You are on page 1of 15

international journal of hydrogen energy 34 (2009) 72087222

Available at www.sciencedirect.com

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

Thermodynamic analyses of adsorption-enhanced steam


reforming of glycerol for hydrogen production
Haisheng Chen, Tianfu Zhang, Bilin Dou, Valerie Dupont, Paul Williams, Mojtaba Ghadiri,
Yulong Ding*
School of Process, Environmental and Materials Engineering, University of Leeds, Leeds LS2 9JT, UK

article info

abstract

Article history:

A non-stoichiometric thermodynamic analysis is performed on the adsorption-enhanced

Received 19 May 2009

steam reforming of glycerol for hydrogen production based on the principle of minimising

Received in revised form

the Gibbs free energy. The effects of temperature (6001000 K), pressure (14 bar), water to

25 June 2009

glycerol feed ratio (3:112:1), percentage of CO2 adsorption (099%) and molar ratio of

Accepted 27 June 2009

carrier gas to feed reactants (1:15:1) on the reforming reactions and carbon formation are

Available online 23 July 2009

examined. The results show that the use of a CO2 adsorbent enhances glycerol conversion
to hydrogen and the maximum number of moles of hydrogen produced per mole of glyc-

Keywords:

erol can be increased from 6 to 7 due to the CO2 adsorption. The analyses suggest that the

Hydrogen production

most favourable temperature for steamglycerol reforming is between 800 and 850 K in the

Steam reforming of glycerol

presence of a CO2 adsorbent, which is about 100 K lower than that for reforming without

Adsorption-enhanced

CO2 adsorption. Although high pressures are favourable for CO2 adsorption, a lower

reaction process

operating pressure gives a higher overall hydrogen conversion. The most favourable water

Thermodynamic analyses

to glycerol feed ratio is found to be w9.0 above which the benefit becomes marginal.

Carbon formation

Carbon formation could occur at low water to glycerol feed ratios, and the use of a CO2
adsorbent can suppress the formation reaction and substantially reduce the lower limit of
the water to glycerol feed ratio for carbon formation.
2009 International Association for Hydrogen Energy. Published by Elsevier Ltd. All rights
reserved.

1.

Introduction

Due to environmental concerns, the global demand for


hydrogen is expected to greatly increase in the future.
Currently, over 60% of the worlds feedstock for hydrogen
production is natural gas. Although such a feedstock is
regarded as the most environmentally benign fossil fuel, it
makes a considerable contribution to the global warming. One
of the solutions to this issue is to use biomass based raw
materials to produce hydrogen [1]. This work is concerned
about the use of glycerol, a major by-product of biodiesel
production [2]. Although a small amount of glycerol from

biodiesel production is purified for pharmaceutical and food


applications, the majority is taken as a waste. As a consequence, efficient and effective use of such a material represents a challenge.
Several approaches have been proposed to convert glycerol
to hydrogen over the past few years. These include aqueous
phase reforming [3], bioconversion using genetically engineered enzymes [2], auto-thermal reforming over Rh catalysts
[4], and catalytic steam reforming over Yttria/Zirconia supported Ru catalysts [5]. There is also a theoretical study on
steamglycerol reforming through thermodynamic analyses
[6], which suggests that favourable conditions for steaming

* Corresponding author. Tel.: 44 113 3432747; fax: 44 113 3432405.


E-mail address: y.ding@leeds.ac.uk (Y. Ding).
0360-3199/$ see front matter 2009 International Association for Hydrogen Energy. Published by Elsevier Ltd. All rights reserved.
doi:10.1016/j.ijhydene.2009.06.070

7209

international journal of hydrogen energy 34 (2009) 72087222

reforming of glycerol for hydrogen production are at


900w950 K with a water (steam) to glycerol feed ratio (WGFR
[6]) of 9:1, which is equivalent to a steam to carbon ratio (S/C)
of 3:1. This work seeks to understand whether the operating
parameters, such as temperature, pressure, WGFR and output
hydrogen concentration, for the steamglycerol reforming can
be improved by using the concept of sorption enhanced
reaction. This concept is based on the Le Chateliers principle
and can be illustrated through the following overall reaction:
C3 H8 O3 3H2 O 3CO2 7H2 ;

DH298 128 kJ=mol

(1)

If the reaction is equilibrium-limited, the glycerol conversion and the rate of the forward reaction can be enhanced by
removing CO2 as soon as it is formed from the reaction zone.
Sorption enhanced reaction processes have been the subject
of numerous investigations since the 1980s [7]. Among the
published work, the sorption enhanced steam reforming of
methane is of particular relevance to the current work, which
has been demonstrated both theoretically and experimentally
[810]. It was found in these studies that the operating
temperature can be reduced substantially from around 1000 K
to about 700 K by using an adsorbent. One of the objectives of
this work is, therefore, to find out the level of temperature
reduction for steamglycerol reforming if a CO2 adsorbent is
used. From an economic point of view, the use of a 9:1 water to
glycerol feed ratio (S/C 3:1) is high. It is also known that
reduction in the ratio can lead to the formation of carbon. As
a consequence, another important objective of the current
work is to find out if the use of a sorption enhanced reaction
process can alleviate such a problem. A combination of stoichiometric and non-stoichiometric thermodynamic analyses
will be used to achieve the above objectives, which are based
on the minimisation of Gibbs free energy for chemical reactions. Such analyses will also provide information for optimal
operating conditions to guide further experimental work.
This paper is structured in the following manner. Section 2
gives the details of the fundamentals associated with the
analyses. The results are then discussed in Section 3 and
finally conclusions are summarised in Section 4.

2.

Theoretical considerations

2.1.

Non-stoichiometric approach

(2)

It is known from thermodynamics that the differential form


of the Gibbs free energy can be written as:
dG S dT V dP

N
X

mi dni

N
X

mi dni

(4)

i1

At equilibrium, the Gibbs free energy is at a minimum,


which implies the derivative of G with respect to n is zero. The
question now is to obtain G as a function of ni and to obtain
a set of ni that minimises the value of G. Two methodologies
could be used to address the issue: stoichiometric and nonstoichiometric approaches. For the stoichiometric approach,
the system is described by a set of stoichiometric reactions,
which typically involves selection of an arbitrary set of
possible reactions according to experience [8]. The non-stoichiometric approach involves the direct minimisation of the
Gibbs free energy to calculate the composition of the system
for a given set of species. This work uses the non-stoichiometric method for determining the gas phase composition.
Although such a method requires solving a set of non-linear
equations, it has several advantages: no need for a preset
selection of possible chemical reactions, easy achievement of
convergence in computation, no need for an accurate estimation of initial equilibrium compositions, and easy incorporation of adsorption and carrier gas effects [12]. The nonstoichiometric method is briefly discussed below.
Starting from Equation (4), the total Gibbs energy of the
system can be given as:
G

N
X

mi n i

(5)

i1

To minimise the Gibbs energy, the following constraint is


introduced based on the elemental balances:
N
X

aji ni bj ;

j 1; .; M

(6)

i1

where aji is the number of j-type atoms in the i-species and bj


refers to the number of j-type atoms in the feed. By introducing
the Lagrangian multipliers, lj, a new function G0 is defined:
G0

