovative spouted-
carbon dioxide. Appl Energy 2011;88:1613–9. bed solar reactor: experimental proof of concept and parametric study. Energy
[48] Pereira EG, Da Silva JN, De Oliveira JL, MacHado CS. Sustainable energy: a review Fuels 2017;31:10933–45.
of gasification technologies. Renew Sustain Energy Rev 2012;16:4753–62. [81] Lopez G, Cortazar M, Alvarez J, Amutio M, Bilbao J, Olazar M. Assessment of a
[49] Niu Y, Han F, Chen Y, Lyu Y, Wang L. Experimental study on steam gasification of conical spouted with an enhanced fountain bed for biomass gasification. Fuel
pine particles for hydrogen-rich gas. J Energy Inst 2017;90:715–24. 2017;203:825–31.
[50] Luo S, Xiao B, Guo X, Hu Z, Liu S, He M. Hydrogen-rich gas from catalytic steam [82] Yao J, Kraussler M, Benedikt F, Hofbauer H. Techno-economic assessment of hy-
gasification of biomass in a fixed bed reactor: influence of particle size on gasifi- drogen production based on dual fluidized bed biomass steam gasification, biogas
cation performance. Int J Hydrogen Energy 2009;34:1260–4. steam reforming, and alkaline water electrolysis processes. Energy Convers
[51] Luo S, Xiao B, Hu Z, Liu S, Guo X, He M. Hydrogen-rich gas from catalytic steam Manage 2017;145:278–92.
gasification of biomass in a fixed bed reactor: influence of temperature and steam [83] Göransson K, Söderlind U, He J, Zhang W. Review of syngas production via bio-
on gasification performance. Int J Hydrogen Energy 2009;34:2191–4. mass DFBGs. Renew Sustain Energy Rev 2011;15:482–92.
[52] Li J, Xiao B, Yan R, Xu X. Development of a supported tri-metallic catalyst and [84] Wilk V, Hofbauer H. Conversion of mixed plastic wastes in a dual fluidized bed
evaluation of the catalytic activity in biomass steam gasification. Bioresour steam gasifier. Fuel 2013;107:787–99.
Technol 2009;100:5295–300. [85] Corella J, Toledo JM, Molina G. A review on dual fluidized-bed biomass gasifiers.
[53] Fremaux S, Beheshti S, Ghassemi H, Shahsavan-Markadeh R. An experimental Ind Eng Chem Res 2007;46:6831–9.
study on hydrogen-rich gas production via steam gasification of biomass in a re- [86] Kaushal P, Tyagi R. Steam assisted biomass gasification-an overview. Can J Chem
search-scale fluidized bed. Energy Convers Manage 2015;91:427–32. Eng 2012;90:1043–58.
[54] Shie J, Tsou F, Lin K. Steam plasmatron gasification of distillers grains residue [87] Asadullah M. Biomass gasification gas cleaning for downstream applications: a
from ethanol production. Bioresour Technol 2010;101:5571–7. comparative critical review. Renew Sustain Energy Rev 2014;40:118–32.
[55] Wei L, Xu S, Zhang L, Liu C, Zhu H, Liu S. Steam gasification of biomass for hy- [88] Erkiaga A, Lopez G, Amutio M, Bilbao J, Olazar M. Syngas from steam gasification
drogen-rich gas in a free-fall reactor. Int J Hydrogen Energy 2007;32:24–31. of polyethylene in a conical spouted bed reactor. Fuel 2013;109:461–9.
[56] Wei L, Xu S, Liu J, Lu C, Liu S, Liu C. A novel process of biomass gasification for [89] Lopez G, Erkiaga A, Artetxe M, Amutio M, Bilbao J, Olazar M. Hydrogen pro-
hydrogen-rich gas with solid heat carrier: preliminary experimental results. Energy duction by high density polyethylene steam gasification and in-line volatile re-
Fuels 2006;20:2266–73. forming. Ind Eng Chem Res 2015;54:9536–44.
[57] Erkiaga A, Lopez G, Amutio M, Bilbao J, Olazar M. Influence of operating condi- [90] Lopez G, Erkiaga A, Amutio M, Bilbao J, Olazar M. Effect of polyethylene co-
tions on the steam gasification of biomass in a conical spouted bed reactor. Chem feeding in the steam gasification of biomass in a conical spouted bed reactor. Fuel
Eng J 2014;237:259–67. 2015;153:393–401.
[58] Xie Y, Xiao J, Shen L, Wang J, Zhu J, Hao J. Effects of Ca-based catalysts on [91] Carpenter DL, Bain RL, Davis RE, Dutta A, Feik CJ, Gaston KR, et al. Pilot-scale
biomass gasification with steam in a circulating spout-fluid bed reactor. Energy gasification of corn stover, switchgrass, wheat straw, and wood: 1. Parametric
Fuels 2010;24:3256–61. study and comparison with literature. Ind Eng Chem Res 2010;49:1859–71.
[59] Gil J, Corella J, Aznar MP, Caballero MA. Biomass gasification in atmospheric and [92] Wang G, Xu S, Wang C, Zhang J. Biomass gasification and hot gas upgrading in a
bubbling fluidized bed: effect of the type of gasifying agent on the product dis- decoupled dual-loop gasifier. Energy Fuels 2017;31:8181–92.
tribution. Biomass Bioenergy 1999;17:389–403. [93] Franco C, Pinto F, Gulyurtlu I, Cabrita I. The study of reactions influencing the
[60] Michel R, Rapagnà S, Di Marcello M, Burg P, Matt M, Courson C, et al. Catalytic biomass steam gasification process. Fuel 2003;82:835–42.
steam gasification of Miscanthus X giganteus in fluidised bed reactor on olivine [94] Karmakar MK, Datta AB. Generation of hydrogen rich gas through fluidized bed
based catalysts. Fuel Process Technol 2011;92:1169–77. gasification of biomass. Bioresour Technol 2011;102:1907–13.
[61] Rapagnà S, Jand N, Kiennemann A, Foscolo PU. Steam-gasification of biomass in a [95] Gao N, Li A, Quan C. A novel reforming method for hydrogen production from
fluidised-bed of olivine particles. Biomass Bioenergy 2000;19:187–97. biomass steam gasification. Bioresour Technol 2009;100:4271–7.

