You are on page 1of 9

Fuel Processing Technology 134 (2015) 107–115

Contents lists available at ScienceDirect

Fuel Processing Technology


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

A thermodynamic analysis of hydrogen production via aqueous phase


reforming of glycerol
A. Seretis, P. Tsiakaras ⁎
Department of Mechanical Engineering, School of Engineering, University of Thessaly, Volos, Greece

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

Article history: The thermodynamic analysis of glycerol (C3H8O3) aqueous phase reforming (APR) has been performed following
Received 25 November 2014 the Gibbs free energy minimization method. The effect
  of the operating parameters, i.e. water to glycerol mass
Received in revised form 20 January 2015 ratio (W/G = 4 − 14), pressure ratio P=P Hsat2 O ¼ 1−2 and temperature (T = 300 − 550K) on: i) hydrogen pro-
Accepted 20 January 2015
duction, ii) CH4 production and iii) carbon formation was investigated. For the range of the examined conditions,
Available online 5 March 2015
the conversion of glycerol approached 100%. It was found that the maximum H2 selectivity reaches (1st case) the
Keywords:
maximum stoichiometric value of 70%, while the methanation reaction (2nd case) can be reduced at high
Aqueous phase reforming temperatures and low pressures but not eliminated, indicating that methanation is thermodynamically favored
Glycerol thermodynamic analysis over H2 production. It was also found (3rd case) that almost 80% of the amount of glycerol is converted into
H2 production carbon, while at pressure ratios P=P Hsat2 O ≤ 1:4 and temperature values T N 400 K carbon formation can be
eliminated. For all the cases studied, the optimal W/G for H2 production was found equal to 9 under thermodynamic
equilibrium conditions and the best conditions to optimize H2 production and minimize CH4 and carbon formation
are as follows: 450 ≤ T ≤ 550K, 1≤ P=P Hsat2 O ≤1:2 and 9 ≤ W/G ≤ 14.
© 2015 Elsevier B.V. All rights reserved.

1. Introduction (Eq. (2)) is thermodynamically favorable; thus in a single stage reactor


it is possible to produce H2 gas, with negligible amounts of CO, appropri-
Renewable fuels are becoming increasingly important because of ate for fuel cell applications. Moreover, low T minimizes the undesirable
depletion of petroleum reserves and increasing environmental concerns decomposition reactions [10].
associated with fossil fuel utilization. Biodiesel is considered one prom- The production of H2 via glycerol's: i) steam reforming, ii) partial ox-
ising renewable fuel, the production and the combustion process of idation), iii) autothermal reforming, iv) supercritical water reforming
which are considered as carbon neutral [1–3]. However biodiesel pre- processes [7,11] and v) APR has been experimentally investigated
sents one significant drawback: for every 10 l of biodiesel produced, [12–26]. Despite the fact that the thermodynamic analysis for the first
roughly 1 l of glycerol is created as a byproduct. Although glycerol four processes has been extensively reported [27–36], there are only
does have its industrial uses, current biodiesel production has already few investigations concerning the thermodynamic study of the reaction
exceeded market demand, leaving large amounts of practically worth- of glycerol's APR for H2 production. Luo et al. [37] studied thermody-
less glycerol in the manufacturers' hands [4]. It is therefore necessary namically the APR of glycerol taking into account the reforming of
to find effective and economical approaches to utilize glycerol [5,6]. glycerol, the WGS reaction and the reaction of methanation at: T =
Glycerol can be used to produce a variety of chemicals and fuels [5,6] 300–500 K and P = 1–50 atm. The combination of oxidation in APR re-
including H2 [7,8], the simplest and the most abundant element not free garding with and without the methanation reaction was also examined
in nature. The H2 demand is continuously growing due to the technolog- thermodynamically by Luo et al. [38], demonstrating that the molar
ical advancements in fuel cell industry. ratio of oxygen/glycerol plays an important role in controlling the gas-
Hydrogen can be produced from glycerol via steam reforming, eous phase H2/CO ratio. However, in both investigations no information
partial oxidation, gasification, autothermal reforming, supercritical about solid carbon formation is presented. Therefore, a determination of
water reforming processes and aqueous-phase reforming (APR) [9]. suitable operating conditions that minimize the formation of carbon is
Among them, APR is the most promising process due to the following interesting. Moreover, the temperature range and the system's pressure
advantages: i) it can be carried out at low T and ii) it can reduce process have been extended up to 550 K and 122 bar and the water to glycerol
cost because it is not necessary to vaporize water and glycerol. In addi- mass ratio effect is investigated for first time for the APR of glycerol.
tion, APR occurs at T and P values where the water–gas shift reaction Furthermore, in the few published comparative studies on theoreti-
cal and experimental analyses of steam-glycerol reforming for hydrogen
⁎ Corresponding author. Tel.: +30 24210 74065. production [39–41], only one investigates the aqueous phase reforming
E-mail address: tsiak@uth.gr (P. Tsiakaras). of glycerol [42]. The comparative results showed that the distribution of