N
X
i1

mi ni

M
X
j1

lj

N
X

!
(7)

aji ni  bj

i1

A comparison between Equations (6) and (7) shows that the


second term in the right-hand side of Equation (7) is null,
hence G is identical to G0 . To find the composition at which G0
is at its minimum value, its derivative with respective to ni
must be zero. This leads to:

Consider a chemical system. The total Gibbs free energy (G)


depends on temperature (T ), pressure (P) and molar quantities
of components (n) in the system [11]:
G GT; P; n

dG

(3)

 0
M
X
vG
mi
lj aji 0;
vni T;P;nj si
j1

i 1; .; N

In gaseous chemical systems, Equation (8) can be further


expressed as:
f
bi
DGi RT ln yi P RT ln f

M
X

lj aji 0;

i 1; .; N

(9)

lj aji =RT 0; i 1;.;N

(10)

j1

or

i1

where S is the entropy, V is the volume, ni is the number of


moles of component i in the system and mi is the chemical
potential of component i. If the temperature and pressure of
the system are constant, one has:

(8)

bi
DGi =RT lnni =nT ln P ln f

M
X
j1

b is the fugacity coefficient of the gas mixture, which


where f
can be calculated using the RedlichKwong equation of state,

7210

international journal of hydrogen energy 34 (2009) 72087222

nT

N
X

ni ;

i 1; .; N

(11)

i1

Equations (6), (10) and (11) represent a (N M 1) non-linear


algebraic equation system. These equations can be solved for
all the unknowns including yi, li and ni at equilibrium, see
Section 2.5 for details.

2.2.

Species in steamglycerol system

In the non-stoichiometric formulation, the species coexisting in


the system at equilibrium must first be defined. The number of
compounds in the steamglycerol system, resulting from the
atomic combination among C, H and O, is large. There is
therefore a need to determine the most important species to
make the problem manageable. Normally, a selection of the
species is determined by considering the following: (i) the Gibbs
free energy of formation of the compounds, (ii) the proportion of
the atoms in the system, and (iii) previous experience with
similar systems [12]. These are discussed below.
In a homologous series, the Gibbs free energy of formulation increases with the number of carbon atoms. Compounds
with more than three carbon atoms are therefore not likely to
exist in the steamglycerol system [12]. At the initial stage of
this work, calculations were carried out with the typical
methyl (carbon monoxide - CO, carbon dioxide - CO2, methane
- CH4, methanal - CH2O, methanol - CH3OH), ethyl (ethylene C2H4, ethane - C2H6, ethanal - CH3COH, ethanol - C2H5OH) and
propyl (propane - C3H8, propene - C3H6, propanal CH3CH2CHO, propanone - CH3COCH3) compounds, as well as
glycerol (C3H5OH3), steam (H2O) and hydrogen (H2). The mole
fractions of the compounds under a typical set of operating
conditions of T 800K, P 1 bar, water to glycerol feed
ratio 9:1 (S/C 3:1) are listed in Table 1. The reasons for
selecting such a set of conditions are associated with the most
favourable conditions for steamglycerol reforming for
hydrogen production without adsorption, and that, in general,
the adsorption process can give a lower working temperature;
see Section 3 for more discussion. It was found that the
concentrations of methanal, methanol, ethylene, ethane,
ethanal, ethanol, propane, propene, propanal and propanone
are all negligible with mole fractions below 106. As a consequence, only the methyl group, CO, CO2, CH4 plus glycerol,
steam and H2, are considered in the subsequent calculations.
On the other hand, it is noted that the C:H:O ratio of glycerol
is 3:8:3. According to the CHO equilibrium phase diagram [15],
the carbon to oxygen ratio (1:1) is high enough to break glycerol
to CO and H2 via the pyrolysis reaction C3H8O3 4 4H2 3CO.
Furthermore, in the steam reforming of glycerol, the steam feed
could provide further addition of hydrogen and oxygen and
shift the pyrolysis reaction further via the water-gas shift:
CO H2O 4 CO2 H2. This seems to indicate that there is no
need to consider solid carbon formation during steamglycerol
reforming. However, the phase diagram only assumes an ideal

Table 1 Mole fractions of the compounds at typical


conditions.
Species

Molar fraction
1.2  108
1.3  109
1.5  109
1.0  1011
2.0  1011
2.0  1013
9.4  1013
2.0  1013
1.5  1015
6.7  1015

Methanal (CH2O)
Methanol (CH3OH)
Ethylene (C2H4)
Ethane (C2H6)
Ethanal (CH3COH)
Ethanol (C2H5OH)
Propane (C3H8)
Propene (C3H6)
Propanal (CH3CH2CHO)
Propanone (CH3COCH3)

case which is not always satisfied [16] and the existence of


carbon has been found under some conditions by both thermodynamic analyses and experiments in similar systems
[12,17,18]. As a consequence, carbon formation is considered in
this work. As the vapour pressure of carbon is negligible under
the temperature and pressure conditions of this work, only
solid-phase carbon needs to be considered. For doing so, the
following equation could be used as Equations (8)(10) are not
applicable for carbon:
f

nC DGC 0

(12)
GCf is

the Gibbs
where nC is the number of moles of carbon, and
free energy of carbon formation, which is zero. However, the use
of Equation (12) can lead to numerical instabilities due to
a difference in the numbers of unknowns and equations [12,19].
To avoid this numerical issue, carbon formation is investigated
by using a combined method of non-stoichiometric and stoichiometric thermodynamic analyses. Calculations are first
conducted using the non-stoichiometric methodology with
a high water to glycerol feed ratio with which the carbon
formation is unlikely to occur. The resulting species mole fractions are then used to investigate stoichiometrically the

H2, Adhikari et al. 2007


H2, Current work
CH4, Adhikari et al. 2007
CH4, Current work

CO, Adhikari et al. 2007


CO, Current work
CO2, Adhikari et al. 2007
CO2, Current work

7
6

Molar Number

yi is the mole fraction of component i ,DGfi is the standard


Gibbs free energy of formation, which is only a function of
temperature and is available in handbooks of thermodynamic
data [11,13,14], R is the universal gas constant, and nT is the
total number of moles of all species defined by:

5
4
3
2
1
0
600

700

800

900

1000

Temperature (K)
Fig. 1 Equilibrium compositions: comparison between
this work and Ref [6] for T [ 6001000 K, P [ 1 bar and
WGFR [ 9:1.

7211

international journal of hydrogen energy 34 (2009) 72087222

8
H2

H2
6

CH4

CO2

Molar Number

Molar Number

CH4
CO

CO

4
3
2

CO2 before adsorption


CO2 after adsorption

4
3
2

0
1

0
0.0

0.2

0.4

0.6

0.8

1.0

CO2 removal fraction

Pressure (bar)
Fig. 2 Equilibrium compositions: effect of pressure for
T [ 800 K, P [ 14 bar, and WGFR [ 9:1.

