Vous êtes sur la page 1sur 12

Fuel Processing Technology 87 (2006) 461 472

www.elsevier.com/locate/fuproc

An overview of hydrogen production from biomass


Meng Ni, Dennis Y.C. Leung *, Michael K.H. Leung, K. Sumathy
Department of Mechanical Engineering, The University of Hong Kong, 7/F Haking Wong Building, Pokfulam Road, Hong Kong, China
Received 15 July 2004; received in revised form 8 November 2004; accepted 12 November 2005

Abstract
Hydrogen production plays a very important role in the development of hydrogen economy. One of the promising hydrogen production
approaches is conversion from biomass, which is abundant, clean and renewable. Alternative thermochemical (pyrolysis and gasification) and
biological (biophotolysis, water gas shift reaction and fermentation) processes can be practically applied to produce hydrogen. This paper gives
an overview of these technologies for hydrogen production from biomass. The future development will also be addressed.
D 2005 Elsevier B.V. All rights reserved.
Keywords: Biomass; Pyrolysis; Gasification; Supercritical water; Fermentation; Biophotolysis

1. Introduction
Dependence on fossil fuels as the main energy sources has
led to serious energy crisis and environmental problems, i.e.
fossil fuel depletion and pollutant emission. It has been
reported that United Arab Emirates, one of the major oil
export countries, would fail to meet the share in the oil and
natural gas demands by 2015 and 2042, respectively [1]. The
fossil fuel resources in Egypt would be exhausted within one to
two decades [2]. In China, Mao claimed that the imported oil
amounted to 31% to meet the energy demand in 2000 and the
demand would reach 45 55% in 2010 [3]. The increasing
energy demands will speed up the exhaustion of the finite fossil
fuel. Moreover, combustion of fossil fuel produces substantial
greenhouse and toxic gases, such as CO2, SO2, NOx and other
pollutants, causing global warming and acid rain.
In response to the two problems stated above, continuous
effort has been made in exploration of clean, renewable
alternatives for a sustainable development. Biomass is one of
the most abundant renewable resources. It is formed by fixing
carbon dioxide in the atmosphere during the process of plant
photosynthesis and, therefore, it is carbon neutral in its
lifecycle. Biomass has been used for centuries. Currently,
biomass contributes about 12% of todays world energy supply,
while in many developing countries it contributes 40 50%
* Corresponding author. Tel.: +852 2859 7911; fax: +852 2858 5415.
E-mail address: ycleung@hkucc.hku.hk (D.Y.C. Leung).
0378-3820/$ - see front matter D 2005 Elsevier B.V. All rights reserved.
doi:10.1016/j.fuproc.2005.11.003

energy supply [4]. Biomass research is recently receiving


increasing attention because of the probable waste-to-energy
application. For instance, 150 GT of vegetable bio-matter
generated globally every year can produce about 1.08  1010
GJ energy [5]. One of the major drawbacks is the low
efficiency of utilizing biomass. In China, biomass is widely
used for cooking and heating through biomass burning with a
thermal efficiency only between 10% and 30%. Alternatively,
converting biomass into gaseous and aqueous fuels, electricity
and especially hydrogen is possibly a more efficient way of
biomass utilization. This paper aims to give an overview of
various methods of producing hydrogen from biomass and
their development potential.
2. Energy from biomass
A variety of biomass resources can be used to convert to
energy. They can be divided into four general categories:
(i) Energy crops: herbaceous energy crops, woody energy
crops, industrial crops, agricultural crops and aquatic
crops.
(ii) Agricultural residues and waste: crop waste and animal
waste.
(iii) Forestry waste and residues: mill wood waste, logging
residues, trees and shrub residues.
(iv) Industrial and municipal wastes: municipal solid waste
(MSW), sewage sludge and industry waste.

462

M. Ni et al. / Fuel Processing Technology 87 (2006) 461 472

The available energy production processes from biomass


can be divided into two general categories: thermochemical
and biological processes. Combustion, pyrolysis, liquefaction
and gasification are the four thermochemical processes. Direct
biophotolysis, indirect biophotolysis, biological water gas
shift reaction, photo-fermentation and dark-fermentation are
the five biological processes. Combustion is the direct
burning of biomass in air to convert the biomass chemical
energy into heat, mechanical power or electricity using
equipment such as stoves, furnaces, boilers or steam turbines,
respectively. As the energy efficiency is low (10 30%) and
the pollutant emissions are the by-products; combustion is not
a suitable hydrogen production for sustainable development.
In biomass liquefaction, biomass is heated to 525 600 K in
water under a pressure of 5 20 MPa in the absence of air.
Solvent or catalyst can be added in the process. The
disadvantages of biomass liquefaction are difficulty to achieve
the operation conditions and low production of hydrogen.
Therefore, liquefaction is not favorable for hydrogen production. Other thermochemical processes (pyrolysis and gasification) and biological processes (biophotolysis, biological
water gas shift reaction and fermentation) are feasible and
receiving much attention in hydrogen production in recent
years.

3. Thermochemical processes for hydrogen production


3.1. Hydrogen from biomass pyrolysis
Pyrolysis is the heating of biomass at a temperature of
650 800 K at 0.1 0.5 MPa in the absence of air to convert
biomass into liquid oils, solid charcoal and gaseous
compounds. Pyrolysis can be further classified into slow
pyrolysis and fast pyrolysis. As the products are mainly
charcoal, slow pyrolysis is normally not considered for
hydrogen production. Fast pyrolysis is a high temperature
process, in which the biomass feedstock is heated rapidly in
the absence of air, to form vapor and subsequently
condensed to a dark brown mobile bio-liquid. The products
of fast pyrolysis can be found in all gas, liquid and solid
phases [6]:
(i) Gaseous products include H2, CH4, CO, CO2 and other
gases depending on the organic nature of the biomass for
pyrolysis.
(ii) Liquid products include tar and oils that remain in liquid
form at room temperature like acetone, acetic acid, etc.
(iii) Solid products are mainly composed of char and almost
pure carbon plus other inert materials.
Although most pyrolysis processes are designed for biofuel production, hydrogen can be produced directly through
fast or flash pyrolysis if high temperature and sufficient
volatile phase residence time are allowed as follows:
Biomass heatYH2 CO CH4 other products

Methane and other hydrocarbon vapors produced can be


steam reformed for more hydrogen production:
CH4 H2 OYCO 3H2

In order to increase the hydrogen production, water gas shift


reaction can be applied as follows:
CO H2 OYCO2 H2 :

Besides the gaseous products, the oily products can also be


processed for hydrogen production [7]. The pyrolysis oil can be
separated into two fractions based on water solubility. The
water-soluble fraction can be used for hydrogen production
while the water-insoluble fraction for adhesive formulation.
The material flow is summarized in Fig. 1. Experimental study
has shown that when Ni-based catalyst is used, the maximum
yield of hydrogen can reach 90%. With additional steam
reforming and water gas shift reaction, the hydrogen yield can
be increased significantly.
Temperature, heating rate, residence time and type of
catalyst used are important pyrolysis process control parameters. In favor of gaseous products especially in hydrogen
production, high temperature, high heating rate and long
volatile phase residence time are required [8]. These parameters
can be regulated by selection among different reactor types and
heat transfer modes, such as gas solid convective heat transfer
and solid solid conductive heat transfer. The heat transfer
modes and features of various reactors are listed in Table 1 [9].
It can be seen from the table that fluidized bed reactor type
exhibits higher heating rates and thus it appears to be the
promising reactor type for hydrogen production from biomass
pyrolysis.
Some inorganic salts, such as chlorides, carbonates and
chromates, exhibit beneficial effect on pyrolysis reaction rate
[10]. As tar is difficult to be gasified, extensive studies on the
catalytic effect of inexpensive dolomite and CaO on the

Biomass

Pyrolysis

Bio-Fuel

Separation

H2O

Catalytic Steam
Reforming

Carbon

Co-Product

Phenolic
Intermediates

Resins

H2 and CO2
Fig. 1. Biomass to hydrogen based on pyrolysis with a co-products strategy [7].