http://dx.doi.org/10.1016/j.fuproc.2015.01.021
0378-3820/© 2015 Elsevier B.V. All rights reserved.
108 A. Seretis, P. Tsiakaras / Fuel Processing Technology 134 (2015) 107–115

production, ii) CH4 production and iii) carbon formation, was inves-
Nomenclature tigated. Finally, the appropriate conditions which offer the highest
H2 production and eliminate the methanation and the carbon forma-
a attraction parameter in the Peng–Robinson (P–R) equa- tion are determined.
tion, N m4/mol2
ai attraction parameter for species i in the P–R equation, 2. Theory
Nm4/mol2
aij binary interaction coefficient of the attraction parame- 2.1. Assumptions—definitions
ter a for a mixture of species i and j, Nm4/mol2
aik number of atoms of the kth element present in each The APR process involves the following two main reactions:
molecule of species i, –
Ak total number of atomic masses of the kth element in the C–C cleavage leading to CO and H2,
system, – 0
ðC3 H8 O3 Þl ⇄3CO þ 4H2 ΔH 250 C ¼ 338:02kJ=mol ð1Þ
aj attraction parameter of the P–R equation of species j, N
m4/mol2
am attraction parameter of the P–R equation in the mixture, Water–gas shift reaction (WGS),
Nm4/mol2
A,B constants defined in Eq. (22), – CO þ H 2 O⇄CO2 þ H 2
0
ΔH250 C ¼ −41:17kJ=mol ð2Þ
b repulsion parameter of the P–R equation, m3/mol
bm repulsion parameter of the P–R equation in the mixture, Overall reaction of aqueous phase reforming (1) + (2)
m3/mol
0
bi repulsion parameter of the P–R equation of species i, ðC3 H8 O3 Þl þ 3H2 O⇄3CO2 þ 7H 2 ΔH250 C ¼ 214:51kJ=mol ð3Þ
m3/mol
c,d function defined by Eq. (18), – Methanation reactions,
^f the fugacity of species i, Pa
i
foi the standard state fugacity of species i, Pa 0
CO þ 3H2 ⇄CH 4 þ H 2 O ΔH250 C ¼ −206:11kJ=mol ð4Þ
Gt total Gibbs free energy, J/mol
Goi the standard Gibbs free energy for species i, J/mol
CO2 þ 4H2 ⇄CH 4 þ 2H 2 O
0
ΔH 250 C ¼ −164:94kJ=mol ð5Þ
Gi partial molar Gibbs free energy for species i, J/mol
ΔGofi standard Gibbs function of species i formation, J/mol
ΔGof CðsÞ standard Gibbs function of solid carbon formation, J/mol Carbon formation reactions,
m function of the acentric factor ω defined in Eq. (16), –
nc moles of solid carbon, mol 2CO⇄CO2 þ C
0
ΔH250 C ¼ −172:43kJ=mol ð6Þ
ni moles of species i, mol
P pressure, Pa 0
CO þ H 2 ⇄C þ H 2 O ΔH250 C ¼ −131:26kJ=mol ð7Þ
Pc critical pressure, Pa
Po standard pressure, Pa CO2 þ 2H2 ⇄C þ 2H2 O
0
ΔH 250 C ¼ −90:09kJ=mol ð8Þ
P Hsat2 O saturation vapor pressure
R gas constant, J/mol K
CH4 ⇄C þ 2H 2
0
ΔH250 C ¼ 74:85kJ=mol ð9Þ
T temperature, K
Tc critical temperature, K
Tr reduced temperature, – Aspen Plus 11.1 software was used for the thermodynamic calcula-
V volume, m3 tions and the RGibbs reactor was selected for the calculations using
Vt volume of mixture, m3 the Peng–Robinson with Boston–Mathias Modifications as property meth-
yi molar fraction of species i in the gas phase, – od [43,44]. This equation is flexible and suitable for high system pres-
yj molar fraction of species j in the gas phase, – sures. The products C3H8O3, H2O, H2, CO, CO2, CH4, and C (graphite)
Z compressibility factor, – were taken into account in the thermodynamic analysis, and the follow-
ψ Lagrange multiplier combination function, – ing three cases were considered separately:
δij binary interaction coefficient of species i and species j in
the mixtures, – Case 1) C3H8O3, H2O, H2, CO, CO2 (overall reaction of APR)
ϕ^ fugacity coefficient of species i, –
i Case 2) C3H8O3, H2O, H2, CO, CO2, CH4 (plus methane production)
λκ Lagrange multiplier, – Case 3) C3H8O3, H2O, H2, CO, CO2, CH4, C (plus carbon formation)
μi chemical potential of species i, J/mol
ω acentric factor, – Case 1 includes only the compounds involved in the main reforming
reaction without methanation and carbon formation. In Case 2, the
methanation reaction was included, while in Case 3 both reactions of
methanation and carbon formation were included.
The aqueous phase reforming of glycerol is carried out at a tem-
gas products was very close to that predicted by thermodynamic calcu- perature range of up to 550 K where the water–gas shift reaction
lation (without methanation) when used a platinum catalyst, and (Eq. (2)) is thermodynamically favorable while at higher temperatures
methanation could be suppressed by selection of catalyst. the steam reforming of glycerol can be carried out, with the advantages
In the present study, the thermodynamic analysis of glycerol's APR and disadvantages of this process [45]. In all cases, W/G = 9 and T =
for H2 production based on Minimum Gibbs Free Energy Principle has 300 − 550K have been used. One of the basic characteristics of APR is
been realized. The effect of the operating parameters, such as water/ that the pre-existing H2O in the system is liquid, which means that the
 