Fig. 4 Equilibrium compositions: effect of CO2 removal for


WGFR [ 9:1, P [ 1 bar, T [ 800 K.

possibility of carbon formation. For doing so, the following four


possible reactions of carbon formation are considered [17]:
2CO CO2 C

(13)

CH4 2H2 C

(14)

CO H2 H2 O C

(15)

CO2 2H2 2H2 O C

(16)

The activities of the four reactions can be expressed as:


A13 K13 Py2CO =yCO2

(17)

A14 K14 P1 yCH4 =y2H2

(18)

A15 K15 PyCO yH2 =yH2 O

(19)

A16 K16 PyCO2 y2H2 =y2H2 O

(20)

7
H2
6

CH4
CO

Molar Number

CO2

4
3
2
1
0

12

Water to Glycerol Feed Ratio


Fig. 3 Equilibrium compositions: effect of water to glycerol
feed ratio for T [ 800 K, P [ 1 bar and WGFR [ 3:112:1.

Fig. 5 Hydrogen production: effect of temperature for


P [ 1 bar and WGFR [ 9:1, (a) Number of moles, (b) Relative
increment.

7212

international journal of hydrogen energy 34 (2009) 72087222

1.8

1.5
f=0.0

f=0.0
1.5

f=0.2

f=0.2

1.2

f=0.4

f=0.6

1.2

Molar Number

Molar Number

f=0.4
f=0.8
f=0.99

0.9

f=1.0
0.6

f=0.6
f=0.8

0.9

f=0.99
f=1.0
0.6

0.3
0.3
0.0

0.0
600

700

800

900

600

1000

700

1.0
f=0.2

900

1000

f=0.6
f=0.8

0.6

900

1000

1.0

0.8

f=0.4

Decrement

Decrement

0.8

f=0.99
f=1.0
0.4

0.6

0.4

0.2

0.2

0.0
600

700

800

900

1000

0.0
600

700

Fig. 6 Methane production: effect of temperature for


P [ 1 bar and WGFR [ 9:1, (a) Number of moles, (b) Relative
decrement.

where Ki (i 1316) are the equilibrium constants for Reactions (13)(16) respectively and yi is the mole fraction of
component i (i CO, CO2, CH4, H2O). For steamglycerol
reforming, carbon formation is possible when the maximum
of the four activities, AC, is greater than unity:
AC maxA13 ; A14 ; A15 ; A16 > 1

(21)

If the activity of carbon formation is smaller than unity, the


water to glycerol feed ratio will be decreased further until AC
equals 1. As carbon is not desirable, this work will mainly deal
with the possibility of carbon formation as well as the ranges
of conditions under which solid carbon starts to appear (zero
carbon conditions); see Section 3.6 for more discussion.

CO2 adsorption

From Equation (10), one has the following for CO2:




f
b i lC 2lO =RT 0
DGCO2 =RT ln nCO2 =nT ln P ln f

800

Temperature (K)

Temperature (K)

2.3.

800

Temperature (K)

Temperature (K)

(22)

When there is an adsorbent for CO2, Equations (22) can be


modified to give:

f=0.2

f=0.6

f=0.99

f=0.4

f=0.8

f=1.0

Fig. 7 Carbon monoxide production: effect of temperature


for P [ 1 bar and WGFR [ 9:1, (a) Number of moles, (b)
Relative decrement.



f
b i lC 2lO =RT0
DGCO2 =RTln nCO2 $1f =nT lnPln f

(23)

where f is the fraction of CO2 produced by the steamglycerol


reforming that is removed by the adsorption. Accordingly, the
total number of moles of all species in the gas phase is reduced
by nCO2 $f and is given by:
nT

N
X

ni  nCO2 $f ;

i 1; .; N

(24)

i1

As a theoretical study, CO2 adsorbents and sorption equilibria


and kinetics will not be considered here. Interested readers can
refer to [710,2028] for information on adsorbents and their
characteristics including hydrotalicate [710,2022], metal
oxides [2325], double salt [26], and lithium base oxides [27,28].

2.4.

The use of carrier gases

The majority of laboratory studies on similar reaction systems


involved the use of inert carrier gases; see for example Refs

7213

international journal of hydrogen energy 34 (2009) 72087222

3.0

f=0.2
f=0.4
f=0.6
f=0.8
f=0.99
f=1.0

1.0

2.5

0.8

2.0

Increment

Molar Number

1.2

f=0.0
f=0.2
f=0.4
f=0.6
f=0.8
f=0.99
f=1.0

1.5
1.0
0.5
0.0
600

700

800

900

0.6
0.4
0.2
0.0

1000

600

700

Temperature (K)

800

900

1000

900

1000

Temperature (K)

3.0

1.0

2.5

Decrement

Molar Number

0.8
2.0
1.5
1.0
0.5

0.6

0.4

0.2

0.0
600

700

800

900

1000

0.0
600

700

Temperature (K)
f=0.0
f=0.2

f=0.4
f=0.6

f=0.8
f=0.99

800

Temperature (K)
f=1.0

f=0.2
f=0.4

f=0.6
f=0.8

f=0.99
f=1.0

Fig. 8 Carbon dioxide production and remaining CO2: effect of temperature for P [ 1 bar and WGFR [ 9:1, (a) Number of
moles of CO2 production, (b) Relative increment in CO2 production, (c) Number of moles of remaining CO2, (d) Relative
decrement in the remaining CO2.

[7,10,22,25,29,30]. It would be interesting to look at this effect,


which may offer guidance for the laboratory scale experiments. In this work, nitrogen is used as an example carrier gas
(analyses on the use of other inert carrier gases are similar).
For doing so, two non-linear algebraic equations are needed,
one for the atomic balance and the other for the Gibbs free
energy:
2nN2 21 WGFR$nf

(25)



f
b i 2lN =RT 0
DGN2 =RT ln nN2 =nT ln P ln f

(26)

where the nf refers to the molar ratio of carrier gas (nitrogen)


to the feed reactants (H2O glycerol). Also, the number of
moles of the carrier nitrogen, nN2 , should be added to the total
number of moles of all species, nT, in the calculations

2.5.

Numerical calculations

In total, four atoms, C, H, O, N, and seven compounds, C3H8O3,


H2O, H2, CO, CO2, CH4, N2, are considered in the equilibrium

system, and Equations (6), (10), (22)(26) take the following


explicit forms:
nCH4 nCO nCO2 3nC3 H8 O3 3
4nCH4 2nH2 2nH2 O 8nC3 H8 O3 2WGFR 8
nH2 O nCO 2nCO2 3nC3 H8 O3 WGFR 3
2nN2 21 WGFR$nf
f
b CH4 lC 4lH =RT 0
DGCH4 =RT lnnCH4 =nT ln P ln f


f
b H2 O 2lH lO =RT 0
DGH2 O =RT ln nH2 O =nT ln P ln f


f
b H2 2lH =RT 0
DGH2 =RT ln nH2 =nT ln P ln f
f
b CO lC lO =RT 0
DGCO =RT lnnCO =nT ln P ln f


f
b CO lC 2lO =RT 0
DGCO2 =RT ln nCO2 $1  f =nT ln P ln f
2


f
bC H O
DGC3 H8 O3 =RT ln nC3 H8 O2 =nT ln P ln f
3 8 3
3lC 8lH 3lO =RT 0


f
b N 2lN =RT 0
DGN2 =RT ln nN2 =nT ln P ln f
2
nT nCH4 nH2 nH2 O nCO nCO2 $1  f nC3 H8 O3 nN2
(27)
The task now is to solve the above 12 non-linear equations,
which contains 12 unknowns for the 4 atoms, 7 compounds
and the total number of moles (nT). For doing so, an in-house
Fortran Code based on the subroutine DNEQNF of the IMSL