M. Ni et al. / Fuel Processing Technology 87 (2006) 461 472


Table 1
Pyrolysis reactor types, heat transfer modes and typical features [9]
Reactor type Mode of heat transfer Typical features
Ablative

95% conduction
4% convection
1% radiation

Fluidized bed 90% conduction


9% convection
1% radiation

Circulating
fluidized
bed

80% conduction

19% convection

4% conduction

High char abrasion from biomass and


char erosion
Leading to high char in product
Char/solid heat carrier separation
required
Solids recycle required
Increased complexity of system
Maximum particle sizes up to 6 mm
Possible liquids cracking by hot solids
Possible catalytic activity from hot char
Greater reactor wear possible
Low heat transfer rates

95% convection
1% radiation

Particle size limit <2 mm


Limited gas/solid mixing

1% radiation

Entrained
flow

Accepts large size feedstocks


Very high mechanical char abrasion
from biomass
Compact design
Heat supply problematical
Heat transfer gas not required
Particulate transport gas not always
required
High heat transfer rates
Heat supply to fluidizing gas or to bed
directly
Limited char abrasion
Very good solids mixing
Particle size limit <2 mm in smallest
dimension
Simple reactor configuration
High heat transfer rates

decomposition of hydrocarbon compounds in tar have been


conducted [11,12]. The catalytic effects of other catalysts (Nibased catalysts [13], Y-type zeolite [14], K2CO3, Na2CO3 and
CaCO3 [15]) and various metal oxides (Al2O3, SiO2, ZrO2,
TiO2 [16] and Cr2O3 [15]) have also been investigated. Among
the different metal oxides, Al2O3 and Cr2O3 exhibit better
catalytic effect than the others. Among the catalysts, Na2CO3 is
better than K2CO3 and CaCO3. Although noble metals Ru and
Rh are more effective than Ni catalyst and less susceptible to
carbon formation, they are not commonly used due to their
high costs [17].
In order to evaluate hydrogen production through pyrolysis
of various types of biomass, extensive experimental investigations have been conducted in recent years. Agricultural
residues [18], peanut shell [7], post-consumer wastes such as
plastics, trap grease, mixed biomass and synthetic polymers
[19] and rapeseed [20] have been widely tested for pyrolysis
for hydrogen production. In order to solve the problem of
decreasing reforming performance caused by char and coke
deposition on the catalyst surface and in the bed itself, fluidized
catalyst beds are usually used to improve hydrogen production
from biomass-pyrolysis-derived bio-fuel [21,22]. Yeboah et al.
[23] constructed a demonstration plant for hydrogen production
from peanut shells pyrolysis and steam reforming in a fluidized

463

bed reactor and the production rates of 250 kg H2/day was


achieved. Padro and Putsche [24] estimated the hydrogen
production cost of biomass pyrolysis to be in the range of
US$8.86/GJ to US$15.52/GJ depending on the facility size and
biomass type. For comparison, the costs of hydrogen production by wind-electrolysis systems and PV-electrolysis systems
are US$20.2/GJ and US$41.8/GJ, respectively. It can be seen
that biomass pyrolysis is a competitive method for renewable
hydrogen production.
3.2. Hydrogen from biomass gasification
Biomass can be gasified at high temperatures (above 1000
K). The biomass particles undergo partial oxidation resulting in
gas and charcoal production. The charcoal is finally reduced to
form H2, CO, CO2 and CH4. This conversion process can be
expressed as:
Biomass heat steamYH2 CO CO2
CH4 light and heavy hydrocarbons char:

Unlike pyrolysis, gasification of solid biomass is carried out


in the presence of O2. Besides, gasification aims to produce
gaseous products while pyrolysis aims to produce bio-oils and
charcoal. The gases produced can be steam reformed to
produce hydrogen and this process can be further improved
by water gas shift reactions as discussed in the previous
section. The gasification process is applicable to biomass
having moisture content less than 35% [25].
One of the major issues in biomass gasification is to deal
with the tar formation that occurs during the process. The
unwanted tar may cause the formation of tar aerosols and
polymerization to a more complex structure, which are not
favorable for hydrogen production through steam reforming.
Currently, three methods are available to minimize tar
formation: (i) proper design of gasifier, (ii) proper control
and operation and (iii) additives/catalysts.
The operation parameters, such as temperature, gasifying
agent and residence time, play an important role in formation
and decomposition of tar. It has been reported that tar could be
thermally cracked at temperature above 1273 K [26]. The use
of some additives (dolomite, olivine and char) inside the
gasifier also helps tar reduction [27]. When dolomite is used,
100% elimination of tar can be achieved [28]. Catalysts not
only reduce the tar content, but also improve the gas product
quality and conversion efficiency. Dolomite, Ni-based catalysts
and alkaline metal oxides are widely used as gasification
catalysts. Process modifications by two-stage gasification and
secondary air injection in the gasifier are also useful for tar
reduction [13].
Another problem of biomass gasification is the formation
of ash that may cause deposition, sintering, slagging, fouling
and agglomeration [29,30]. To resolve the ash-associated
problems, fractionation and leaching have been employed to
reduce ash formation inside the reactor [30 33]. Though
fractionation is effective for ash removal, it may deteriorate
the quality of the remaining ash. On the other hand, leaching

464

M. Ni et al. / Fuel Processing Technology 87 (2006) 461 472

Table 2
Investigations on biomass gasification for hydrogen production
Feedstock

Reactor type

Catalyst used

Hydrogen production (vol.%)

References

Sawdust

Not known

Na2CO3

[35]

Sawdust
Wood
Sawdust
Not known
Sawdust

Circulating fluidized bed


Fixed bed
Fluidized bed
Fluidized bed
Fluidized bed

Pine sawdust
Bagasse
Cotton stem
Eucalyptus gobulus
Pinus radiata
Sewage sludge
Almond shell

Fluidized bed

Not used
Not used
Not known
Ni
K2CO3
CaO
Na2CO3
Not known

Downdraft
Fluidized bed

Switchgrass

Moving bed

48.31 at 700 -C
55.4 at 800 -C
59.8 at 900 -C
10.5 at 810 -C
7.7 at 550 -C
57.4 at 800 -C
62.1 at 830 -C
11.27 at 964 -C
13.32 at 1008 -C
14.77 at 1012 -C
26 42 at 700 800
29 38 at 700 800
27 38 at 700 800
35 37 at 700 800
27 35 at 700 800
10 11
62.8 at 800 -C
63.7 at 900 -C
27.1

Not known
La Ni Fe
Perovskite
Cu Zn Al

-C
-C
-C
-C
-C

[41]

[43]
[44]
[45]