glycerol mass ratio (W/G = 4 − 14), pressure ratio P=P Hsat2 O ¼ 1−2 pressure of the system (P) needs to be kept above the saturation vapor
 
and operating temperature (T = 300 − 550K) on: i) the H2 pressure P Hsat2 O at the reaction temperature (T). A well-established
A. Seretis, P. Tsiakaras / Fuel Processing Technology 134 (2015) 107–115 109

method [37,46] was used to replace the P with the P=P Hsat2 O. The calcu- bðT Þ ¼ bðT c Þ ð14Þ
lations were carried out at P=P Hsat2 O
¼ 1−2. Choosing this pressure
range, the values of the system's pressure are far enough for the anal- where a(Tr, ω) is a dimensionless function of the reduced T and the
ysis reaching up to 122 bars, as given in Table 1. Additionally the acentric factor. At T lower than the critical point the following equation
above three cases for W/G = 4 − 14 and T = 300 − 550K at a con- is used:
stant P=P Hsat2 O ¼ 1:1 were studied and compared.   
1=2 1=2
It was assumed: i) glycerol in liquid phase, ii) water both in liquid a ðT r ; ωÞ ¼ 1 þ m 1−T r ð15Þ
and gaseous phase in equilibrium, iii) carbon in solid phase and iv)
the remaining in gaseous phase. It is also important to note, that except where,
the methanation and carbon formation, other undesired reactions
as dehydration, and/or hydrogenation/dehydrogenation to produce 2
m ¼ 0:37464 þ 1:54226ω−0:26992ω ð16Þ
liquid products (methanol, ethanol, 1-propanol, acetol, ethylene glycol,
propylene glycol etc.) are possible. Appreciating that the selectivity of At T higher than the critical point Eq. (15) gives unrealistic results.
the process can be directed by the operating conditions and the nature For this case Boston and Mathias derived the following alternative [44]:
of catalyst used, the liquid products were not assumed in this study.
C n h  io2
Carbon formation is defined as C ¼ 3Glyf , where Cf is the outlet molar d
in aðT r ; ωÞ ¼ exp c 1−T r ð17Þ
flow rate (kmol/h) of solid carbon formed and Glyin is the inlet molar
flow rate of glycerol, which means that if the entire quantity of glycerol
d ¼ 1 þ m=2; c ¼ 1−1=d ð18Þ
is converted to carbon then C = 1. The three in the denominator corre-
sponds to the number of carbons in glycerol.
The quantity of the other products is defined as molar fraction in dry Based on the van der Waals classical rule of mixtures, the mixture
basis (the mole fractions are calculated without taking into account the parameters am and bm can be written as follows [43]:
amount of steam).
X
N
bm ¼ yi bi ð19Þ
2.2. Peng–Robinson equation of state i¼1