7214

international journal of hydrogen energy 34 (2009) 72087222

1.5
f=0.4
f=0.6

f=0.0
f=0.2

f=1.0

f=0.8
f=0.99

1.2

Molar Number

Molar Number

6
5
4
3
2

0.6

0.3

1
0

0.9

f=0.0

f=0.4

f=0.8

f=0.2

f=0.6

f=0.99

f=1.0

0.0
4

Pressure (bar)

Pressure (bar)

2.0

f=0.2

1.0

f=0.4
f=0.6

1.5

0.8

f=0.99

Decrement

Increment

f=0.8
f=1.0
1.0

0.6
f=0.2
f=0.4

f=0.6
f=0.8

f=0.99
f=1.0

0.4

0.2

0.5

0.0
0.0

Pressure (bar)

Pressure (bar)

Fig. 9 Hydrogen production: effect of pressure for


T [ 800 K and WGFR [ 9:1, (a) Number of moles, (b)
Relative increment.

Fig. 10 Methane production: effect of pressure for


T [ 800 K and WGFR [ 9:1, (a) Number of moles, (b)
Relative decrement.

library was written. DNEQNF solves non-linear equations


using a modified Powell hybrid algorithm and a finite difference approximation to the Jacobian. The program requires the
following as the input at the beginning: (i) number of atoms,
(ii) number of compounds, (iii) temperature, (iv) pressure, (v)
values of Gibbs free energy, and (vi) guessed initial values of
the unknowns. All the thermodynamic data used in the
calculations were obtained from Refs [11,13,14]. The initial
quantity of glycerol was taken as 1 mole, the molar quantities
of H2O and N2 depended on the water to glycerol feed ratio
(WGFR in the equations) and nitrogen feed (nf), and the fraction of adsorbed CO2 was also used as an input parameter.

equilibrium compositions of H2, CO2, CH4 and CO for WGFR 9:1


(S/C 3:1), P 1 bar and T 6001000 K with the data from Ref
[6]. One can see that excellent agreement has been obtained,
indicating the appropriateness of the methodology used in this
work. Under the conditions of this work, the conversion of
glycerol was found always to be greater than 99.99%, and hence
the conversion of glycerol is close to completion.
Fig. 1 shows that the number of moles of hydrogen at
equilibrium increases rapidly with increasing temperature
from 600 K to about 850 K, beyond which the increasing trend
slows down substantially and the number of moles of
hydrogen peaks at around 950 K. A further increase in the
temperature leads to a decrease in the number of moles of
hydrogen. The upper limit of hydrogen production under this
set of conditions is w6.0 moles per unit mole of glycerol rather
than the theoretical number of moles of hydrogen (7.0) contained in a mole of glycerol (Equation (1)). As will be shown in
Section 3.2, it is possible to achieve the theoretical number of
moles of hydrogen under similar conditions using a CO2
adsorbent. The competitive product, CH4, however, decreases
with increasing temperature and the formation of CH4 is
almost inhibited when the temperature is above 950 K. Similar

3.

Results and discussion

3.1.
Steamglycerol reforming in the absence of
adsorption
Calculations were first conducted on the steam reforming of
glycerol without adsorption for which thermodynamic data are
available for comparison [6,16]. Fig. 1 compares the calculated

international journal of hydrogen energy 34 (2009) 72087222

Molar Number

3.2.
Steamglycerol reforming with CO2 adsorption for
typical sets of conditions

0.5
f=0.4
f=0.6

f=0.0
f=0.2

0.4

f=1.0

f=0.8
f=0.99

0.3

0.2

0.1

0.0
1

Pressure (bar)

1.0
f=0.2
f=0.4

Decrement

0.8

f=0.6
f=0.8

7215

f=0.99
f=1.0

0.6

0.4

Fig. 4 shows the effect of CO2 removal on the equilibrium


compositions for WGFR 9:1 (S/C 3:1), P 1 bar and
T 800 K. One can see that, an increase in fractional CO2
adsorption enhances hydrogen and CO2 production, but
reduces CH4 and CO production. The enhancement of
hydrogen production is particularly significant when over
80% of CO2 is removed. In the ideal scenario of 100% CO2
removal, 7.0 moles of hydrogen are obtained per unit mole of
glycerol with 3.0 moles of CO2 (fully adsorbed), and no CH4 or
CO. Accordingly, the maximum hydrogen concentration
increases from 66.4% [6] to 100% due to CO2 adsorption. This
agrees with the theoretical stoichiometry of glycerol, indicating again the appropriateness of the methodology used in
this work.
Having established confidence in the methodology,
further and more comprehensive analyses on the adsorptionenhanced hydrogen production have been performed over
T 6001000 K, P 14 bar; WGFR 3:112:1 (S/C 1:14:1),
f 0.20.99, and nf 1:15:1. For some cases, results with
100% CO2 removal ( f 1.0) are also discussed as a reference
although f w0.99 is more practicable. The results of the
analyses on the effects of temperature, pressure and water to
glycerol feed ratio are discussed in the following (Sections
3.33.5).

0.2

0.0

3.3.
Effect of temperature on adsorption-enhanced
hydrogen production
1

Pressure (bar)
Fig. 11 Carbon monoxide production: effect of pressure
for T [ 800 K and WGFR [ 9:1, (a) Number of moles, (b)
Relative decrement.

to hydrogen, the number of moles of CO2 increases with


increasing temperature first and peaks at w850 K beyond
which there is a decreasing trend. On the other hand, CO is
seen only to increase with temperature. The reason for the
above could be explained by the reforming reaction for
hydrogen production (Equation (1)) and CO production via the
reverse water-gas shift reaction (CO2 H2 4 CO H2O), both
of which are favoured by high temperatures.
Fig. 2 shows the effect of pressure on the molar quantities
of equilibrium compositions at T 800 K and WGFR 9:1 (S/
C 3:1). With increasing pressure, the molar quantities of H2,
CO and CO2 decrease, whereas the number of moles of the
competitive product, CH4, increases. These are as expected
from Equation (1). The effect of the water to glycerol feed ratio
is shown in Fig. 3. Again, as expected, an excess amount of
steam is beneficial for hydrogen production although the
effect seems to be small for ratios of above 9.
A comparison of the results shown in Figs. 13 suggests
that the most favourable conditions for steamglycerol
reforming for hydrogen production in the absence of adsorption is at a temperature of w900 K, a pressure of w1 bar and
a WGFR of 9:1 (S/C 3:1).