715
A. Arregi et al. Energy Conversion and Management 165 (2018) 696–719

[96] Song T, Wu J, Shen L, Xiao J. Experimental investigation on hydrogen production [128] Lu YJ, Jin H, Guo LJ, Zhang XM, Cao CQ, Guo X. Hydrogen production by biomass
from biomass gasification in interconnected fluidized beds. Biomass Bioenergy gasification in supercritical water with a fluidized bed reactor. Int J Hydrogen
2012;36:258–67. Energy 2008;33:6066–75.
[97] Devi L, Ptasinski KJ, Janssen FJJG. A review of the primary measures for tar [129] Kruse A. Supercritical water gasification. Biofuel Bioprod Biorefining
elimination in biomass gasification processes. Biomass Bioenergy 2002;24:125–40. 2008;2:415–37.
[98] Anis S, Zainal ZA. Tar reduction in biomass producer gas via mechanical, catalytic [130] Azadi P, Farnood R. Review of heterogeneous catalysts for sub- and supercritical
and thermal methods: a review. Renew Sustain Energy Rev 2011;15:2355–77. water gasification of biomass and wastes. Int J Hydrogen Energy
[99] Abdoulmoumine N, Adhikari S, Kulkarni A, Chattanathan S. A review on biomass 2011;36:9529–41.
gasification syngas cleanup. Appl Energy 2015;155:294–307. [131] Balat M, Balat M, Kirtay E, Balat H. Main routes for the thermo-conversion of
[100] Valderrama Rios ML, González AM, Lora EES, Almazán del Olmo OA. Reduction of biomass into fuels and chemicals. Part 1: pyrolysis systems. Energy Convers
tar generated during biomass gasification: a review. Biomass Bioenergy Manage 2009;50:3147–57.
2018;108:345–70. [132] Anex RP, Aden A, Kazi FK, Fortman J, Swanson RM, Wright MM, et al. Techno-
[101] Orío A, Corella J, Narváez I. Performance of different dolomites on hot raw gas economic comparison of biomass-to-transportation fuels via pyrolysis, gasifica-
cleaning from biomass gasification with air. Ind Eng Chem Res 1997;36:3800–8. tion, and biochemical pathways. Fuel 2010;89:S29–35.
[102] Wilk V, Schmid JC, Hofbauer H. Influence of fuel feeding positions on gasification [133] Wright MM, Daugaard DE, Satrio JA, Brown RC. Techno-economic analysis of
in dual fluidized bed gasifiers. Biomass Bioenergy 2013;54:46–58. biomass fast pyrolysis to transportation fuels. Fuel 2010;89:S2–10.
[103] Brachi P, Chirone R, Miccio F, Miccio M, Picarelli A, Ruoppolo G. Fluidized bed co- [134] Bridgwater AV. Review of fast pyrolysis of biomass and product upgrading.
gasification of biomass and polymeric wastes for a flexible end-use of the syngas: Biomass Bioenergy 2012;38:68–94.
focus on bio-methanol. Fuel 2014;128:88–98. [135] Amutio M, Lopez G, Artetxe M, Elordi G, Olazar M, Bilbao J. Influence of tem-
[104] Narváez I, Orío A, Aznar MP, Corella J. Biomass gasification with air in an at- perature on biomass pyrolysis in a conical spouted bed reactor. Resour Conserv
mospheric bubbling fluidized bed. Effect of six operational variables on the quality Recycling 2012;59:23–31.
of the produced raw gas. Ind Eng Chem Res 1996;35:2110–20. [136] Gollakota ARK, Reddy M, Subramanyam MD, Kishore N. A review on the upgra-
[105] De Lasa H, Salaices E, Mazumder J, Lucky R. Catalytic steam gasification of bio- dation techniques of pyrolysis oil. Renew Sustain Energy Rev 2016;58:1543–68.
mass: catalysts, thermodynamics and kinetics. Chem Rev 2011;111:5404–33. [137] Jahirul MI, Rasul MG, Chowdhury AA, Ashwath N. Biofuels production through
[106] Guan G, Kaewpanha M, Hao X, Abudula A. Catalytic steam reforming of biomass biomass pyrolysis – a technological review. Energies 2012;5:4952–5001.
tar: prospects and challenges. Renew Sustain Energy Rev 2016;58:450–61. [138] Tanksale A, Beltramini JN, Lu GM. A review of catalytic hydrogen production
[107] Shen Y, Yoshikawa K. Recent progresses in catalytic tar elimination during bio- processes from biomass. Renew Sustain Energy Rev 2010;14:166–82.
mass gasification or pyrolysis – a review. Renew Sustain Energy Rev [139] Trane R, Dahl S, Skjøth-Rasmussen MS, Jensen AD. Catalytic steam reforming of
2013;21:371–92. bio-oil. Int J Hydrogen Energy 2012;37:6447–72.
[108] Shahbaz M, yusup S, Inayat A, Patrick DO, Ammar M. The influence of catalysts in [140] Ayalur Chattanathan S, Adhikari S, Abdoulmoumine N. A review on current status
biomass steam gasification and catalytic potential of coal bottom ash in biomass of hydrogen production from bio-oil. Renew Sustain Energy Rev
steam gasification: a review. Renew Sustain Energy Rev 2017;73:468–76. 2012;16:2366–72.
[109] Chan FL, Tanksale A. Review of recent developments in Ni-based catalysts for [141] Li D, Li X, Gong J. Catalytic reforming of oxygenates: state of the art and future
biomass gasification. Renew Sustain Energy Rev 2014;38:428–38. prospects. Chem Rev 2016;116:11529–653.
[110] Miccio F, Piriou B, Ruoppolo G, Chirone R. Biomass gasification in a catalytic [142] Takanabe K, Aika K, Seshan K, Lefferts L. Catalyst deactivation during steam re-
fluidized reactor with beds of different materials. Chem Eng J 2009;154:369–74. forming of acetic acid over Pt/ZrO2. Chem Eng J 2006;120:133–7.