This reaction is exothermic and high yield of hydrogen can be


achieved at relatively low temperature (923 973 K). As
compared with conventional gasification, the HyPr-RING
process, as illustrated in Fig. 3, can be conducted in a much
simpler manner as the reaction for hydrogen production and
gas separation are carried out in one single reactor at a lower
temperature [48]. This novel gasification process has been
analyzed theoretically and demonstrated experimentally to be a
very efficient technique for hydrogen production from biomass.
When biomass has high moisture content above 35%, it is
likely to gasify biomass in a supercritical water condition.
Under the severe conditions obtained by heating water to a
temperature above its critical temperature (647 K) and
compressing it to a pressure above its critical pressure (22
MPa), biomass is rapidly decomposed into small molecules or
gases in a few minutes at a high efficiency [53]. Supercritical
water gasification is a promising process to gasify biomass
11
10.23

10

Hydrogen Cost ($/GJ)

can remove biomass inorganic fraction, as well as improve


the quality of the remaining ash [30]. More recently,
gasification of leached olive oil waste in a circulating fluidized
bed reactor was reported for gas production that demonstrated
the feasibility of leaching as a pretreatment technique for gas
production [34].
Hydrogen can be produced from the gasification gaseous
products through the same procedure of steam reforming and
water gas shift reaction as discussed in the pyrolysis section.
As the products of gasification are mainly gases, this process is
more favorable for hydrogen production than pyrolysis. In
order to optimize the process for hydrogen production, a
number of efforts have been made by researchers to test
hydrogen production from biomass gasification with various
biomass types and at various operating conditions, as listed in
Table 2 [35 45]. Using a fluidized bed gasifier along with
suitable catalysts, it is possible to achieve hydrogen production
about 60 vol.%. Such high conversion efficiency makes
biomass gasification an attractive hydrogen production alternative. In addition, the costs of hydrogen production by biomass gasification are competitive with natural gas reforming as
illustrated in Fig. 2 [46]. Taking into account the environmental
benefit as well, hydrogen production from biomass gasification
should be a promising option based on both economic and
environmental considerations.
In recent years, a novel gasification method, namely,
Hydrogen Production by Reaction Integrated Novel Gasification (HyPr-RING) was proposed by Lin et al. [47 52]. HyPrRING method is an integration of the water-hydrocarbon
reaction, water gas shift reaction and absorption of CO2 and
other pollutants in a single reactor under both sub-critical and
supercritical water conditions. The main reaction for this new
method can be expressed as:

[36]
[37]
[38]
[39]
[40]

8.76

8.74

Switchgrass

Bagasse

7.76
7.54
7.46

Natural Gas $4.5/GJ


6.67

6
Natural gas $3/GJ
5
500

1000

5.85

1500

2000

Biomass feed to gasifier (tonnes/day)

0
C 2H2 O CaOYCaCO3 2H2 ; DH298
 88kJ =mol

Fig. 2. Estimated cost comparison of hydrogen production by biomass


gasification and natural gas steam reforming [46].

M. Ni et al. / Fuel Processing Technology 87 (2006) 461 472

465

(a) HyPr-RING process


Coal, Heavy oil,
biomass, water

High pressure reactor


(923-973K)

H2

C + 2H2O + CaO CaCO3 + 2H2

Water
clean Up

Sorbent
regenerator

CaCl2, NaS

Ash

CO2

(b) Conventional gasification hydrogen


process for hydrogen production
Sorbent
regenerator

Sorbent
regenerator

CO2

Steam

Coal,
Heavy
Oil,
Biomass

Synthesis
gas, CO,
CO2, H2
H2O,H2S
Gasifier

Clean gas,
CO,CO2,
H2

Gas
clean up

Shift
Reaction
(<673K)

CO2,
H2

CO2
absorptor

H2

> 1573K
C + CO2 2CO
C + H2O CO + H2

Ash
O2
Air

CO + H2O CO2 + H2

Air
Separator

N2
Fig. 3. Comparison of HyPr-RING and conventional gasification hydrogen production process [48].

with high moisture contents due to the high gasification ratio


(100% achievable) and high hydrogen volumetric ratio (50%
achievable).
In recent years, extensive research has been carried out to
evaluate the suitability of various wet biomass gasification in
supercritical water conditions. However, the works have been
mostly on a laboratory scale and in an early development
stage; hence, the principles and basic mechanisms are not
well understood yet. The solubility of biomass components in
hot compressed water has been first studied by Mok et al.
[54]. The results show that in hot compressed waters, about
40 60% of the biomass sample is soluble though the reaction
is maintained slightly below the critical water condition.
Minowa et al. [55 57] reported hydrogen production from
cellulose gasification in hot compressed water (subcritical)
using nickel catalyst. Yu et al. [58] reported that the
gasification of glucose at supercritical water condition, such

as 873 K and 34.5 MPa, was different from the nonsupercritical water condition. One advantage is that, during
gasification, neither tar nor char formation occurs. This early
finding stimulated extensive interests in supercritical water
research. Recent attempts to produce hydrogen from biomass
at supercritical water conditions are summarized in Table 3
[59 62]. Using glucose as a model compound, hydrogen
yield of more than 50 vol.% can be achieved with the use of
proper catalysts in supercritical water condition. Tubular
reactors are widely used in supercritical water gasification
because of their robust structures to withstand high pressure.
Although supercritical water gasification is still at its early
development stage, the technology has already shown its
economic competitiveness with other hydrogen production
methods. Spritzer and Hong [63] have estimated the cost of
hydrogen produced by supercritical water gasification to be
about US$3/GJ (US$0.35/kg).

466

M. Ni et al. / Fuel Processing Technology 87 (2006) 461 472

Table 3
Recent studies on hydrogen production by biomass gasification in supercritical water conditions
Feedstock

Gasifier type

Catalyst used

Temperature and pressure

Glucose
Glucose
Glucose
Glucose
Glucose
Glycerol
Glycerol/methanol
Corn starch
Sawdust/corn
starch mixture
Glucose

Not
Not
Not
Not
Not
Not
Not
Not
Not

Not used
Activated
Activated
Activated
Activated
Activated
Activated
Activated
Activated

600
600
600
550
500
665
720
650
690

Tubular reactor

KOH

600 -C, 25 MPa

Catechol

Tubular reactor

KOH

600 -C, 25 MPa

Sewage
Glucose
Glucose
Glucose
Glucose
Glucose
Sawdust

Autoclave
Tubular reactor
Tubular reactor
Tubular reactor
Tubular reactor
Tubular reactor
Tubular reactor

K2CO3
Not used
Not used
Not used
Not used
Not used
Sodium carboxymethylcellulose (CMC)

450
600
500
550
650
650
650

known
known
known
known
known
known
known
known
known

carbon
carbon
carbon
carbon
carbon
carbon
carbon
carbon

-C,
-C,
-C,
-C,
-C,
-C,
-C,
-C,
-C,

-C,
-C,
-C,
-C,
-C,
-C,
-C,

34.5 MPa
34.5 MPa
25.5 MPa
25.5 MPa
25.5 MPa
28 MPa
28 MPa
28 MPa
28 MPa

31.5 35 MPa
25 MPa
30 MPa
30 MPa
32.5 MPa
30 MPa
22.5 MPa

Hydrogen yield
0.56 mol H2/mol
2.15 mol H2/mol
1.74 mol H2/mol
0.62 mol H2/mol
0.46 mol H2/mol
48 vol.%
64 vol.%
48 vol.%
57 vol.%