The advantage of this equation is that it can accurately and easily N X


X N
represent the relation among T, P, and phase compositions in binary am ¼ yi y j ai j ð20Þ
and multicomponent systems. It only requires the critical properties i¼1 j¼1
and acentric factor for the generalized parameters and little computer
time and leads to good phase equilibrium prediction. The Peng–Robinson To increase the calculation accuracy, the binary interaction coeffi-
equation of state is superior in the prediction of liquid phase densities cient δij can be introduced to Eq. (20):
[43]. Therefore, it is extensively used in gas–liquid equilibrium calcula-
 
tions at moderate or high system pressures [37,38,46]. The functional 1=2 1=2
ai j ¼ 1−δi j ai a j ð21Þ
form of the Peng–Robinson equation of state of a pure fluid can be
expressed as follows [43]:
When i = j, δij = 0 and when i ≠ j, δij is an empirically determined
RT a binary interaction for species i and species j and its value usually can
P¼ − ð10Þ
V−b V ðV þ bÞ þ bðV−bÞ be obtained by fitting experimental data of mixtures.
Eq. (10) can be rewritten as follows [43]:
where a represents a measure of the intermolecular attraction force,
   
and b is related to the size of the molecule. When Eq. (10) is applied 3 2 2 2 3
Z −ð1−BÞZ þ A−3B −2B Z− AB−B −B ¼ 0 ð22Þ
at the critical point, the following equations are obtained:

R2 T 2c where,
aðT c Þ ¼ 0:45724 ð11Þ
Pc
am P b P PV
A¼ ;B ¼ m ;Z ¼ ð23Þ
RT R2 T 2 RT RT
bðT c Þ ¼ 0:07780 c ð12Þ
Pc
According to the expression of the fugacity coefficient of species i:
At T other than the critical point, these equations are as follows:
ZV t "   #
^ ¼ 1 RT ∂P
aðT Þ ¼ aðT c ÞaðT r ; ωÞ ð13Þ ln ϕ ι − dV t − ln Z ð24Þ
RT Vt ∂ni T;V t ;n jð j≠iÞ

Table 1
Pressure of the system in bar (values in italic) at different pressure ratios and temperature Using the Peng–Robinson equation of state to represent the fluid
values.
^ of species i in a mixture can be calculated
phase, the fugacity coefficient ϕ ι
Pressure ratio Temperature (K) according to the following equation:
P=P Hsat2 O
300 350 400 450 500 550

1 0.031 0.41 2.47 9.35 26.50 61.20


^ ¼ bi
1.2 0.038 0.50 2.96 11.22 31.80 73.44 ln ϕι ðZ−1Þ− ln ðZ−BÞ
bm pffiffiffi  1
1.4 0.044 0.58 3.45 13.09 37.10 85.68 2 3 0
1.6 0.050 0.67 3.95 14.96 42.40 97.92
A 4 bi 2 X
N Zþ 2þ1 B
1.8 0.057 0.75 4.44 16.83 47.70 110.16 þ pffiffiffi − yi ai j 5 ln @ pffiffiffi  A ð25Þ
2 0.063 0.83 4.94 18.70 53.00 122.40 2 2B bm am j¼1 Z− 2−1 B
110 A. Seretis, P. Tsiakaras / Fuel Processing Technology 134 (2015) 107–115

Fig. 1. Mole fraction of a) H2, b) CO and c) CO2 (dry basis) as a function of T and P=P sat
H2 O for constant W/G = 9 without methanation and carbon formation (case 1).