Fig. 5(a) and (b) illustrates the effect of temperature on


adsorption-enhanced hydrogen production for P 1 bar and
WGFR 9:1 (S/C 3:1). From Fig. 5(a), hydrogen production is
seen to be enhanced over the entire range of temperatures
considered. At a given fraction of CO2 adsorption (except for
100% CO2 removal for which the temperature has little
effect), the number of moles of hydrogen at equilibrium
increases approximately linearly with increasing temperature until a critical temperature is reached above which the
slope of the increase decreases rapidly and a peak occurs
followed by a decreasing trend with a further increase in
temperature. The critical temperature decreases with
increasing fraction of CO2 removal via adsorption, whereas
the peak temperature is around 950 K and does not appear to
depend on the fractional removal of CO2. With 99% CO2
removal ( f 0.99), an operating temperature over 800 K leads
to hydrogen production close to the ideal scenario of 7 moles
per mole of glycerol. If the relative change (increment) of
hydrogen production (RCH2 nH2 $ad  nH2 =nH2 ) is used to
quantify adsorption enhanced steamglycerol reforming,
Fig. 5(b) is obtained, where the subscript ad refers to
adsorption. It can be seen that the relative increment of
hydrogen production decreases with increasing temperature
and the decrease tends to level off at above 850900 K. There
seems to be a minimum in the relative increment at w900 K.
As low temperatures are generally more favourable for high
capacities of CO2 adsorbents, Fig. 5(a) and 5(b) suggest that
the most favourable temperature for steamglycerol reforming for hydrogen production is w800850 K. At temperatures

7216

international journal of hydrogen energy 34 (2009) 72087222

3.0

1.0
f=0.2
f=0.4

f=0.99
f=1.0

f=0.6
f=0.8

0.8

Increment

Molar Number

2.7

2.4

2.1

0.4

0.2

1.8
f=0.4
f=0.6

f=0.0
f=0.2
1.5

0.6

f=0.8
f=0.99

f=1.0

0.0

Pressure (bar)

3.0
f=0.4
f=0.6

f=0.0
f=0.2

2.5

f=0.8
f=0.99

f=1.0

1.0
f=0.2
f=0.4

0.8

f=0.99
f=1.0

f=0.6
f=0.8

2.0

Decrement

Molar Number

Pressure (bar)

1.5
1.0
0.5

0.6

0.4

0.2

0.0

0.0
1

Pressure (bar)

Pressure (bar)

Fig. 12 Carbon dioxide production: effect of pressure for T [ 800 K and WGFR [ 9:1, (a) Number of moles, (b) Increment, (c)
Number of moles of remaining CO2, (d) remaining CO2 decrement.

below 800 K, the absolute number of moles of hydrogen


production is low (Fig. 5a) even with 99% CO2 removal by
adsorption ( f 0.99). At temperatures above w850 K,
however, the relative increment of hydrogen production due
to CO2 adsorption is small (Fig. 5b). Additionally, high
temperature operations would also have issues such as
capital costs, energy integration and adsorbent capacity.
Fig. 6 shows the effect of temperature on competitive
methane production at P 1 bar and WGFR 9:1 (S/C 3:1).
Contrary to hydrogen production, the number of moles of
methane at equilibrium decreases with increasing fraction of
CO2 adsorption over the entire range of temperatures
considered in this work. For the case of full removal of CO2,
virtually no methane is produced. At a given fraction of CO2
adsorption (except for 100% CO2 removal), the number of
moles of methane at equilibrium decreases rapidly with
increasing temperature and becomes negligible at a temperature depending on the percentage of CO2 removal. For
example, when 99% CO2 is removed ( f 0.99), almost no
methane production occurs above w800 K.
The relative change (decrement) of methane production
defined as RCCH4 nCH4 >ad  nCH4 =nCH4 is shown in Fig. 6(b) as
a function of temperature for various extents of CO2 removal.
It is interesting to note that with a CO2 removal fraction below

f 1, a maximum decrement occurs and the temperature


corresponding to the maximum is a function of the fraction of
CO2 removal, ranging from w800 K for 20% CO2 removal to
w950 K for 99% CO2 removal. This suggests that low temperatures are more favourable for steamglycerol reforming with
low CO2 removal fractions and vice-versa. The results shown
in Fig. 6(a) and 6(b) suggest that the most favourable temperature for suppressing methane production is w800850 K; this
is consistent with the conclusions from Fig. 5(a) and 5(b) for
hydrogen production.
The effect of temperature on carbon monoxide production is
shown in Fig. 7(a) and 7(b) for P 1 bar and WGFR 9:1 (S/
C 3:1). As can be seen from Fig. 7(a), the number of moles of CO
decreases significantly due to CO2 adsorption at all temperatures investigated. However, the relative change (decrement) of
CO mole fraction, defined as RCCO nCO$ad  nCO =nCO and
shown in Fig. 7(b), is a weak function of temperature for all CO2
removal fractions.
Fig. 8 shows the effect of temperature on the amount of
carbon dioxide produced and that remaining after CO2
adsorption for P 1 bar and WGFR 9:1 (S/C 3:1). The
amount of CO2 produced is seen to increase with increasing
fractional CO2 removal at all temperatures, as shown in
Fig. 8(a). For a given CO2 removal fraction, the number of

7217

international journal of hydrogen energy 34 (2009) 72087222

1.2
f=0.0
f=0.2

f=0.8
f=0.99

f=1.0

0.9

Molar Number

Molar Number

f=0.4
f=0.6

5
4
3

0.6

0.3

f=0.4
f=0.6

f=0.0
f=0.2

f=0.8
f=0.99

f=1.0
0.0

12

Water to Glycerol Feed Ratio

2.0

b
f=0.2
f=0.4

f=0.6
f=0.8

f=0.99
f=1.0

1.5

12

1.0

0.8

Decrement

Increment

Water to Glycerol Feed Ratio

1.0

0.6

f=0.2
f=0.4

f=0.6
f=0.8

f=0.99
f=1.0

0.4

0.5
0.2

0.0

0.0
3

12

Water to Glycerol Feed Ratio

12

Water to Glycerol Feed Ratio

Fig. 13 Hydrogen production: effect of water to glycerol


feed ratio for T [ 800 K and P [ 1 bar, (a) Number of moles,
(b) Increment.

Fig. 14 Methane production: effect of water to glycerol


feed ratio for T [ 800 K and P [ 1 bar, (a) Number of moles,
(b) Decrement.

moles of CO2 at equilibrium increases first with increasing


temperature, then peaks at w850 K, a decrease then occurs
with a further increase in the temperature. For a 99% CO2
removal ( f 0.99), the amount of CO2 produced above 800 K
is very close to the theoretical value of 3.0 moles. As a
consequence, high temperatures are mostly favoured for
steamglycerol reforming in terms of CO2 production. This is
reflected more clearly in Fig. 8(b), where the relative change
(increment) of CO2 production due to adsorption
(RCCO2 nCO2 $ad  nCO2 =nCO2 ) is plotted against temperature.
For fractions of CO2 removal below about 0.99, the relative CO2
increment increases with increasing temperature. The results
shown in Fig. 8(a) and 8(b) suggest that temperatures above
800 K are favourable for steamglycerol reforming in terms of
CO2 production.
Due to adsorption, carbon dioxide produced can exist in both
the gas and the adsorbed phases. CO2 in the gaseous phase is
called the remaining CO2, which is shown in Fig. 8(c) and 8(d). As
expected, the remaining CO2 decreases with increasing fraction
of CO2 removal. For a fraction of CO2 adsorption at 99% ( f 0.99),
the remaining CO2 in the gas phase is very small at all

temperatures investigated and the temperature effect is


marginal. The relative decrement of CO2 remaining in the gas
phase, defined as RCCO2 $remain nCO2  nCO2 $ad $1  f =nCO2 , is
shown in Fig. 8(d), which decreases slightly with increasing
temperature for f 0.20.8. For f 0.991.0, the change of
RCCO2 $remain with respect to temperature is negligible.