References
of
of
of
of
of

feed
feed
feed
feed
feed

59.7 vol.%
(9.1 mol H2/mol glucose)
61.5 vol.%
(10.6 mol H2/mol catechol)
47 vol.%
41.8 vol.%
32.9 vol.%
33.1 vol.%
40.8 vol.%
41.2 vol.%
30.5 vol.%

[59]

[60]

[61]

[62]

Hydrogen production from biomass thermochemical


processes has already been shown to be attractive
economically and demonstrated to be a feasible option.
However, it should be noted that hydrogen gas is normally
produced together with other gas constituents. Thus,
separation and purification of hydrogen gas are required.
Nowadays, several methods, such as CO2 absorption,
drying/chilling and membrane separation [64 65], have
been successfully developed for hydrogen gas purification.
It is expected that biomass thermochemical conversion
processes, especially the newly developed gasification
types, will be available for large-scale hydrogen production
in the near future.

these works are in laboratory scale and the practical applications still need to be demonstrated. Biological hydrogen
production can be classified into five different groups: (i)
direct biophotolysis, (ii) indirect biophotolysis, (iii) biological
water gas shift reaction, (iv) photo-fermentation and (v) darkfermentation [66]. All processes are controlled by the
hydrogen-producing enzymes, such as hydrogenase and
nitrogenase. The major components of nitrogenase are MoFe
protein and Fe protein. Nitrogenase has the ability to use
magnesium adenosine triphosphate (MgATP) and electrons to
reduce a variety of substrates (including protons). This
chemical reaction yields hydrogen production by a nitrogenase-based system [67]:

4. Biological process for hydrogen production

2e 2H 4ATPYH2 4ADP 4Pi

4.1. General aspects

where ADP and Pi refer to adenosine diphosphate and


inorganic phosphate, respectively.
Hydrogenases exist in most of the photosynthetic microorganisms and they can be classified into two categories: (i)
uptake hydrogenases and (ii) reversible hydrogenases. Uptake
hydrogenases, such as NiFe hydrogenases and NiFeSe hydro-

The phenomenon of biological hydrogen production was


observed one century ago. When the oil crisis broke out in
1970s, the technology started receiving attention, especially in
hydrogen production by photosynthetic process. However,

Table 4
Comparative properties of nitrogenase and hydrogenase [51]
Property

Nitrogenase

Hydrogenase

Substrates
Products
Number of proteins
Metal components
Optimum temperature
Optimum pH
Inhibitors
Simulators

ATP, H+ or N2, electrons


H2 or NH+4
Two (Mo Fe and Fe)
Mo, Fe
30 -C (A. vinelandii)
7.1 7.3 (A. vinelandii)
N2 (of H2 production), NH+4 , O2, high N to C ratio
Light

H+, H2
ATP, H+, H2, electrons
One
Ni, Fe, S
55 -C (R. rubrum), 70 -C (R. capsulatus)
6.5 7.5 (R. sulfidophilus)
CO, EDTA (ethylenediaminetetraacetic acid), O2, presence of organic compound
H2 (R. sphaeroides), absence of organic compounds (R. rubrum, R. capsulatus)

M. Ni et al. / Fuel Processing Technology 87 (2006) 461 472

467

genases, act as important catalysts for hydrogen consumption


as follows:

Table 5
Comparison of hydrogen production by wild-type microorganisms and mutants

H2 Y2e 2H

Name of
microorganisms

Hydrogen production
comparison between
mutants and wild-type
microorganisms

References

Azotobacter vinelandii

H2 oxidation/H2 evolution
For wild-type: 46
For mutants: 1.5
Hydrogen from malate
For wild-type: 25%
For mutants: 68%
Mutants show higher
oxygen tolerance
than wild-type Anabaena
hydrogen production not known
Hydrogen production of the
mutants 1.5 times higher than
that of the wild-type
50% more hydrogen production
for the mutants than the wild-type

[70]

Reversible hydrogenases, as indicated by its name, have the


ability to produce H2 as well as consume H2 depending on the
reaction condition,
H2 62e 2H :

Koku et al. [68] have reported that though a variety of


substrates can be used for Rhodobacter sphaeroides growth,
only some of them are suitable for hydrogen production. The
properties of nitrogenase and hydrogenase are summarized in
Table 4.

Rhodobacter capsulatus

Anabaena

Rhodobacter sphaeroides

4.2. Direct biophotolysis


Rhodobacter sphaeroides

Direct biophotolysis of hydrogen production is a biological


process using microalgae photosynthetic systems to convert
solar energy into chemical energy in the form of hydrogen:
solar energy

2H2 O Y 2H2 O2

Two photosynthetic systems are responsible for photosynthesis


process: (i) photosystem I (PSI) producing reductant for CO2
reduction and (ii) photosystem II (PSII) splitting water and
evolving oxygen. In the biophotolysis process, two photons
from water can yield either CO2 reduction by PSI or hydrogen
formation with the presence of hydrogenase. In green plants,
due to the lack of hydrogenase, only CO2 reduction takes place.
On the contrary, microalgaes, such as green algae and
Cyanobacteria (blue-green algae), contain hydrogenase and,
thus, have the ability to produce hydrogen. In this process,
electrons are generated when PSII absorbs light energy. The
electrons are then transferred to the ferredoxin (Fd) using the
solar energy absorbed by PSI. The hydrogenase accepts the
electrons from Fd to produce hydrogen as shown in Fig. 4.
Since hydrogenase is sensitive to oxygen, it is necessary to
maintain the oxygen content at a low level under 0.1% so that
hydrogen production can be sustained [67]. This condition can
be obtained by the use of green algae Chlamydomonas
reinhardtii that can deplete oxygen during oxidative respiration
[69]. However, due to the significant amount of substrate being
respired and consumed during this process, the efficiency is
H2
Solar
Energy

2e-

PSI and
PSII

2H2O

Fd

Hydrogenase

O2
Fig. 4. Schematics of direct biophotolysis.

2H+

[71]

[72]

[73]

[74]

low. Recently, mutants derived from microalgae were reported


to have good O2 tolerance and thus higher hydrogen
production. The works reported on mutants for hydrogen
production are listed in Table 5 [70 74]. It can be seen that,
using mutants for hydrogen production, the efficiency can be
increased significantly.
Benemann [75] estimated the cost of direct biophotolysis for
hydrogen production to be $20/GJ assuming that the capital
cost is about US$60/m2 with an overall solar conversion
efficiency of 10%. Hallenbeck and Benemann [67] performed
similar cost estimation and reported that the capital cost of
US$100/m2. In their estimation, some practical factors were
neglected, such as gas separation and handling.
4.3. Indirect biophotolysis
According to Gaudernack [76], the concept of indirect
biophotolysis involves the following four steps as illustrated
in Fig. 5: (i) biomass production by photosynthesis, (ii)
biomass concentration, (iii) aerobic dark fermentation yielding 4 mol hydrogen/mol glucose in the algae cell, along with
2 mol of acetates, and (iv) conversion of 2 mol of acetates
into hydrogen. In a typical indirect biophotolysis, Cyanobacteria are used to produce hydrogen via the following
reactions [66]:
12H2 O 6CO2 YC6 H12 O6 6O2

10

C6 H12 O6 12H2 OY12H2 6CO2 :

11

Markov et al. [77] investigated the indirect biophotolysis


with Cyanobacterium anabaena variabilis exposed to light
intensities of 45 55 Amol 1 m 2 and 170 180 Amol 1 m 2
in the first stage and second stage, respectively. Photoproduction of hydrogen at a rate of about 12.5 ml H2/gcdw h (cdw:
cell dry weight) was found. In the study on indirect
biophotolysis with Cyanobacterium gloeocapsa alpicola by
Troshina et al. [78], it was found that maintaining the medium

468

M. Ni et al. / Fuel Processing Technology 87 (2006) 461 472

Stage 1
Solar
Energy

CO2
2eFd

PSI and PSII

H2

Solar
Energy
2eFd

Bacterial
Photosystem

Cell
Material

Nitrogenase

4ATP
O2
2H2O
Stage 2

2H+

Organic Acids
H2

Cell
Material

Fermentation

Fig. 6. Photo-fermentation schematics.