[111] Nanda S, Isen J, Dalai AK, Kozinski JA. Gasification of fruit wastes and agro-food [143] Takanabe K, Aika K, Inazu K, Baba T, Seshan K, Lefferts L. Steam reforming of
residues in supercritical water. Energy Convers Manage 2016;110:296–306. acetic acid as a biomass derived oxygenate: bifunctional pathway for hydrogen
[112] Norouzi O, Safari F, Jafarian S, Tavasoli A, Karimi A. Hydrothermal gasification formation over Pt/ZrO2 catalysts. J Catal 2006;243:263–9.
performance of Enteromorpha intestinalis as an algal biomass for hydrogen-rich gas [144] Basagiannis AC, Verykios XE. Reforming reactions of acetic acid on nickel catalysts
production using Ru promoted Fe–Ni/γ-Al2O3 nanocatalysts. Energy Convers over a wide temperature range. Appl Catal A 2006;308:182–93.
Manage 2017;141:63–71. [145] Bimbela F, Oliva M, Ruiz J, García L, Arauzo J. Hydrogen production by catalytic
[113] Yakaboylu O, Harinck J, Smit KG, De Jong W. Supercritical water gasification of steam reforming of acetic acid, a model compound of biomass pyrolysis liquids. J
biomass: a literature and technology overview. Energies 2015;8:859–94. Anal Appl Pyrol 2007;79:112–20.
[114] Matsumura Y, Minowa T, Potic B, Kersten SRA, Prins W, Van Swaaij WPM, et al. [146] Bimbela F, Chen D, Ruiz J, García L, Arauzo J. Ni/Al coprecipitated catalysts
Biomass gasification in near- and super-critical water: status and prospects. modified with magnesium and copper for the catalytic steam reforming of model
Biomass Bioenergy 2005;29:269–92. compounds from biomass pyrolysis liquids. Appl Catal B 2012;119–120:1–12.
[115] Guo Y, Wang SZ, Xu DH, Gong YM, Ma HH, Tang XY. Review of catalytic super- [147] Hu X, Lu G. Comparative study of alumina-supported transition metal catalysts for
critical water gasification for hydrogen production from biomass. Renew Sustain hydrogen generation by steam reforming of acetic acid. Appl Catal B
Energy Rev 2010;14:334–43. 2010;99:289–97.
[116] Calzavara Y, Joussot-Dubien C, Boissonnet G, Sarrade S. Evaluation of biomass [148] Assaf PGM, Nogueira FGE, Assaf EM. Ni and Co catalysts supported on alumina
gasification in supercritical water process for hydrogen production. Energy applied to steam reforming of acetic acid: representative compound for the aqu-
Convers Manage 2005;46:615–31. eous phase of bio-oil derived from biomass. Catal Today 2013;213:2–8.
[117] Safari F, Salimi M, Tavasoli A, Ataei A. Non-catalytic conversion of wheat straw, [149] Pant KK, Mohanty P, Agarwal S, Dalai AK. Steam reforming of acetic acid for
walnut shell and almond shell into hydrogen rich gas in supercritical water media. hydrogen production over bifunctional Ni-Co catalysts. Catal Today
Chin J Chem Eng 2016;24:1097–103. 2013;207:36–43.
[118] Yanik J, Ebale S, Kruse A, Saglam M, Yüksel M. Biomass gasification in super- [150] Wang S, Zhang F, Cai Q, Li X, Zhu L, Wang Q, et al. Catalytic steam reforming of
critical water: part 1. Effect of the nature of biomass. Fuel 2007;86:2410–5. bio-oil model compounds for hydrogen production over coal ash supported Ni
[119] Lu YJ, Guo LJ, Ji CM, Zhang XM, Hao XH, Yan QH. Hydrogen production by catalyst. Int J Hydrogen Energy 2014;39:2018–25.
biomass gasification in supercritical water: a parametric study. Int J Hydrogen [151] Wang S, Cai Q, Zhang F, Li X, Zhang L, Luo Z. Hydrogen production via catalytic
Energy 2006;31:822–31. reforming of the bio-oil model compounds: acetic acid, phenol and hydro-
[120] Lu Y, Guo L, Zhang X, Ji C. Hydrogen production by supercritical water gasifi- xyacetone. Int J Hydrogen Energy 2014;39:18675–87.
cation of biomass: explore the way to maximum hydrogen yield and high carbon [152] Gil MV, Fermoso J, Rubiera F, Chen D. H2 production by sorption enhanced steam
gasification efficiency. Int J Hydrogen Energy 2012;37:3177–85. reforming of biomass-derived bio-oil in a fluidized bed reactor: an assessment of
[121] Yanik J, Ebale S, Kruse A, Saglam M, Yüksel M. Biomass gasification in super- the effect of operation variables using response surface methodology. Catal Today
critical water: II. Effect of catalyst. Int J Hydrogen Energy 2008;33:4520–6. 2014:19–34.
[122] Ding N, Azargohar R, Dalai AK, Kozinski JA. Catalytic gasification of cellulose and [153] Yang X, Weng Y, Li M, Sun B, Li Y, Wang Y. Enhanced hydrogen production by
pinewood to H2 in supercritical water. Fuel 2014;118:416–25. steam reforming of acetic acid over a Ni catalyst supported on mesoporous MgO.
[123] Jin H, Lu Y, Guo L, Zhang X, Pei A. Hydrogen production by supercritical water Energy Fuels 2016;30:2198–203.
gasification of biomass with homogeneous and heterogeneous catalyst. Adv [154] Chen G, Tao J, Liu C, Yan B, Li W, Li X. Hydrogen production via acetic acid steam
Condens Matter Phys 2014. reforming: a critical review on catalysts. Renew Sustain Energy Rev
[124] Elif D, Nezihe A. Hydrogen production by supercritical water gasification of fruit 2017;79:1091–8.
pulp in the presence of Ru/C. Int J Hydrogen Energy 2016;41:8073–83. [155] Zhao X, Xue Y, Yan C, Huang Y, Lu Z, Wang Z, et al. Promoted activity of porous
[125] Osada M, Yamaguchi A, Hiyoshi N, Sato O, Shirai M. Gasification of sugarcane silica coated Ni/CeO2–ZrO2 catalyst for steam reforming of acetic acid. Int J
bagasse over supported ruthenium catalysts in supercritical water. Energy Fuels Hydrogen Energy 2017;42:21677–85.