2.3. Gibbs free energy minimization The combination function ψ can be constructed using the Lagrange
multiplier method as follows:
The Gibbs free energy minimization method is believed to be an ef-
fective way to determine the equilibrium compositions of each compo-
nent in a complex system [47]. As the system reaches equilibrium at the t
X
M
ψ¼G þ λk aik ð28Þ
appropriate T and P, Gibbs free energy minimization can be used directly k¼1
to determine the equilibrium composition of each component in the
mixture. The main benefit of the Gibbs free energy minimization is
that there is no need to check the specific information for each reaction The minimum value of the combination function ψ is equal to the
in the system. The total Gibbs free energy of the system is equal to the primitive function Gt. In Eq. (28), the Lagrange multiplier λk is constant,
sum of the standard Gibbs free energies of all the pure components and its value can be calculated. For reaction equilibrium in the system, if
and that of the mixed system at equilibrium. This can be expressed as ^f ¼ y ϕ
^ P; f o ¼ P o ; Go ¼ ΔGo and derivation of the combination func-
i i i i i fi
follows:
tion ψ with respect to ni is zero, then this equation can be written as
^f follows:
t
X
N X
N X
N
o
X
N
i
G ¼ ni Gi ¼ ni μ i ¼ ni Gi þ RT ni ln ð26Þ
f oi
i¼1 i¼1 i¼1 i¼1 !
∂ψ X N
o yϕ^P X M
¼ ni ΔG f i þ RT ln i oi þ λk aik ¼0 ð29Þ
The minimum value of Eq. (26) can be obtained by finding the set of ∂ni i¼1
P k¼1
ni that minimizes Gt at constant T and P. This minimization is subject to
the constraints of elemental balance of all mixture components.
Constraint of Eq. (27) gives a set of (N + M) equations; thus
X
ni aik ¼ Ak ; k ¼ 1; 2…; M ð27Þ the values of ni and λk can be obtained under the minimization
i condition.
A. Seretis, P. Tsiakaras / Fuel Processing Technology 134 (2015) 107–115 111

When carbon formation is considered in the reaction system [46,48], without taking into account the amount of steam. As it can be seen, for a
then the equilibrium condition is given by the following equation and wide range of conditions the molar fraction of CO is practically zero,
the constraints are the same as Eq. (27): while the molar fractions of H2 and CO2 approach the stoichiometric
! maximum theoretical value 0.7 and 0.3 respectively. Certainly at higher
X
N −1 ^P X
yi ϕ M   T the fraction of CO slightly increases, and therefore the fractions of
o i o
ni ΔG f i þ RT ln o þ λk aik þ nc ΔG f CðsÞ ¼ 0 ð30Þ
P H2 and CO2 are reduced and the effect becomes more pronounced at
i¼1 k¼1
high P=P Hsat2 O . This is because the APR as endothermic is favored with in-
creasing T while the WGSR as exothermic is not favored with increasing
3. Results and discussion T. Luo et al. [37] reported a similar behavior for the mole fractions of
gaseous products in equilibrium. On the other hand, in Fig. 1a, when
Thermodynamic analysis of the effect of reaction variables on the pro- the yH2 axis is illustrated in a larger scale (0.6–0.7), it is noticed that
duction of H2, CH4 and solid carbon was studied. For the range of T, P=P Hsat2 O the mole fraction of H2 remains practically constant.
and W/G values examined, the conversion of glycerol approached always
100%. The three cases for T = 300 − 550K and P=P Hsat2 O ¼ 1−2 at a con- 3.2. Case 2 (plus methane production)
stant W/G = 9 were analyzed below.
For the case 2 all the compounds involved in the case 1 plus CH4 have
3.1. Case 1 (mainly hydrogen production) been considered as the reaction products. Therefore, the possible reac-
tions that are taking place in this case, are the APR of glycerol, the
Case 1 includes only the compounds involved in the main reforming WGS and the methanation reactions [Eqs. (1),(2),(4) and (5)].
reaction without methanation and solid carbon formation. Therefore, Fig. 2 shows the mole fractions of gaseous products in equilibrium as
the possible reactions that are taking place in this case, are the aqueous a function of T and P=P Hsat2 O (dry basis). In this case, the maximum value of
phase reforming of glycerol and the water–gas shift [Eqs. (1) and (2)]. CO is significantly lower than that in case 1 (0.000027 to 0.0034)
In Fig. 1 the mole fractions of gaseous products in equilibrium as a because it is obviously consumed in the first reaction of methanation
function of T and P=P Hsat2 O are depicted. The mole fractions are calculated (30). The mole fraction of CO increases with increasing T, while the