2eFd

Hydrogenase
2H+

Fig. 5. Indirect biophotolysis for hydrogen production.

at pH value between 6.8 and 8.3 yielded optimal hydrogen


production. Increasing the temperature from 30 -C to 40 -C can
increase the hydrogen production twice as much. The hydrogen
production rate through indirect biophotolysis is comparable to
hydrogenase-based hydrogen production by green algae. The
estimated overall cost is US$10/GJ of hydrogen [67]. However,
it should be pointed out that indirect biophotolysis technology
is still under active research and development. The estimated
cost is subject to a significant change depending on the
technological advancement.
4.4. Biological water gas shift reaction
Some photoheterotrophic bacteria, such as Rhodospirillum
rubrum can survive in the dark by using CO as the sole carbon
source to generate ATP by coupling the oxidation of CO to the
reduction of H+ to H2 [79]:
CO H2 O6CO2 H2 ; DG0  20:1kJ =mol:

12

In equilibrium, the dominating products are CO2 and H2.


Therefore, this process is favorable for hydrogen production.
Organisms growing at the expense of this process are the
gram-negative bacteria, such as R. rubrum and Rubrivax
gelatinosus, and the gram-positive bacteria, such as Carboxydothermus hydrogenoformans [80]. Under anaerobic conditions, CO induces the synthesis of several proteins, including
CO dehydrogenase, Fe S protein and CO-tolerant hydrogenase. Electrons produced from CO oxidation are conveyed
via the Fe S protein to the hydrogenase for hydrogen
production [81].
Biological water gas shift reaction for hydrogen production
is still under laboratory scale and only few works have been
reported. The common objectives of these works were to
identify suitable microorganisms that had high CO uptake and
to estimate the hydrogen production rate. Kerby et al. [79]
observed that under dark, anaerobic conditions in the presence
of sufficient nickel, the doubling time of R. rubrum was less

than 5 h by the oxidation of CO to CO2 coupled with the


reduction of protons to hydrogen. However, R. rubrum requires
light to grow and hydrogen production is inhibited by medium
CO partial pressure above 0.2 atm. An alternative new
chemoheterotrophic bacterium Citrobacter sp. Y19 was tested
by Jung et al. [82] for hydrogen production using water gas
shift reaction. The maximum hydrogen production activity was
found to be 27 mmol/g cell h, which is about three times higher
than R. rubrum.
Recently, Wolfrum et al. [83] have conducted a detailed
study to compare the biological water gas shift reaction with
conventional water gas shift processes. Their analysis showed
that biological water gas shift process was economically
competitive when the methane concentration was under 3%.
The hydrogen production cost from biological water gas shift
reaction ranged from US$1.75/kg (US$14.6/GJ) to around
US$2.25/kg (US$18.8/GJ) for a methane concentration between 1% and 10%. Compared with thermochemical water
gas shift processes, the cost of biological water gas shift
processes are lower due to the elimination of reformer and
associated equipment.
4.5. Photo-fermentation
Photosynthetic bacteria have the capacity to produce
hydrogen through the action of their nitrogenase using solar
energy and organic acids or biomass. This process is known as
photo-fermentation, as shown in Fig. 6. In recent years, some
attempts have been made for hydrogen production from
industrial and agricultural wastes to effect waste management.
As summarized in Table 6 [84 87], hydrogen can be produced
by photo-fermentation of various types of biomass wastes.
Table 6
Example studies on hydrogen production by photo-fermentation
Biomass type
Lactic acid

Bacteria system

Rhodobacter
immobilized
Lactate feedstock Rhodobacter
Wastewater
Rhodobacter
immobilized
Sugar refinery
Rhodobacter
wastewater
O.U.001

Hydrogen conversion References


efficiency

sphaeroides 86%

[84]

capsulatus 30%
sphaeroides 53%

[85]
[86]

spaeroides 0.005 l H2/h/l


culture

[87]

M. Ni et al. / Fuel Processing Technology 87 (2006) 461 472

H2

469

CO2

Gas Separation

Fermentative
Substrates such as
Biomass, Agricultural
Products or other
organic wastes

Pretreatment

Fermentation

Fig. 7. Hydrogen production by dark fermentation.

However, these processes have three main drawbacks: (i) use


of nitrogenase enzyme with high-energy demand, (ii) low solar
energy conversion efficiency and (iii) demand for elaborate
anaerobic photobioreactors covering large areas [84]. Hence, at
the present time, photo-fermentation process is not a competitive method for hydrogen production.
4.6. Dark fermentation
Fermentation by anaerobic bacteria as well as some
microalgaes, such as green algae on carbohydrate-rich substrates, can produce hydrogen at 30 -C to 80 -C especially in a
dark condition [88]. This process using dark fermentation for
hydrogen production is illustrated in Fig. 7. Unlike a
biophotolysis process that produces only H2, the products of
dark fermentation are mostly H2 and CO2 combined with other
gases, such as CH4 or H2S, depending on the reaction process
and the substrate used. With glucose as the model substrate,
maximum 4 mol H2 is produced per mole glucose when the end
product is acetic acid:
C6 H12 O6 2H2 OY2CH3 COOH 4H2 2CO2

13

When the end product is butyrate, 2 mol H2 is produced:


C6 H12 O6 2H2 OYCH2 CH2 CH2 OOH 2H2 2CO2
14
However, in practice, the 4 mol H2 production/mol glucose
cannot be achieved because the end products normally contain
both acetate and butyrate [89].
The amount of hydrogen production by dark fermentation
highly depends on the pH value, hydraulic retention time
(HRT) and gas partial pressure. For the optimal hydrogen
production, pH should be maintained between 5 and 6 [90
92]. Partial pressure of H2 is yet another important parameter
affecting the hydrogen production. When hydrogen concentration increases, the metabolic pathways shift to produce more
reduced substrates, such as lactate, ethanol, acetone, butanol or
alanine, which in turn decrease the hydrogen production [93].
Besides the pH value and partial pressure, HRT also plays an
important role in hydrogen production. Ueno et al. [94] have
reported that an optimal HRT of 0.5 day could effect maximum

hydrogen production (14 mmol/g carbohydrate) from wastewater by anaerobic microflora in the presence of chemostat
culture. When HRT was increased from 0.5 day to 3 days,
hydrogen production rate was reduced from 198 to 34 mmol
l 1 day 1, while the carbohydrates in the wastewater were
decomposed at an increasing efficiency from 70% to 97%. Due
to the fact that solar radiation is not a requirement, hydrogen
production by dark fermentation does not demand much land
and is not affected by the weather condition. Hence, the
feasibility of the technology yields a growing commercial
value.
5. Conclusions
Hydrogen is recognized as one of the most promising
energy carriers in the future. Many investigations on various
hydrogen production methods have been conducted over the
past several decades. Biomass is potentially a reliable energy
resource for hydrogen production. Biomass is renewable,
abundant and easy to use. Over the life cycle, net CO2
emission is nearly zero due to the photosynthesis of green
plants. The thermochemical pyrolysis and gasification hydrogen production methods are economically viable and will
become competitive with the conventional natural gas reforming method. Biological dark fermentation is also a promising
hydrogen production method for commercial use in the future.
With further development of these technologies, biomass will
play an important role in the development of sustainable
hydrogen economy.
Acknowledgements
The authors gratefully acknowledge the funding support
from the CLP Research Institute Ltd.
References
[1] A. Kazim, T.N. Veziroglu, Utilization of solar-hydrogen energy in the
UAE to maintain its share in the world energy market for the 21st century,
Renewable Energy 24 (2001) 259.
[2] M.A.H. Abdallah, S.S. Asfour, T.N. Veziroglu, Solar-hydrogen energy
system for Egypt, International Journal of Hydrogen Energy 24 (1999) 505.