2012;26:3179–86. [156] Wang M, Zhang F, Wang S. Effect of La2O3 replacement on γ-Al2O3 supported
[126] Güngören Madenoglu T, Boukis N, Saglam M, Yüksel M. Supercritical water ga- nickel catalysts for acetic acid steam reforming. Int J Hydrogen Energy
sification of real biomass feedstocks in continuous flow system. Int J Hydrogen 2017;42:20540–8.
Energy 2011;36:14408–15. [157] Hu R, Li D, Xue H, Zhang N, Liu Z, Liu Z. Hydrogen production by sorption-en-
[127] Antal Jr. MJ, Allen SG, Schulman D, Xu X, Divilio RJ. Biomass gasification in hanced steam reforming of acetic acid over Ni/CexZr1−xO2-CaO catalysts. Int J
supercritical water. Ind Eng Chem Res 2000;39:4040–53. Hydrogen Energy 2017;42:7786–97.

716
A. Arregi et al. Energy Conversion and Management 165 (2018) 696–719

[158] Zhang F, Wang M, Zhu L, Wang S, Zhou J, Luo Z. A comparative research on the oil aqueous fraction over Ni/CeO2-ZrO2 catalysts. Int J Hydrogen Energy
catalytic activity of La2O3 and γ-Al2O3 supported catalysts for acetic acid steam 2010;35:11693–9.
reforming. Int J Hydrogen Energy 2017;42:3667–75. [189] Ortiz-Toral PJ, Satrio J, Brown RC, Shanks BH. Steam reforming of bio-oil frac-
[159] Wang D, Czernik S, Montané D, Mann M, Chornet E. Biomass to hydrogen via fast tions: effect of composition and stability. Energy Fuels 2011;25:3289–97.
pyrolysis and catalytic steam reforming of the pyrolysis oil or its fractions. Ind Eng [190] Bimbela F, Oliva M, Ruiz J, García L, Arauzo J. Hydrogen production via catalytic
Chem Res 1997;36:1507–18. steam reforming of the aqueous fraction of bio-oil using nickel-based coprecipi-
[160] Wang D, Czernik S, Chornet E. Production of hydrogen from biomass by catalytic tated catalysts. Int J Hydrogen Energy 2013;38:14476–87.
steam reforming of fast pyrolysis oils. Energy Fuels 1998;12:19–24. [191] Liu S, Chen M, Chu L, Yang Z, Zhu C, Wang J, et al. Catalytic steam reforming of
[161] Rioche C, Kulkarni S, Meunier FC, Breen JP, Burch R. Steam reforming of model bio-oil aqueous fraction for hydrogen production over Ni-Mo supported on mod-
compounds and fast pyrolysis bio-oil on supported noble metal catalysts. Appl ified sepiolite catalysts. Int J Hydrogen Energy 2013;38:3948–55.
Catal B 2005;61:130–9. [192] Remiro A, Valle B, Aguayo AT, Bilbao J, Gayubo AG. Operating conditions for
[162] Hu X, Lu G. Investigation of the steam reforming of a series of model compounds attenuating Ni/La2O3-αAl2O3 catalyst deactivation in the steam reforming of bio-
derived from bio-oil for hydrogen production. Appl Catal B 2009;88:376–85. oil aqueous fraction. Fuel Process Technol 2013;115:222–32.
[163] Navarro RM, Guil-Lopez R, Gonzalez-Carballo JM, Cubero A, Ismail AA, Al-Sayari [193] Remón J, Medrano JA, Bimbela F, García L, Arauzo J. Ni/Al-Mg-O solids modified
SA, et al. Bimetallic MNi/Al2O3-La catalysts (M = Pt, Cu) for acetone steam re- with Co or Cu for the catalytic steam reforming of bio-oil. Appl Catal B
forming: role of M on catalyst structure and activity. Appl Catal A 2013;132–133:433–44.
2014;474:168–77. [194] Valle B, Remiro A, Aguayo AT, Bilbao J, Gayubo AG. Catalysts of Ni/α-Al2O3 and
[164] Navarro RM, Guil-Lopez R, Ismail AA, Al-Sayari SA, Fierro JLG. Ni- and PtNi- Ni/La2O3-αAl2O3 for hydrogen production by steam reforming of bio-oil aqueous
catalysts supported on Al2O3 for acetone steam reforming: effect of the mod- fraction with pyrolytic lignin retention. Int J Hydrogen Energy 2013;38:1307–18.
ification of support with Ce, La and Mg. Catal Today 2014;474:168–77. [195] Yao D, Wu C, Yang H, Hu Q, Nahil MA, Chen H, et al. Hydrogen production from
[165] Xie H, Yu Q, Wei M, Duan W, Yao X, Qin Q, et al. Hydrogen production from steam catalytic reforming of the aqueous fraction of pyrolysis bio-oil with modified Ni-Al
reforming of simulated bio-oil over Ce-Ni/Co catalyst with in continuous CO2 catalysts. Int J Hydrogen Energy 2014;39:14642–52.
capture. Int J Hydrogen Energy 2015;40:1420–8. [196] Bimbela F, Ábrego J, Puerta R, García L, Arauzo J. Catalytic steam reforming of the
[166] Xie H, Yu Q, Yao X, Duan W, Zuo Z, Qin Q. Hydrogen production via steam re- aqueous fraction of bio-oil using Ni-Ce/Mg-Al catalysts. Appl Catal B
forming of bio-oil model compounds over supported nickel catalysts. J Energy 2017;209:346–57.