Fig. 2. Mole fraction of a) H2, b) CO, c) CO2 and d) CH4 (dry basis) as a function of T and P=P Hsat2 O for constant W/G = 9 without carbon formation (case 2).
112 A. Seretis, P. Tsiakaras / Fuel Processing Technology 134 (2015) 107–115

effect of the pressure is very small. The amount of H2 tends to increase at Finally, Fig. 4 shows the vapor fraction (water plus gaseous prod-
T N 450 K and P=P Hsat2 O ≤1:2. Nevertheless the maximum value of H2 is ucts) of the system for the three cases. As it is below the unity means
also significantly reduced compared with the case 1 (0.045 vs. 0.7) be- that there is liquid phase which is of course water, as glycerol has
cause it is consumed in the methanation reactions, which is undesirable. completely reacted. At P=P Hsat2 O N 1 for all T values the liquid phase exists,
Moreover, in Fig. 2c and d, it is obvious that the amount of glycerol is confirming the APR process at these conditions. The vapor fraction of
converted entirely into CO2 and CH4 (0.4166 and 0.5833 respectively) the gas phase in the second and third case is smaller than that in the
with the exception of high T and low P=P Hsat2 O . Therefore, two of the first case, which means that the liquid phase is to a greater extent in
keys of glycerol's APR are: i) to choose a selective catalyst to depress these cases because an excess of water is produced by the reactions of
the methanation reaction and ii) to conduct the process at high T and methanation and carbon formation.
low P=P Hsat2 O as far as possible in order to have a liquid phase. Having studied the above three cases in a range of P=P Hsat2 O and T, we
concluded that the best conditions for producing H2 from catalytic
reforming of glycerol in the aqueous phase are low P=P Hsat2 O and high T
3.3. Case 3 (plus carbon formation) and more specifically 450 ≤ T ≤ 550K and 1 ≤P=P Hsat2 O ≤1:2. At these
operating conditions where carbon formation was eliminated, the
In this case both the reactions of methanation and carbon formation experimental results [49,42] with platinum catalyst (that suppress the
were considered. The possible reactions taking place in this case are the methanation reactions) are very close to that predicted thermodynam-
APR of glycerol, the WGS, the methanation reactions and the carbon ically by case 1.
formation reactions [Eqs. (1), (2) and (4–9)].
In Fig. 3, the mole fractions of the gaseous products in equilibrium
and the carbon formation (Fig. 3e) as defined in Section 2.1 are present- 3.4. Water to glycerol ratio effect (three cases)
ed as a function of T and P=P Hsat2 O. It should be noted that about 80% of the
Another important issue is the water to glycerol weight ratio. Below
amount of glycerol is converted into carbon at P=P Hsat2 O ≥1; 4 and
we study the three cases at T = 300 − 550K, for three W/G = 14, 9, 4
T b 400 K and the gaseous products in this range of conditions consist
and constant P=P Hsat2 O ¼ 1:1. This ratio was chosen because the range 1≤
mainly of CH4. At high T the exothermic carbon formation reactions
and methanation reactions are not favored, while at low P=P Hsat2 O the spe- P=P Hsat2 O ≤1:2 was found to be the optimal for hydrogen production. Fig. 5
cific reactions shift towards the left side, which is desirable. Therefore, at shows the mole fractions of H2 (dry basis) in equilibrium as a function of T
and W/G for the three cases, respectively. In all cases, it is observed that
P=P Hsat2 O ≤ 1:4 and T N 400 K, carbon formation is not favored and case 3 is
the mole fractions of H2 are increased by increasing the W/G (at high T).
converted in case 2 where it is necessary to have the presence of a selec-
This is because increasing the W/G ratio, i.e. increasing the amount of
tive catalyst that will prevent the methanation. At T N 450 K and P=P Hsat2 O water, the WGS reaction (Eq. (2)) shifts to the right. Moreover, for
≤1:2 (P ≥P Hsat2 O in order to have liquid phase) the mole fraction of H2 W/G ≤ 9 the mole fraction of H2 sharply decreases, while for W/G ≥ 9
tends to increase. the mole fraction of H2 slightly changes. Taking into consideration that