470

M. Ni et al. / Fuel Processing Technology 87 (2006) 461 472

[3] Z.Q. Mao, Hydrogena future clean energy in china, Symposium on


Hydrogen Infrastructure Technology for Energy and Fuel Applications, The Hong Kong Polytechnic University, Hong Kong, 2003
(November 18).
[4] A. Demirbas, Biomass resource facilities and biomass conversion
processing for fuels and chemicals, Energy Conversion and Management
42 (2001) 1357.
[5] J. Larminie, A. Dicks, Fuel Cell Systems Explained, Wiley, Toronto,
2000.
[6] R.K. Jalan, V.K. Srivastava, Studies on pyrolysis of a single biomass
cylindrical pellet-kinetic and heat transfer effects, Energy Conversion and
Management 40 (1999) 467.
[7] R. Evans, L. Boyd, C. Elam, S. Czernik, R. French, C. Feik, S. Philips, E.
Chaornet, Y. Parent, Hydrogen from biomass-catalytic reforming of
pyrolysis vapors, FY 2003 Progress Report, National Renewable Energy
Laboratory, 2003.
[8] A. Demirbas, Gaseous products from biomass by pyrolysis and gasification: effects of catalyst on hydrogen yield, Energy Conversion and
Management 43 (2002) 897.
[9] A.V. Bridgwater, Principles and practice of biomass fast pyrolysis
processes for liquids, Journal of Analytical and Applied Pyrolysis 51
(1999) 3.
[10] M.A. Rabah, S.M. Eldighidy, Low cost hydrogen production from waste,
International Journal of Hydrogen Energy 14 (1989) 221.
[11] P.A. Simell, N.A.K. Hakala, H.E. Haario, A. Krause, Catalytic decomposition of gasification gas tar with benzene as the model compound,
Industrial and Engineering Chemistry Research 36 (1997) 42.
[12] P.A. Simell, E.K. Hirvensalo, V.T. Smolander, A. Krause, Steam reforming
of gasification gas tar over dolomite with benzene as a model compound,
Industrial and Engineering Chemistry Research 38 (1999) 1250.
[13] I. Narvaez, J. Corella, A. Orio, Fresh tar (from a biomass gasifier)
elimination over a commercial steam-reforming catalyst. Kinetics and
effect of different variables of operation, Industrial and Engineering
Chemistry Research 36 (1997) 317.
[14] P.T. Williams, A.J. Brindle, Catalytic pyrolysis of tyres: influence of
catalyst temperature, Fuel 81 (2002) 2425.
[15] G. Chen, J. Andries, H. Spliethoff, Catalytic pyrolysis of biomass for
hydrogen rich fuel gas production, Energy Conversion and Management
44 (2003) 2289.
[16] D. Sutton, B. Kelleher, J. Ross, Catalytic conditioning of organic volatile
products produced by peat pyrolysis, Biomass and Bioenergy 23 (2002)
209.
[17] L. Garcia, R. French, S. Czernik, S. Chornet, Catalytic steam reforming of
bio-oils for the production of hydrogen: effects of catalyst composition,
Applied Catalysis A: General 201 (2000) 225.
[18] J. Abedi, Y.D. Yeboah, M. Realff, D. McGee, J. Howard, K.B. Bota, An
integrated approach to hydrogen production from agricultural residues for
use in urban transportation, Proceedings of the 2001 DOE Hydrogen
Program Review, NREL/CP-570-30535, National Renewable Energy
Laboratory, 2001.
[19] S. Czernik, R. French, R. Evans, E. Chornet, Hydrogen from postconsumer residues, U.S.DOE Hydrogen and Fuel Cells Merit Review
Meeting, Berkeley, CA, 2003 (May 19 23).
[20] O. Onay, O.M. Kockar, Fixed-bed pyrolysis of rapeseed (Brassica napus
L.), Biomass and Bioenergy 26 (2004) 289.
[21] K.A.M. Bair, S. Czernik, R. French, Y. Parent, M. Ritland, E. Chornet,
Fluidizable catalysts for producing hydrogen by steam reforming biomass
pyrolysis liquids, Proceedings of the 2002 U.S. DOE Hydrogen Program
Review, NREL/CP-610-32405, National Renewable Energy Laboratory,
2002.
[22] K.A.M. Bair, S. Czernik, R. French, E. Chornet, Fluidizable catalysts for
hydrogen production from biomass pyrolysis/steam reforming, FY 2003
Progress Report, National Renewable Energy Laboratory, 2003.
[23] Y. Yeboah, K. Bota, D. Day, D. McGee, M. Realff, R. Evans, E.
Chornet, S. Czernik, C. Feik, R. French, S. Philips, J. Patrick, Hydrogen
from biomass for urban transportation, Hydrogen, Fuel Cells and
Infrastructure Technologies Program Review Meeting, Berkeley, CA,
2003 (May 18 22).