3.4.
Effect of pressure on adsorption-enhanced hydrogen
production
The effect of pressure on hydrogen production is illustrated in
Fig. 9(a) and 9(b) for T 800 K and WGFR 9:1 (S/C 3:1). For
a given pressure, both the number of moles of hydrogen at
equilibrium and relative change (increment) in the amount of
hydrogen production increase with increasing fractional CO2
removal (Fig. 9(a)). For a given fractional CO2 removal, the
amount of hydrogen produced decreases with increasing
pressure, whereas the relative change (increment) in the
amount of hydrogen production due to CO2 adsorption (RCH2 )
shows an opposite trend (Fig. 9(b)). A comparison of the results

7218

international journal of hydrogen energy 34 (2009) 72087222

f=0.4
f=0.6

f=0.0
f=0.2

0.4

Molar Number

independent of the operating pressure. At 99% CO2 removal


( f 0.99), little CO is produced under all pressures considered
in this work.
The effect of pressure on CO2 production is presented in
Fig. 12. Both the equilibrium amount of CO2 (Fig. 12(a)) and the
remaining CO2 (Fig. 12(c)) decrease with increasing pressure
for a given fraction of CO2 removal; the relative change
(increment) in the amount of CO2 (RCCO2 , Fig. 12(b)) and that of
the remaining CO2 (Fig. 12(d)) are almost independent of
pressure for a given fraction of CO2 adsorption.

0.5
f=1.0

f=0.8
f=0.99

0.3

0.2

0.1

3.5.
Effect of water to glycerol feed ratio (WGFR) on
adsorption-enhanced hydrogen production

0.0
3

12

Water to Glycerol Feed Ratio

b
1.0
f=0.2
f=0.4

Decrement

0.8

f=0.6
f=0.8

f=0.99
f=1.0

0.6

0.4

0.2

0.0

12

Water to Glycerol Feed Ratio


Fig. 15 Carbon monoxide production: effect of water to
glycerol feed ratio for T [ 800 K and P [ 1 bar, (a) Number
of moles, (b) Decrement.

shown in Fig. 9(a) and 9(b) suggests that low pressure operations are favourable for hydrogen production, in agreement
with the results in the absence of CO2 adsorbent.
Fig. 10 shows the effect of pressure on methane production
for T 800 K, and WGFR 9:1 (S/C 3:1). For a given operating
pressure, the equilibrium amount of methane decreases with
increasing fractional CO2 adsorption (Fig. 10(a)), whereas the
relative change (decrement) in the equilibrium amount shows
an opposite trend (Fig. 10(b)). With increasing pressure, the
equilibrium amount increases (Fig. 10(a)) but the relative
change (decrement) in the equilibrium amount decreases
(Fig. 10(b)), if the fractional CO2 removal is fixed. These results
suggest that low pressures are more favourable for hydrogen
production through steamglycerol reforming; this is in
agreement with the results shown in Fig. 9.
The effect of pressure on carbon monoxide production is
illustrated in Fig. 11 for T 800 K and WGFR 9:1 (S/C 3:1).
The number of moles of carbon monoxide is seen to decrease
with increasing fraction of CO2 adsorption for a given operating pressure, and decrease with increasing pressure if the
fraction of CO2 adsorption is fixed. However, the relative
change (decrement) in CO production (RCCO) is almost

The effect of the water to glycerol feed ratio on hydrogen


production is presented in Fig. 13(a) and 13(b) for P 1 bar and
T 800 K. One can see from Fig. 13(a) that the equilibrium
amount of hydrogen increases with an increase in the water to
glycerol feed ratio or an increase in the fractional CO2
removal. At f 0.99, the number of moles of hydrogen is close
to 7.0 at a water to glycerol feed ratio over w9 (S/C w3).
However the relative change (increment) in the number of
moles of hydrogen at equilibrium due to CO2 adsorption (RCH2 )
decreases with increasing water to glycerol feed ratio. These
results suggest that a water to glycerol feed ratio of w9 (S/
C w3) is most favourable for hydrogen production. A water
to glycerol feed ratio higher than 9 (S/C 3) does give a higher
number of moles of hydrogen, but it also gives a low relative
increment in hydrogen and requires additional energy
consumption for upstream heating and downstream separation of hydrogen from water.
The effect of water to glycerol feed ratio in the feed on the
number of moles of methane at equilibrium is shown in Fig. 14
for T 800 K and P 1 bar. It is shown that the number of
moles of methane at equilibrium decreases with an increase
in either the fractional CO2 removal or the water to glycerol
feed ratio, whereas the relative change (decrement) of
methane production due to adsorption increases with
increasing water to glycerol feed ratio. The relative increment
of methane production due to adsorption levels off when the
water to glycerol feed ratio exceeds w9 (S/C w3). These
results suggest a water to glycerol feed ratio of w9 (S/C w3)
is most favourable for preventing methane production and
a further increase in the water to glycerol feed ratio only gives
a small benefit.
Fig. 15 shows the number of moles of carbon monoxide at
equilibrium as a function of water to glycerol feed ratio for
T 800 K and P 1 bar. For a given fractional CO2 removal, an
increase in the water to glycerol feed ratio is seen to
substantially reduce the equilibrium amount of CO. However,
the relative change (decrement) in equilibrium CO production
is only a weak function of the water to glycerol feed ratio.
Fig. 16 shows the effect of the water to glycerol feed
ratio on the number of moles of carbon dioxide produced
and that remains in the system after adsorption for
T 800 K and P 1 bar. From Fig. 16(a) and 16(b), one can
see that the number of moles of CO2 produced increases
with increasing fraction of CO2 adsorption or increasing
water to glycerol feed ratio, whereas the relative increment
in CO2 production due to adsorption decreases with

7219

international journal of hydrogen energy 34 (2009) 72087222

3.0

1.0
f=0.2
f=0.4

f=0.99
f=1.0

f=0.6
f=0.8

0.8

Increment

Molar Number

2.7

2.4

2.1

1.8

1.5

f=0.0
f=0.2
f=0.4
3

f=0.6
f=0.8
f=0.99

0.4

0.2

f=1.0

0.6

0.0

12

Water to Glycerol Feed Ratio

3.0
f=0.4
f=0.6

f=0.0
f=0.2

2.5

f=0.8
f=0.99

12

f=1.0

1.0
f=0.2
f=0.4

0.8

f=0.99
f=1.0

f=0.6
f=0.8

2.0

Decrement

Molar Number

Water to Glycerol Feed Ratio

1.5
1.0
0.5

0.6

0.4

0.2

0.0
3

12

Water to Glycerol Feed Ratio

0.0

12

Water to Glycerol Feed Ratio

Fig. 16 Carbon dioxide production: effect of water to glycerol feed ratio for T [ 800 K and P [ 1 bar, (a) Number of moles, (b)
Increment, (c) Number of moles of remaining CO2, (d) Decrement in the remaining CO2.

increasing water to glycerol feed ratio. From Fig. 16(c) and


16(d), the amount of carbon dioxide that remains in the
system after adsorption is seen to decrease with increasing
fraction of CO2 adsorption and the relative decrement of
CO2 left in the system is only a weak function of the water
to glycerol feed ratio.