Chem 2015;24:299–308. [197] Czernik S, Evans R, French R. Hydrogen from biomass-production by steam re-
[167] Alberton AL, Souza MMVM, Schmal M. Carbon formation and its influence on forming of biomass pyrolysis oil. Catal Today 2007;129:265–8.
ethanol steam reforming over Ni/Al2O3 catalysts. Catal Today 2007;123:257–64. [198] Wang Z, Pan Y, Dong T, Zhu X, Kan T, Yuan L, et al. Production of hydrogen from
[168] Trane-Restrup R, Dahl S, Jensen AD. Steam reforming of ethanol: effects of support catalytic steam reforming of bio-oil using C12A7-O−–based catalysts. Appl Catal A
and additives on Ni-based catalysts. Int J Hydrogen Energy 2013;38:15105–18. 2007;320:24–34.
[169] González-Gil R, Chamorro-Burgos I, Herrera C, Larrubia MA, Laborde M, Mariño F, [199] Wu C, Huang Q, Sui M, Yan Y, Wang F. Hydrogen production via catalytic steam
et al. Production of hydrogen by catalytic steam reforming of oxygenated model reforming of fast pyrolysis bio-oil in a two-stage fixed bed reactor system. Fuel
compounds on Ni-modified supported catalysts. Simulation and experimental Process Technol 2008;89:1306–16.
study. Int J Hydrogen Energy 2015;40:11217–27. [200] Hou T, Yuan L, Ye T, Gong L, Tu J, Yamamoto M, et al. Hydrogen production by
[170] Quitete CPB, Tavares RPA, Bittencourt RCP, Souza MMVM. Coking study of nickel low-temperature reforming of organic compounds in bio-oil over a CNT-promoting
catalysts using model compounds. Catal Lett 2016;146:1435–44. Ni catalyst. Int J Hydrogen Energy 2009;34:9095–107.
[171] Tan Z, Xu X, Liu Y, Zhang C, Zhai Y, Liu P, et al. Upgrading bio-oil model com- [201] Lan P, Xu Q, Zhou M, Lan L, Zhang S, Yan Y. Catalytic steam reforming of fast
pounds phenol and furfural with in situ generated hydrogen. Environ Prog pyrolysis bio-oil in fixed bed and fluidized bed reactors. Chem Eng Technol
Sustainable Energy 2014;33:751–5. 2010;33:2021–8.
[172] Artetxe M, Nahil MA, Olazar M, Williams PT. Steam reforming of phenol as bio- [202] Xu Q, Lan P, Zhang B, Ren Z, Yan Y. Hydrogen production via catalytic steam
mass tar model compound over Ni/Al2O3 catalyst. Fuel 2016;184:629–36. reforming of fast pyrolysis bio-oil in a fluidized-bed reactor. Energy Fuels
[173] Yan B, Li W, Tao J, Xu N, Li X, Chen G. Hydrogen production by aqueous phase 2010;24:6456–62.
reforming of phenol over Ni/ZSM-5 catalysts. Int J Hydrogen Energy [203] Salehi E, Azad FS, Harding T, Abedi J. Production of hydrogen by steam reforming
2017;42:6674–82. of bio-oil over Ni/Al2O3 catalysts: effect of addition of promoter and preparation
[174] Marquevich M, Czernik S, Chornet E, Montané D. Hydrogen from biomass: steam procedure. Fuel Process Technol 2011;92:2203–10.
reforming of model compounds of fast-pyrolysis oil. Energy Fuels 1999;13:1160–6. [204] Seyedeyn-Azad F, Salehi E, Abedi J, Harding T. Biomass to hydrogen via catalytic
[175] Wu C, Liu R. Carbon deposition behavior in steam reforming of bio-oil model steam reforming of bio-oil over Ni-supported alumina catalysts. Fuel Process
compound for hydrogen production. Int J Hydrogen Energy 2010;35:7386–98. Technol 2011;92:563–9.
[176] Xu Q, Xie D, Wang F, Yan Y. Mechanism of hydrogen production by the catalytic [205] Zhang Y, Li W, Zhang S, Xu Q, Yan Y. Steam reforming of bio-oil for hydrogen
steam reforming of bio-oil. Energy Sources Part A 2013;35:1028–38. production: effect of Ni-Co bimetallic catalysts. Chem Eng Technol
[177] Lan P, Xu Q, Lan L, Ren Z, Zhang S, Yan Y. A model for carbon deposition during 2012;35:302–8.
hydrogen production by the steam reforming of bio-oil. Energy Sources Part A [206] Seyedeyn Azad F, Abedi J, Salehi E, Harding T. Production of hydrogen via steam
2014;36:250–8. reforming of bio-oil over Ni-based catalysts: effect of support. Chem Eng J
[178] Garcia-Garcia I, Acha E, Bizkarra K, Martinez De Ilarduya J, Requies J, Cambra JF. 2012;180:145–50.
Hydrogen production by steam reforming of m-cresol, a bio-oil model compound, [207] Fu P, Yi W, Li Z, Bai X, Zhang A, Li Y, et al. Investigation on hydrogen production
using catalysts supported on conventional and unconventional supports. Int J by catalytic steam reforming of maize stalk fast pyrolysis bio-oil. Int J Hydrogen
Hydrogen Energy 2015;40:14445–55. Energy 2014;39:13962–71.
[179] Mei Y, Wu C, Liu R. Hydrogen production from steam reforming of bio-oil model [208] Lan P, Lan LH, Xie T, Liao AP. The preparation of syngas by the reforming of bio-
compound and byproducts elimination. Int J Hydrogen Energy 2016;41:9145–52. oil in a fluidized-bed reactor. Energy Sources Part A 2014;36:242–9.