[24] C.E.G.. Padro, V. Putsche, Survey of the economics of hydrogen


technologies, National Renewable Energy Laboratory Technical Report,
NREL/TP_570_27079, September, 1999.
[25] A. Demirbas, Hydrogen production from biomass by the gasification
process, Energy Sources 24 (2002) 59.
[26] T.A. Milne., N. Abatzoglou, R.J. Evans, Biomass gasifier _tars_: their
nature, formation, and conversion, National Renewable Energy Laboratory Technical Report, NREL/TP_570_25357, 1998.
[27] J. Corella, M.P. Aznar, J. Gil, M.A. Caballero, Biomass gasification in
fluidised bed: where to locate the dolomite to improve gasification?
Energy and Fuels 13 (1999) 1122.
[28] D. Sutton, B. Kelleher, J. Ross, Review of literature on catalysts for
biomass gasification, Fuel Processing Technology 73 (2001) 155.
[29] J.M. Wornat, R.H. Hurt, N.Y.C. Yang, T.J. Headley, Structural and
compositional transformations of biomass chars during combustion,
Combustion and Flame 100 (1995) 131.
[30] S. Arvelakis, E.G. Koukios, Physicochemical upgrading of agroresidues
as feedstocks for energy production via thermochemical conversion
methods, Biomass and Bioenergy 22 (2002) 331.
[31] S. Arvelakis, P. Vourliotis, E. Kakaras, E.G. Koukios, Effect of leaching
on the ash behavior of wheat straw and olive residue during fluidized bed
combustion, Biomass and Bioenergy 20 (2001) 459.
[32] S. Arvelakis, H. Gehrmann, M. Beckmann, E.G. Koukios, Effect of
leaching on the ash behavior of olive residue during fluidized bed
gasification, Biomass and Bioenergy 22 (2002) 55.
[33] S.Q. Turn, C.M. Kinoshita, D. Ishimura, Removal of inorganic constituents of biomass feedstocks by mechanical dewatering and leaching,
Biomass and Bioenergy 12 (1997) 241 252.
[34] P. Garcia-Ibanez, A. Cabanillas, J.M. Sanchez, Gasification of leached
orujillo (olive oil waste) in a pilot plant circulating fluidized bed reactor.
Preliminary results, Biomass and Bioenergy 27 (2004) 183.
[35] Y. Yongjie, Exploring energy from biomassthe gasification of residues
from hydrolyzed sawdust, Acta Energiae Solaris Sinica 17 (1996) 209 (in
Chinese).
[36] W. Chuangzhi, Y. Xiuli, X. Bingyan, L. Zhengfan, L. Ping, The
performance study of biomass gasification with oxygen-rich air, Acta
Energiae Solaris Sinica 18 (1997) 237 (in Chinese).
[37] Y. Xia, W. Dunsong, Study of gasification treatment of biomass in fixed
bed gasifier, Mei Qi Yu Re Li 20 (2000) 243 (in Chinese).
[38] S. Turn, C. Kinoshita, Z. Zhang, D. Ishimura, J. Zhou, An experimental
investigation of hydrogen production from biomass gasification, International Journal of Hydrogen Energy 23 (1998) 641.
[39] S. Rapagna, N. Jand, P.U. Foscolo, Catalytic gasification of biomass to
produce hydrogen rich gas, International Journal of Hydrogen Energy 23
(1998) 551.
[40] J. Jian-chun, J. Chun, Z. Jin-Ping, Y. Hao, D. Wei-di, T. Yuan-bo,
Study on industrial applied technology for biomass catalytic gasification, Chemistry and Industry of Forest Products 21 (2001) 21 (in
Chinese).
[41] W. Zhiwei, T. Songtao, S. Xueyong, L. Zian, C. Congming, L. Dingkai, A
study on model for biomass pyrolysis and gasification in fluidized bed,
Journal of Fuel Chemistry and Technology 30 (2002) 342 (in Chinese).
[42] C. Courson, L. Udron, D. Swierczynski, C. Petit, A. Kiennemann,
Hydrogen production from biomass gasification on nickel catalysts tests
for dry reforming of methane, Catalysis Today 76 (2002) 75.
[43] A. Midilli, M. Dogru, G. Akay, C.R. Howarth, Hydrogen production from
sewage sludge via a fixed bed gasifier product gas, International Journal of
Hydrogen Energy 27 (2002) 1035.
[44] S. Rapagna, H. Provendier, C. Petit, A. Kiennemann, P.U. Foscolo,
Development of catalysts suitable for hydrogen or syn-gas production
from biomass gasification, Biomass and Bioenergy 22 (2002) 377.
[45] R.C. Brown, Biomass-derived hydrogen from a thermally ballasted
gasifier, FY 2003 Progress Report, National Renewable Energy Laboratory, 2003.
[46] D.A. Bowen, F. Lau, R. Zabransky, R. Remick, R. Slimane, S. Doong,
Techno-economic analysis of hydrogen production by gasification of
biomass, NREL FY 2003 Progress Report, National Renewable Energy
Laboratory, 2003.

M. Ni et al. / Fuel Processing Technology 87 (2006) 461 472


[47] S.Y. Lin, Y. Suzuki, H. Hatano, M. Harada, Hydrogen production from
hyrocarbon by integration of Water-Carbon Reaction and Carbon Dioxide
Removal (HyPr-RING) Method, Energy and Fuels 15 (2001) 339.
[48] S.Y. Lin, Y. Suzuki, H. Hatano, M. Harada, Development an innovative
method, HyPr-RING, to produce hydrogen from hydrocarbons, Energy
Conversion and Management 43 (2002) 1283.
[49] S.Y. Lin, M. Harada, Y. Suzuki, H. Hatano, Gasification of organic
material/CaO pellets with high-pressure steam, Energy and Fuel 18 (2004)
1014.
[50] S.Y. Lin, M. Harada, Y. Suzuki, H. Hatano, Continuous experiment
regarding hydrogen production by coal/CaO reaction with steam (I) gas
products, Fuel 83 (2004) 869.
[51] S.Y. Lin, M. Harada, Y. Suzuki, H. Hatano, Process analysis for hydrogen
production by reaction integrated novel gasification (HyPr-RING), Energy
Conversion and Management 46 (2005) 869.
[52] S. Sato, S.Y. Lin, Y. Suzuki, H. Hatano, Hydrogen production from heavy
oil in the presence of calcium hydroxide, Fuel 82 (2003) 561 567.
[53] Y. Matsumura, T. Minowa, Fundamental design of a continuous biomass
gasification process using a supercritical water fluidized bed, International
Journal of Hydrogen Energy 29 (2004) 701.
[54] W.S.L. Mok, M. Jerry, J. Antal, Uncatalyzed solvolysis of whole biomass
hemicellulose by hot compressed liquid water, Industrial and Engineering
Chemistry Research 31 (1992) 1157.
[55] T. Minowa, F. Zhen, T. Ogi, Cellulose decomposition in hot-compressed
water with alkali or nickel catalyst, Journal of Supercritical Fluids 13
(1998) 253.
[56] T. Minowa, T. Ogi, Hydrogen production from cellulose using a reduced
nickel catalyst, Catalysis Today 45 (1998) 411.
[57] T. Minowa, S. Inoue, Hydrogen production from biomass by catalytic
gasification in hot compressed water, Renewable Energy 16 (1999) 1114.
[58] D.H. Yu, M. Aihara, M.J. Antal, Hydrogen production by steam reforming
glucose in supercritical water, Energy and Fuels 7 (1993) 574.
[59] X. Xiaodong, M. Yukihiko, S. Jonny, J.A. Michael, Carbon-catalyzed
gasification of organic feedstocks in supercritical water, Industrial and
Engineering Chemistry Research 35 (1996) 2522.
[60] M.J. Antal, X.D. Xu, Hydrogen production from high moisture content
biomass in supercritical water, Proceedings of the 1998 U.S.DOE
Hydrogen Program Review, NREL/CP-570-25315, 1998.
[61] H. Schmieder, J. Abeln, N. Boukis, E. Dinjus, A. Kruse, M. Kluth, G.
Petrich, E. Sadri, M. Schacht, Hydrothermal gasification of biomass and
organic wastes, Journal of Supercritical Fluids 17 (2000) 145.
[62] X.H. Hao, L.J. Guo, X. Mao, X.M. Zhang, X.J. Chen, Hydrogen
production from glucose used as a model compound of biomass gasified
in supercritical water, International Journal of Hydrogen Energy 28
(2003) 55.
[63] M.H. Spritzer, G.T. Hong, Supercritical water partial oxidation, FY 2003
Progress Report, National Renewable Energy Laboratory, 2003.
[64] Y.S. Lin, Microporous and dense inorganic membranes: current status and
prospective, Separation and Purification Technology 25 (2001) 39.
[65] M.W. Reij, J.T.F. Keurentjes, S. Hartmans, Membrane bioreactors for
waste gas treatment, Journal of Biotechnology 59 (1998) 155.
[66] D.B. Levin, L. Pitt, M. Love, Biohydrogen production: prospects and
limitations to practical application, International Journal of Hydrogen
Energy 29 (2004) 173.
[67] P.C. Hallenbeck, J.R. Benemann, Biological hydrogen production;
fundamentals and limiting processes, International Journal of Hydrogen
Energy 27 (2002) 1185.
[68] H. Koku, I. Eroglu, U. Gunduz, M. Yucel, L. Turker, Aspects of the
metabolism of hydrogen production by Rhodobacter sphaeroides,
International Journal of Hydrogen Energy 27 (2002) 1315.
[69] A. Melis, L.P. Zhang, M. Forestier, M.L. Ghirardi, M. Seibert, Sustained
photobiological hydrogen gas production upon reversible inactivation of
oxygen evolution in the green algae Chlamydomonas reinhardtii, Plant
Physiology 122 (2000) 127.
[70] H. Mctavish, L.A. Sayavedrasoto, D.J. Arp, Substitution of Azotobacter
vinelandii hydrogenase small-subunit cysteines by serines can create
insensitivity to inhibition by O2 and preferentially damages H2 oxidation
over H2 evolution, Journal of Bacteriology 177 (1995) 3960.