3.6.

Carbon formation

The results for carbon formation are shown in Fig. 17(a) and
17(b) for various fractions of CO2 removal through adsorption
at two pressures of 1 bar and 4 bar, respectively. The figures
give the zero carbon formation curves, which correspond to
AC being equal to unity. The region on the left side of the zero
carbon curves is where carbon formation may occur and that
on the right side is the carbon-free region.
Consider first Fig. 17(a) for 1-bar operation, where the
results can be split into three groups according to the fractional CO2 removal ( f ):
 Group I ( f  w0.4) - For this group, the minimum water to
glycerol feed ratio (above which no carbon in theory would
form) decreases with increasing temperature given a CO2
removal fraction. For example, the minimum water to

glycerol feed ratio at f 0 is about 3.8 (S/C 1.26) at 600 K but


the ratio decreases to about 1 (S/C 0.33) at 1000 K. For
a given temperature that is below w900 K, the minimum
water to glycerol feed ratio decreases with increasing fraction of CO2 removal. For temperatures about 900 K, the
effect of fractional CO2 removal on the minimum ratio
becomes negligible.
 Group II (0.4 < f  w0.8) For this group, the minimum water
to glycerol feed ratio, for a given fractional CO2 removal,
increases slightly with increasing temperature up to
w900 K, where a decrease follows. For a given temperature,
the behaviour is similar to the results for Group I, i.e. the
minimum water to glycerol feed ratio decreases with
increasing fractional CO2 removal for temperatures below
900 K, and the effect of the fractional CO2 removal on the
minimum ratio is small but not negligible at temperatures
above 900 K, where a higher fractional CO2 removal is seen
to increase slightly the minimum steamglycerol ratio.
 Group III ( f > w0.8) For this group, the minimum water to
glycerol feed ratio increases with increasing temperature for
a given fractional CO2 removal.
The results at 4 bar are shown in Fig. 17(b), and are
very similar to those at 1 bar. For comparison, the zero

7220

international journal of hydrogen energy 34 (2009) 72087222

a 1000

Carbon Region

P=4bar, f=0.0

Molar Number

Temperature (K)

900

6
No Carbon Region

800
f=0.0
f=0.2
f=0.4
f=0.6
f=0.8
f=0.99

700

600

H2
CO

CO2

3
2
1
0

CH4

800
f=0.0
f=0.2
f=0.4
f=0.6
f=0.8
f=0.99

700

0.8

1.0

5
4
3
2
1

0
0.0

0.2

Water to Glycerol Feed Ratio


Fig. 17 Zero carbon curves of steamglycerol reforming for
hydrogen production, (a) Zero carbon curves at P [ 1 bar,
(b) Zero carbon curves at P [ 4 bar.

carbon curves for 1 bar are included in Fig. 17(b) and those
for 4 bar are also given in Fig. 17(a). One can see that the
difference between the zero carbon curves at the two
pressures is small, indicating a weak pressure effect on
carbon formation. The results shown in Fig. 17(a) and 17(b)
clearly demonstrate that the use of CO2 adsorption
suppresses carbon formation at temperatures below about
900 K.

3.7.

P=1bar, f=0.0

600

No Carbon Region

Molar Number

900

Temperature (K)

Carbon Region

b 1000

Carrier gas to feeding gas ratio

Water to Glycerol Feed Ratio

Effect of carrier gas

The effect of carrier gas on steamglycerol reforming is


illustrated in Fig. 18(a) and 18(b), where the results shown in
Fig. 18(a) are for T 800K, P 1 bar and WGFR 9:1 (S/C 3:1),
and f 0.0 (no CO2 adsorption), and the results shown in
Fig. 18(b) are for nf 3.0 with various CO2 removal fractions
(T 800 K, P 1 bar and WGFR 9:1). The corresponding
results without carrier gas (nf 0.0) are also shown in
Fig. 18(b) for reference. From Fig. 18(a), one can see that the
use of a carrier gas increases the equilibrium molar quantities of hydrogen and CO2 and reduces the production of

0.4

0.6

CO2 removal fraction


H2, nf=3

CO2, nf=3

CO, nf=0

CH4, nf=3

H2, nf=0

CO2, nf=0

CO, nf=3

CH4 , nf=0

Fig. 18 Equilibrium compositions: effects of carrier gas for


T [ 800 K, P [ 1 bar and WGFR [ 9:1, (a) Without CO2
adsorption, (b) With CO2 adsorption.

methane due to the so-called dilution effect. Such a dilution


effect can be explained by using Equation (1). According to
the Le Chateliers principle, if a chemical system at equilibrium experiences a change in concentration, temperature,
volume, or partial pressure, then the equilibrium shifts to
counteract the imposed change. When an inert gas is added
into the equilibrium system at a constant pressure, the
partial pressures of reactive gases will decrease (dilution
effect) resulting in a shift toward the direction with the
greater number of moles of gas.
A comparison between the results for nf 0.0 and those
for nf 3.0 in Fig. 18(b) indicates that the effect of carrier
gas on hydrogen production in the presence of a CO2
adsorbent is similar to that in the absence of CO2 adsorption. However, the amounts of H2 and CO2 produced in the
presence of nitrogen are higher than those without the use
of the carrier gas, whereas the amounts of CO and CH4
production in the presence of the carrier gas are lower than

international journal of hydrogen energy 34 (2009) 72087222

those without the use of the carrier gas. It is these reasons


that lead to the small relative changes due to the presence
of carrier gas.

4.

Concluding remarks

Thermodynamic analyses have been carried out on the


adsorption-enhanced steam reforming of glycerol for
hydrogen production. The effects of temperature, pressure,
water to glycerol feed ratio, fractional CO2 removal via
adsorption, and the use of a carrier gas have been investigated. It is found that the major products of the steam
glycerol system are hydrogen, methane, carbon monoxide,
and carbon dioxide under the conditions of this work and
the use of a CO2 adsorbent enhances glycerol conversion to
hydrogen. The upper limit of the number of moles of
hydrogen produced per mole of glycerol is increased from 6
to 7 due to the CO2 adsorption. Accordingly, the maximum
hydrogen concentration is increased from 66.4% to 100% due
to CO2 adsorption. The most favourable conditions for the
steamglycerol reforming for hydrogen production are
shown to be T w800850 K, P 1 bar and WGFR w9.0 (S/
C w3.0). Carbon formation is possible with a low water to
glycerol feed ratio and CO2 adsorption prevents carbon
formation at temperatures below w900 K. Further work is
underway to experimentally validate the findings of this
work.