[180] Trane-Restrup R, Jensen AD. Steam reforming of cyclic model compounds of bio- [209] Remón J, Broust F, Valette J, Chhiti Y, Alava I, Fernandez-Akarregi AR, et al.
oil over Ni-based catalysts: product distribution and carbon formation. Appl Catal Production of a hydrogen-rich gas from fast pyrolysis bio-oils: comparison be-
B 2015;165:117–27. tween homogeneous and catalytic steam reforming routes. Int J Hydrogen Energy
[181] Yoon SJ, Choi Y, Lee J. Hydrogen production from biomass tar by catalytic steam 2014;39:171–82.
reforming. Energy Convers Manage 2010;51:42–7. [210] Quan C, Xu S, Zhou C. Steam reforming of bio-oil from coconut shell pyrolysis over
[182] Wang L, Hisada Y, Koike M, Li D, Watanabe H, Nakagawa Y, et al. Catalyst Fe/olivine catalyst. Energy Convers Manage 2017;141:40–7.
property of Co-Fe alloy particles in the steam reforming of biomass tar and to- [211] Valle B, Aramburu B, Olazar M, Bilbao J, Gayubo AG. Steam reforming of raw bio-
luene. Appl Catal B 2012;121–122:95–104. oil over Ni/La2O3-αAl2O3: influence of temperature on product yields and catalyst
[183] Artetxe M, Alvarez J, Nahil MA, Olazar M, Williams PT. Steam reforming of dif- deactivation. Fuel 2018;216:463–74.
ferent biomass tar model compounds over Ni/Al2O3 catalysts. Energy Convers [212] Valle B, Aramburu B, Benito PL, Bilbao J, Gayubo AG. Biomass to hydrogen-rich
Manage 2017;136:119–26. gas via steam reforming of raw bio-oil over Ni/La2O3-αAl2O3 catalyst: effect of
[184] Garcia L, French R, Czernik S, Chornet E. Catalytic steam reforming of bio-oils for space-time and steam-to-carbon ratio. Fuel 2018;216:445–55.
the production of hydrogen: effects of catalyst composition. Appl Catal A [213] Remiro A, Valle B, Aramburu B, Aguayo AT, Bilbao J, Gayubo AG. Steam re-
2000;201:225–39. forming of the bio-oil aqueous fraction in a fluidized bed reactor with in situ CO2
[185] Czernik S, French R, Feik C, Chornet E. Hydrogen by catalytic steam reforming of capture. Ind Eng Chem Res 2013;52:17087–98.
liquid byproducts from biomass thermoconversion processes. Ind Eng Chem Res [214] Valle B, Aramburu B, Remiro A, Bilbao J, Gayubo AG. Effect of calcination/re-
2002;41:4209–15. duction conditions of Ni/La2O3-αAl2O3 catalyst on its activity and stability for
[186] Basagiannis AC, Verykios XE. Steam reforming of the aqueous fraction of bio-oil hydrogen production by steam reforming of raw bio-oil/ethanol. Appl Catal B
over structured Ru/MgO/Al2O3 catalysts. Catal Today 2007;127:256–64. 2014;147:402–10.
[187] Li H, Xu Q, Xue H, Yan Y. Catalytic reforming of the aqueous phase derived from [215] Remiro A, Valle B, Oar-Arteta L, Aguayo AT, Bilbao J, Gayubo AG. Hydrogen
fast-pyrolysis of biomass. Renewable Energy 2009;34:2872–7. production by steam reforming of bio-oil/bio-ethanol mixtures in a continuous
[188] Yan C, Cheng F, Hu R. Hydrogen production from catalytic steam reforming of bio- thermal-catalytic process. Int J Hydrogen Energy 2014;39:6889–98.

717
A. Arregi et al. Energy Conversion and Management 165 (2018) 696–719

[216] Lemonidou AA, Kechagiopoulos P, Heracleous E, Voutetakis S. Steam reforming of [244] Shen Y, Zhao P, Shao Q, Ma D, Takahashi F, Yoshikawa K. In-situ catalytic con-
bio-oils to hydrogen. The role of catalysis for the sustainable production of bio- version of tar using rice husk char-supported nickel-iron catalysts for biomass
fuels and bio-chemicals 2013:467–93. pyrolysis/gasification. Appl Catal B 2014;152–153:140–51.
[217] Chen J, Sun J, Wang Y. Catalysts for steam reforming of bio-oil: a review. Ind Eng [245] Shen Y, Chen M, Sun T, Jia J. Catalytic reforming of pyrolysis tar over metallic
Chem Res 2017;56:4627–37. nickel nanoparticles embedded in pyrochar. Fuel 2015;159:570–9.
[218] Yan C, Hu E, Cai C. Hydrogen production from bio-oil aqueous fraction with in situ [246] Wang B, Cao J, Zhao X, Bian Y, Song C, Zhao Y, et al. Preparation of nickel-loaded
carbon dioxide capture. Int J Hydrogen Energy 2010;35:2612–6. on lignite char for catalytic gasification of biomass. Fuel Process Technol
[219] Nahil MA, Wang X, Wu C, Yang H, Chen H, Williams PT. Novel bi-functional Ni- 2015;136:17–24.
Mg-Al-CaO catalyst for catalytic gasification of biomass for hydrogen production [247] Cao J, Liu T, Ren J, Zhao X, Wu Y, Wang J, et al. Preparation and characterization
with in situ CO2 adsorption. RSC Adv 2013;3:5583–90. of nickel loaded on resin char as tar reforming catalyst for biomass gasification. J
[220] Rennard D, French R, Czernik S, Josephson T, Schmidt L. Production of synthesis Anal Appl Pyrol 2017;127:82–90.
gas by partial oxidation and steam reforming of biomass pyrolysis oils. Int J [248] Zhao X, Ren J, Cao J, Wei F, Zhu C, Fan X, et al. Catalytic reforming of volatiles
Hydrogen Energy 2010;35:4048–59. from biomass pyrolysis for hydrogen-rich gas production over limonite ore. Energy
[221] Li Q, Hu G. Supply chain design under uncertainty for advanced biofuel produc- Fuels 2017;31:4054–60.
tion based on bio-oil gasification. Energy 2014;74:576–84. [249] Xiao X, Meng X, Le DD, Takarada T. Two-stage steam gasification of waste biomass
[222] Li Q, Zhang Y, Hu G. Techno-economic analysis of advanced biofuel production in fluidized bed at low temperature: parametric investigations and performance
based on bio-oil gasification. Bioresour Technol 2015;191:88–96. optimization. Bioresour Technol 2011;102:1975–81.
[223] Trippe F, Fröhling M, Schultmann F, Stahl R, Henrich E. Techno-economic as- [250] Efika CE, Wu C, Williams PT. Syngas production from pyrolysis-catalytic steam
sessment of gasification as a process step within biomass-to-liquid (BtL) fuel and reforming of waste biomass in a continuous screw kiln reactor. J Anal Appl Pyrol
chemicals production. Fuel Process Technol 2011;92:2169–84. 2012;95:87–94.