Fig. 3. Mole fraction of a) H2, b) CO, c) CO2, d) CH4 (dry basis) and e) carbon formation as a function of T and P=P Hsat2 O for constant W/G = 9 regarding methanation and carbon formation
(case 3).
A. Seretis, P. Tsiakaras / Fuel Processing Technology 134 (2015) 107–115 113

Fig. 4. Vapor fraction of the system as a function of T and P=P Hsat2 O for constant W/G = 9 for the a) case 1, b) case 2 and c) case 3.

at higher W/G required large amounts of water and therefore more ener- minimization, using the Peng–Robinson with Boston–Mathias Modifications
gy for the APR process, the optimal W/G ratio seems to be 9. method by the aid of Aspen Plus simulation software. By varying the re-
action conditions such as T, P=P sat
H 2 O and W/G ratio, glycerol can be con-
4. Conclusions verted into a H2 rich mixture or in a mixture containing H2, CO2, CO
and CH4 as well as solid carbon. The analysis was performed at:
In this work the thermodynamic equilibrium calculations have i) temperature values of 300–550 K, ii) water to glycerol ratio of 4–14
been accomplished according to the method of Gibbs free energy and iii) P=P Hsat2 O of 1.0–2.0.

Fig. 5. Mole fraction of H2 (dry basis) as a function of T and W/G for constant P=P Hsat2 O ¼ 1:1 for the a) case 1, b) case 2 and c) case 3.
114 A. Seretis, P. Tsiakaras / Fuel Processing Technology 134 (2015) 107–115