Acknowledgements
The authors would like to extend their thanks to UK EPSRC
for financial support under Grants EP/F027389/1. AnSteel
Group of China is acknowledged for supporting Mr T. Zhang
in the form of a visiting research fellowship. Thanks should
be given to Professor M.D. Koretsky for the constructive
discussion.

Nomenclature
AC
aji
bj
C
f
G
H
DH298
i
K
M
N
n
nf
O
P

activity of carbon formation


number of j-type atoms in the i-species
number of j-type atoms
carbon atom
CO2 removal fraction
Gibbs free energy
hydrogen atom
formation enthalpy
number of components
equilibrium constants for reaction
total number of atoms
total number of species
number of moles
molar ratio of carrier gas (nitrogen) to the feed
reactants
oxygen atom
pressure

R
RC
S
S/C
WGFR
T
V
y
m
l
b
f

7221

universal gas constant


relative change of product production
entropy
molar ratio of steam to carbon feed
molar ratio of water to glycerol feed
temperature
volume
mole fraction
chemical potential
Lagrangian multipliers
fugacity coefficient of the gas mixture

references

[1] Cortright RD, Davda RR, Dumesic JA. Hydrogen from


catalystic reforming of biomass-derived hydrocarbons in
liquid water. Nature 2002;418:9647.
[2] Cameron DC, Koutsky JA. Final report to National Biodiesel
Development Board. Internet site: http://www.biodiesel.org/
resources/reportsdatabase/reports/gen/19941001_gen-243.
pdf; 1994.
[3] Huber GW, Shabaker JW, Dumesic JA. Raney NiSn catalyst
for H2 production from biomass-derived hydrocarbons.
Science 2003;300:20757.
[4] Dauenhauer PJ, Salge JR, Schmidt LD. Renewable hydrogen
by autothermal steam reforming of volatile carbohydrates.
Journal of Catalysis 2006;244:23847.
[5] Hirai T, Ikenaga NO, Miyake T, Suzuke T. Production of
hydrogen by steam reforming of glycerin on ruthenium
catalyst. Energy & Fuels 2005;19:17612.
[6] Adhikari S, Fernando S, Gwaltney S, Filip To SD, Bricka RM,
Steele PH, et al. A thermodynamic analysis of hydrogen
production by steam reforming of glycerol. International
Journal of Hydrogen Energy 2007;32:287580.
[7] Cong T. Adsorption enhanced steammethane reforming for
low temperature hydrogen production using solids
circulation. PhD thesis, University of Leeds; 2009.
[8] Hufton JR, Mayorga, Sircar S. Sorption-enhanced reaction
process for hydrogen production. AIChE Journal 1999;45(2):
24856.
[9] Ding Y, Alpay E. Adsorption-enhanced steammethane
reforming. Chemical Engineering Science 2000;55:392940.
[10] Cong T, Chen H, Ding Y. Adsorption enhanced steammethane
reforming reaction for hydrogen production using solids
circulation through a catalyst bed reactor. In: Proceedings of
particulate systems analysis 2008. Warwickshire, UK: Stratfordupon-Avon; 2008. 02-04th Sep. 2008.
[11] Koretsky MD. Engineering and chemical thermodynamics.
John Wiley & Sons, Inc; 2004.
[12] Garcia EY, Larorde MA. Hydrogen production by the steam
reforming of ethanol: thermodynamic analysis. International
Journal of Hydrogen Energy 1991;16(5):30712.
[13] Yaws CL. Handbook of thermodynamic and physical
properties of chemical compounds, knovel. Online version
available at: http://knovel.com/web/portal/browse/display?_
EXT_KNOVEL_DISPLAY_bookid 667&VerticalID 0; 2003.
[14] Lide DR. CRC handbook of chemistry and physics. 86th ed.
Florida, USA: CRC Press; 2005.
[15] Prins M, Ptasinski K, Janssen F. Thermodynamics of gas-char
reaction: first and second law analysis. Chemical
Engineering Science 2003;58:100311.
[16] Slinn M, Kendall K, Mallon C, Andrews J. Steam reforming of
biodiesel by-products to make renewable hydrogen.
Bioresource Technology 2008;99:58518.

7222

international journal of hydrogen energy 34 (2009) 72087222

[17] Mas V, Kipreos R, Amadeo N, Laborde M. Thermodynamic


analysis of ethanol/water system with the stoichiometric
method. International Journal of Hydrogen Energy 2006;31:218.
[18] Adhikari S, Fernando S, Haryanto A. A comparative
thermodynamic and experimental analysis on hydrogen
production by steam reforming of glycerin. Energy & Fuels
2007b;21:230660.
[19] Lwin Y, Daud W, Mohamad A, Yaakob Z. Hydrogen
production from steammethanol reforming:
thermodynamic analysis. International Journal of Hydrogen
Energy 2000;25(1):4753.
[20] Ding Y, Alpay E. Equilibria and kinetics of CO2 adsorption on
hydrotalcite adsorbent. Chemical Engineering Science 2000;
55:392940.
[21] Waldron WF, Hufton JR, Sircar S. Production of hydrogen by
cyclic sorption enhanced reaction process. AIChE Journal
2001;47:14779.
[22] Reijers H, Valster-Schiermeier S, Cobden P, Brink R.
Hydrotalcite as CO2 sorbent for sorption-enhanced steam
reforming of methane. Industrial and Engineering Chemistry
Research 2006;45:252230.
[23] Balasubramanian B, Ortiz AL, Kaytakoglu S, Harrison DP.
Hydrogen from methane in a single-step process. Chemical
Engineering Science 1999;54:354352.

[24] Harrison DP, Peng Z. Low-carbon monoxide hydrogen by


sorption-enhanced reaction. International Journal of
Chemical Reaction Engineering 2003;1:19.
[25] Dou B, Dupont V, Rickett G. Hydrogen production by
sorption-enhanced steam reforming of glycerol. Bioresource
Technology 2009;100:35407.
[26] Mayorga SG. Carbon dioxide adsorbents containing
magnesium oxide suitable for use at high temperatures. U.S.
patent no. 6280503 B1; 2001.
[27] Kato M, Yoshikawa S, Nakagawa K. Carbon dioxide
absorption by lithium orthosilicate in a wide range of
temperature and carbon dioxide concentrations. Journal of
Material Science Letters 2002;21:4857.
[28] Fauth DJ, Frommell EA, Hoffman JS, Reasbeck RP,
Pennline HW. Eutectic salt promoted lithium zirconate:
novel high temperature sorbent for CO2 capture. Fuel
Process Technology 2005;86:150321.
[29] Zhang B, Tang X, Li Y, Xu Y, Shen W. Hydrogen production
from steam reforming of ethanol and glycerol over ceriasupported metal catalysts. International Journal of Hydrogen
Energy 2007;32:236773.
[30] Byrd AJ, Pant KK, Gupta RB. Hydrogen production from
glycerol by reforming in supercritical water over Ru/Al2O3
catalyst. Fuel 2008;87:295660.

You might also like