[224] Yao J, Liu J, Hofbauer H, Chen G, Yan B, Shan R, et al. Biomass to hydrogen-rich [251] Alvarez J, Kumagai S, Wu C, Yoshioka T, Bilbao J, Olazar M, et al. Hydrogen
syngas via steam gasification of bio-oil/biochar slurry over LaCo1- xCuxO3 per- production from biomass and plastic mixtures by pyrolysis-gasification. Int J
ovskite-type catalysts. Energy Convers Manage 2016;117:343–50. Hydrogen Energy 2014;39:10883–91.
[225] Braimakis K, Atsonios K, Panopoulos KD, Karellas S, Kakaras E. Economic eva- [252] Kumagai S, Alvarez J, Blanco PH, Wu C, Yoshioka T, Olazar M, et al. Novel Ni-Mg-
luation of decentralized pyrolysis for the production of bio-oil as an energy carrier Al-Ca catalyst for enhanced hydrogen production for the pyrolysis-gasification of a
for improved logistics towards a large centralized gasification plant. Renew biomass/plastic mixture. J Anal Appl Pyrol 2015;113:15–21.
Sustain Energy Rev 2014;35:57–72. [253] Kaewpanha M, Karnjanakom S, Guan G, Hao X, Yang J, Abudula A. Removal of
[226] Zhang Y, Brown TR, Hu G, Brown RC. Comparative techno-economic analysis of biomass tar by steam reforming over calcined scallop shell supported Cu catalysts.
biohydrogen production via bio-oil gasification and bio-oil reforming. Biomass J Energy Chem 2017;26:660–6.
Bioenergy 2013;51:99–108. [254] Jafarian S, Tavasoli A, Karimi A, Norouzi O. Steam reforming of bagasse to hy-
[227] Panigrahi S, Dalai AK, Chaudhari ST, Bakhshi NN. Synthesis gas production from drogen and synthesis gas using ruthenium promoted Ni-Fe/γ-Al2O3 nano-catalysts.
steam gasification of biomass-derived oil. Energy Fuels 2003;17:637–42. Int J Hydrogen Energy 2017;42:5505–12.
[228] Kan T, Xiong J, Li X, Ye T, Yuan L, Torimoto Y, et al. High efficient production of [255] Shen Y, Areeprasert C, Prabowo B, Takahashi F, Yoshikawa K. Metal nickel na-
hydrogen from crude bio-oil via an integrative process between gasification and noparticles in situ generated in rice husk char for catalytic reformation of tar and
current-enhanced catalytic steam reforming. Int J Hydrogen Energy syngas from biomass pyrolytic gasification. RSC Adv 2014;4:40651–64.
2010;35:518–32. [256] Wang Y, Hu X, Song Y, Min Z, Mourant D, Li T, et al. Catalytic steam reforming of
[229] Latifi M, Berruti F, Briens C. Thermal and catalytic gasification of bio-oils in the cellulose-derived compounds using a char-supported iron catalyst. Fuel Process
Jiggle Bed Reactor for syngas production. Int J Hydrogen Energy Technol 2013;116:234–40.
2015;40:5856–68. [257] Li D, Ishikawa C, Koike M, Wang L, Nakagawa Y, Tomishige K. Production of
[230] Van Rossum G, Kersten SRA, Van Swaaij WPM. Catalytic and noncatalytic gasi- renewable hydrogen by steam reforming of tar from biomass pyrolysis over sup-
fication of pyrolysis oil. Ind Eng Chem Res 2007;46:3959–67. ported Co catalysts. Int J Hydrogen Energy 2013;38:3572–81.
[231] Chhiti Y, Salvador S, Commandré J, Broust F, Couhert C. Wood bio-oil non- [258] Li D, Koike M, Chen J, Nakagawa Y, Tomishige K. Preparation of Ni-Cu/Mg/Al
catalytic gasification: influence of temperature, dilution by an alcohol and ash catalysts from hydrotalcite-like compounds for hydrogen production by steam
content. Energy Fuels 2011;25:345–51. reforming of biomass tar. Int J Hydrogen Energy 2014;39:10959–70.
[232] Park Y, Namioka T, Sakamoto S, Min T, Roh S, Yoshikawa K. Optimum operating [259] Arregi A, Lopez G, Amutio M, Artetxe M, Barbarias I, Bilbao J, et al. Role of op-
conditions for a two-stage gasification process fueled by polypropylene by means erating conditions in the catalyst deactivation in the in-line steam reforming of
of continuous reactor over ruthenium catalyst. Fuel Process Technol volatiles from biomass fast pyrolysis. Fuel 2018;216:233–44.
2010;91:951–7. [260] Lopez G, Artetxe M, Amutio M, Alvarez J, Bilbao J, Olazar M. Recent advances in
[233] Wu C, Williams PT. Pyrolysis-gasification of plastics, mixed plastics and real-world the gasification of waste plastics. A critical overview. Renew Sustain Energy Rev
plastic waste with and without Ni-Mg-Al catalyst. Fuel 2010;89:3022–32. 2018;82:576–96.
[234] Barbarias I, Lopez G, Alvarez J, Artetxe M, Arregi A, Bilbao J, et al. A sequential [261] Barbarias I, Lopez G, Artetxe M, Arregi A, Bilbao J, Olazar M. Valorisation of
process for hydrogen production based on continuous HDPE fast pyrolysis and in- different waste plastics by pyrolysis and in-line catalytic steam reforming for hy-
line steam reforming. Chem Eng J 2016;296:191–8. drogen production. Energy Convers Manage 2018;156:575–84.
[235] Barbarias I, Lopez G, Artetxe M, Arregi A, Santamaria L, Bilbao J, et al. Pyrolysis [262] Arregi A, Amutio M, Lopez G, Artetxe M, Alvarez J, Bilbao J, et al. Hydrogen-rich
and in-line catalytic steam reforming of polystyrene through a two-step reaction gas production by continuous pyrolysis and in-line catalytic reforming of pine
system. J Anal Appl Pyrol 2016;122:502–10. wood waste and HDPE mixtures. Energy Convers Manage 2017;136:192–201.