471

[71] H. Ooshima, S. Takakuwa, T. Katsuda, M. Okuda, T. Shirasawa, M.


Azuma, J. Kato, Production of hydrogen by a hydrogenase-deficient
mutant of Rhodobacter capsulatu, Journal of Fermentation and Bioengineering 85 (1998) 470.
[72] A.A. Tsygankov, L.T. Serebryakova, K.K. Rao, D.O. Hall, Acetylene
reduction and hydrogen photoproduction by wild-type and mutant strains
of Anabaena at different CO2 and O2 concentrations, FEMS Microbiology
Letters 167 (1998) 13.
[73] L. Vasilyeva, M. Miyake, E. Khatipov, T. Wakayama, M. Sekine, M. Hara,
E. Nakada, Y. Asada, J. Miyake, Enhanced hydrogen production by a
mutant of Rhodobacter sphaeroides having an altered light-harvesting
system, Journal of Bioscience and Bioengineering 87 (1999) 619.
[74] T. Kondo, M. Arkawa, T. Hirai, T. Wakayama, M. Hara, J. Miyake,
Enhancement of hydrogen production by a photosynthetic bacterium
mutant with reduced pigment, Journal of Bioscience and Bioengineering 93
(2002) 145.
[75] J.R. Benemann, Process analysis and economics of biophotolysis of water,
IEA/H2/10/TR2-98, 1998.
[76] B. Gaudernack, Photoproduction of hydrogen, IEA Agreement on the
Production and Utilization of Hydrogen Annual Report, IEA/H2/AR-98,
1998.
[77] S.A. Markov, A.D. Thomas, M.J. Bazin, D.O. Hall, Photoproduction of
hydrogen by Cyanobacteria under partial vacuum in batch culture or
in a photobioreactor, International Journal of Hydrogen Energy 22
(1997) 521.
[78] O. Troshina, L. Serebryakova, M. Sheremetieva, P. Lindblad, Production
of H2 by the unicellular Cyanobacterium gloeocapsa alpicola CALU 743
during fermentation, International Journal of Hydrogen Energy 27 (2002)
1283.
[79] R.L. Kerby, P.W. Ludden, G.P. Roberts, Carbon monoxide-dependent
growth of Rhodospirillum rubrum, Journal of Bacteriology 177 (1995)
2241.
[80] B. Soboh, D. Linder, R. Hedderich, Purification and catalytic properties of
a CO-oxidizing: H2-evolving enzyme complex from Carboxydothermus
hydrogenoformans, Europe Journal of Biochemistry 269 (2002) 5712.
[81] S.A. Ensign, P.W. Ludden, Characterization of the CO oxidation/H2
evolution system of Rhodospirillum rubrum. Role of a 22-kDa iron
sulfur protein in mediating electron transfer between carbon monoxide
dehydrogenase and hydrogenase, Journal of Biological Chemistry 266
(1991) 18395.
[82] G.Y. Jung, J.R. Kim, J.Y. Park, S. Park, Hydrogen production by a new
chemoheterotrophic bacterium Citrobacter sp. Y19, International Journal
of Hydrogen Energy 27 (2002) 601.
[83] E. Wolfrum, P.C. Manese, A. Watt, G. Vanzin, J. Huang, S. Smolinski,
Biological water gas shift, DOE Hydrogen, Fuel Cell, and Infrastructure
Technologies Program Review, 2003 (May 19 22).
[84] A.S. Fedorov, A.A. Tsygankov, K.K. Rao, D.O. Hall, Hydrogen
photoproduction by Rhodobacter sphaeroides immobilised on polyurethane foam, Biotechnology Letters 20 (1998) 1007.
[85] A.A. Tsygankov, A.S. Fedorov, T.V. Laurinavichene, I.N. Gogotov, K.K.
Rao, D.O. Hall, Actual and potential rates of hydrogen photoproduction
by continuous culture of the purple non-sulphur bacterium Rhodobacter
capsulatus, Applied Microbiology and Biotechnology 49 (1998) 102.
[86] H.G. Zhu, T. Suzuki, A.A. Tsygankov, Y. Asada, J. Miyake, Hydrogen
production from tofu wastewater by Rhodobacter sphaeroides immobilized in agar gels, International Journal of Hydrogen Energy 24 (1999)
305.
[87] M. Yetis, U. Gunduz, I. Eroglu, M. Yucel, L. Turker, Photoproduction of
hydrogen from sugar refinery wastewater by Rhodobacter sphaeroides
O.U.001, International Journal of Hydrogen Energy 25 (2000) 1035.
[88] C.Y. Lin, C.H. Jo, Hydrogen production from sucrose using an anaerobic
sequencing batch reactor process, Journal of Chemical Technology and
Biotechnology 78 (2003) 678.
[89] F.R. Hawkes, R. Dinsdale, D.L. Hawkes, I. Hussy, Sustainable fermentative hydrogen production: challenges for process optimisation, International Journal of Hydrogen Energy 27 (2002) 1339.
[90] H.H.P. Fang, H. Liu, Effect of pH on hydrogen production from glucose
by a mixed culture, Bioresource Technology 82 (2002) 87.

472

M. Ni et al. / Fuel Processing Technology 87 (2006) 461 472

[91] N. Kumar, D. Das, Enhancement of hydrogen production by Enterobacter


coacae IIT-BT 08, Process Biochemistry 35 (2000) 589.
[92] S.K. Khanal, W.H. Chen, L. Li, S. Sung, Biological hydrogen production:
effects of pH and intermediate products, International Journal of
Hydrogen Energy 29 (2004) 1123.
[93] E.W.J.V. Niel, P.A.M. Claassen, A.J.M. Stams, Substrate and production
inhibition of hydrogen production by the extreme thermophile Caldicel-

lulosiruptor saccharolyticus, Biotechnology and Bioengineering 81


(2003) 255.
[94] Y. Ueno, S. Otsuka, M. Morimoto, Hydrogen production from industrial
wastewater by anaerobic microflora in chemostat culture, Journal of
Fermentation and Bioengineering 82 (1996) 194.

All in-text references underlined in blue are linked to publications on ResearchGate, letting you access and read them immediately.

Vous aimerez peut-être aussi