[236] Erkiaga A, Lopez G, Barbarias I, Artetxe M, Amutio M, Bilbao J, et al. HDPE [263] Alvarez J, Lopez G, Amutio M, Bilbao J, Olazar M. Bio-oil production from rice
pyrolysis-steam reforming in a tandem spouted bed-fixed bed reactor for H2 pro- husk fast pyrolysis in a conical spouted bed reactor. Fuel 2014;128:162–9.
duction. J Anal Appl Pyrol 2015;116:34–41. [264] Amutio M, Lopez G, Alvarez J, Olazar M, Bilbao J. Fast pyrolysis of eucalyptus
[237] Chen F, Wu C, Dong L, Jin F, Williams PT, Huang J. Catalytic steam reforming of waste in a conical spouted bed reactor. Bioresour Technol 2015;194:225–32.
volatiles released via pyrolysis of wood sawdust for hydrogen-rich gas production [265] Elordi G, Olazar M, Lopez G, Artetxe M, Bilbao J. Product yields and compositions
on Fe-Zn/Al2O3 nanocatalysts. Fuel 2015;158:999–1005. in the continuous pyrolysis of high-density polyethylene in a conical spouted bed
[238] Dong L, Wu C, Ling H, Shi J, Williams PT, Huang J. Promoting hydrogen pro- reactor. Ind Eng Chem Res 2011;50:6650–9.
duction and minimizing catalyst deactivation from the pyrolysis-catalytic steam [266] Artetxe M, Lopez G, Amutio M, Elordi G, Bilbao J, Olazar M. Light olefins from
reforming of biomass on nanosized NiZnAlOx catalysts. Fuel 2017;188:610–20. HDPE cracking in a two-step thermal and catalytic process. Chem Eng J
[239] Jin F, Sun H, Wu C, Ling H, Jiang Y, Williams PT, et al. Effect of calcium addition 2012;207–208:27–34.
on Mg-AlOx supported Ni catalysts for hydrogen production from pyrolysis-gasi- [267] Alvarez J, Amutio M, Lopez G, Barbarias I, Bilbao J, Olazar M. Sewage sludge
fication of biomass. Catal Today 2018. http://dx.doi.org/10.1016/j.cattod.2018. valorization by flash pyrolysis in a conical spouted bed reactor. Chem Eng J
01.004. 2015;273:173–83.
[240] Al-Rahbi AS, Williams PT. Hydrogen-rich syngas production and tar removal from [268] Lopez G, Olazar M, Amutio M, Aguado R, Bilbao J. Influence of tire formulation on
biomass gasification using sacrificial tyre pyrolysis char. Appl Energy the products of continuous pyrolysis in a conical spouted bed reactor. Energy Fuels
2017;190:501–9. 2009;23:5423–31.
[241] Zou J, Yang H, Zeng Z, Wu C, Williams PT, Chen H. Hydrogen production from [269] Lopez G, Alvarez J, Amutio M, Mkhize NM, Danon B, van der Gryp P, et al. Waste
pyrolysis catalytic reforming of cellulose in the presence of K alkali metal. Int J truck-tyre processing by flash pyrolysis in a conical spouted bed reactor. Energy
Hydrogen Energy 2016;41:10598–607. Convers Manage 2017;142:523–32.
[242] Waheed QMK, Williams PT. Hydrogen production from high temperature pyr- [270] Fernandez-Akarregi AR, Makibar J, Lopez G, Amutio M, Olazar M. Design and
olysis/steam reforming of waste biomass: rice husk, sugar cane bagasse, and wheat operation of a conical spouted bed reactor pilot plant (25 kg/h) for biomass fast
straw. Energy Fuels 2013;27:6695–704. pyrolysis. Fuel Process Technol 2013;112:48–56.
[243] Cao J, Shi P, Zhao X, Wei X, Takarada T. Catalytic reforming of volatiles and [271] Makibar J, Fernandez-Akarregi AR, Amutio M, Lopez G, Olazar M. Performance of
nitrogen compounds from sewage sludge pyrolysis to clean hydrogen and syn- a conical spouted bed pilot plant for bio-oil production by poplar flash pyrolysis.
thetic gas over a nickel catalyst. Fuel Process Technol 2014;123:34–40. Fuel Process Technol 2015;137:283–9.

718
A. Arregi et al. Energy Conversion and Management 165 (2018) 696–719

[272] Lopez G, Artetxe M, Amutio M, Bilbao J, Olazar M. Thermochemical routes for the yield. J Energy Inst 2016;89:657–67.
valorization of waste polyolefinic plastics to produce fuels and chemicals. A re- [276] Ye M, Tao Y, Jin F, Ling H, Wu C, Williams PT, et al. Enhancing hydrogen pro-
view. Renew Sustain Energy Rev 2017;73:346–68. duction from the pyrolysis-gasification of biomass by size-confined Ni catalysts on
[273] Davidian T, Guilhaume N, Iojoiu E, Provendier H, Mirodatos C. Hydrogen pro- acidic MCM-41 supports. Catal Today 2018;307:154–61.
duction from crude pyrolysis oil by a sequential catalytic process. Appl Catal B [277] Santamaria L, Lopez G, Arregi A, Amutio M, Artetxe M, Bilbao J, et al. Influence of
2007;73:116–27. the support on Ni catalysts performance in the in-line steam reforming of biomass
[274] Wu C, Sui M, Yan Y. A comparison of steam reforming of two model bio-oil fast pyrolysis derived volatiles. Appl Catal B 2018;229:105–13.
fractions. Chem Eng Technol 2008;31:1748–53. [278] Yao D, Hu Q, Wang D, Yang H, Wu C, Wang X, et al. Hydrogen production from
[275] Waheed QMK, Wu C, Williams PT. Pyrolysis/reforming of rice husks with a biomass gasification using biochar as a catalyst/support. Bioresour Technol
Ni–dolomite catalyst: influence of process conditions on syngas and hydrogen 2016;216:159–64.

719

Vous aimerez peut-être aussi