Three different cases have been examined: The first case, in which [16] A.O. Menezes, M.T. Rodrigues, A. Zimmaro, L.E.P. Borges, M.A. Fraga, Production of
renewable hydrogen from aqueous-phase reforming of glycerol over Pt catalysts
no CH4 production and carbon formation were taken into account. In supported on different oxides, Renewable Energy 36 (2011) 595–599.
this case it was observed that in a wide range of conditions the fraction [17] R.L. Manfro, A.F. da Costa, N.F.P. Ribeiro, M.M.V.M. Souza, Hydrogen production by
of CO is practically zero, while fractions of H2 and CO2 approach the aqueous-phase reforming of glycerol over nickel catalysts supported on CeO2, Fuel
Processing Technology 92 (2011) 330–335.
stoichiometric maximum theoretical value of 0.7 and 0.3 respectively. [18] Y. Guo, M.U. Azmat, X. Liu, Y. Wang, G. Lu, Effect of support's basic properties on
The second case, in which except the formation of CO, CO2 and H2, hydrogen production in aqueous-phase reforming of glycerol and correlation be-
the reaction of CH4 production (methanation) was taken into account. tween WGS and APR, Applied Energy 92 (2012) 218–223.
[19] A. Ciftci, B. Peng, A. Jentys, J.A. Lercher, E.J.M. Hensen, Applied catalysis A: general
It was observed that the maximum value of H2 is significantly reduced support effects in the aqueous phase reforming of glycerol over supported platinum
compared with the case 1 (0.045 vs. 0.7) and the amount of glycerol is catalysts, Applied Catalysis A: General 431-432 (2012) 113–119.
converted entirely into CO2 and CH4 (0.4166 and 0.5833 respectively) [20] Y. Guo, X. Liu, M.U. Azmat, W. Xu, J. Ren, Y. Wang, G. Lu, Hydrogen production by
aqueous-phase reforming of glycerol over Ni–B catalysts, International Journal of
with the exception of high T and low P=P Hsat2 O . Therefore, two of the Hydrogen Energy 37 (2012) 227–234.
keys of glycerol's APR are: i) to choose a selective catalyst to depress [21] A. Iriondo, V.L. Barrio, J.F. Cambra, P.L. Arias, M.B. Güemez, R.M. Navarro, M.C.
the methanation reaction and ii) to conduct the process at high T and Sanchez-Sanchez, J.L.G. Fierro, Hydrogen production from glycerol over nickel
catalysts supported on Al2O3 Modified by Mg, Zr, Ce or La, Topics in Catalysis 49
low P=P Hsat2 O as far as possible in order to have a liquid phase. (2008) 46–58.
In the third case, the formation of CO, CO2, H2, and CH4 and the reac- [22] N. Luo, X. Fu, F. Cao, T. Xiao, P. Edwards, Glycerol aqueous phase reforming for
hydrogen generation over Pt catalyst—effect of catalyst composition and reaction
tion of carbon formation were considered. It was found that at P=P Hsat2 O ≤ conditions, Fuel 87 (2008) 3483–3489.
1:4 and T N 400 K, carbon formation is not favored. Finally, for all the [23] K. Lehnert, P. Claus, Influence of Pt particle size and support type on the aqueous-
phase reforming of glycerol, Catalysis Communications 9 (2008) 2543–2546.
examined cases, the optimal W/G ratio for H2 production via APR of [24] J. Shabaker, G. Huber, J. Dumesic, Aqueous-phase reforming of oxygenated hydro-
glycerol was found to be 9. carbons over Sn-modified Ni catalysts, Journal of Catalysis 222 (2004) 180–191.
The present findings (need to choose a selective catalyst to depress [25] D.L. King, L. Zhang, G. Xia, A.M. Karim, D.J. Heldebrant, X. Wang, T. Peterson, Y.
Wang, Aqueous phase reforming of glycerol for hydrogen production over Pt–Re
the methanation reaction, best conditions to minimize carbon forma- supported on carbon, Applied Catalysis B: Environmental 99 (2010) 206–213.
tion, optimal W/G ratio for H2 production) may provide new opportuni- [26] G. Wen, Y. Xu, H. Ma, Z. Xu, Z. Tian, Production of hydrogen by aqueous-phase
ties for the theoretical and experimental design of glycerol's aqueous reforming of glycerol, International Journal of Hydrogen Energy 33 (2008)
6657–6666.
phase reforming, giving a contribution to the advancement of the fuel
[27] T. Pairojpiriyakul, W. Kiatkittipong, A. Soottitantawat, A. Arpornwichanop, N.
reforming technology. Laosiripojana, W. Wiyaratn, et al., Thermodynamic analysis of hydrogen production
from glycerol at energy self-sufficient conditions, Canadian Journal of Chemical En-
gineering 9999 (2011) 1–8.
Acknowledgments [28] X. Wang, M. Li, M. Wang, H. Wang, S. Li, S. Wang, et al., Thermodynamic analysis of
glycerol dry reforming for hydrogen and synthesis gas production, Fuel 88 (2009)
2148–2153.
This research has been co‐financed by the European Union
[29] H. Wang, X. Wang, M. Li, S. Li, S. Wang, X. Ma, Thermodynamic analysis of hydrogen
(European Social Fund—ESF) and Greek national funds through the production from glycerol autothermal reforming, International Journal of Hydrogen
Operational Program “Education and Lifelong Learning” of the Energy 34 (2009) 5683–5690.
National Strategic Reference Framework (NSRF)—Research Funding [30] S. Adhikari, S. Fernando, S. Gwaltney, S. Filipto, R. Markbricka, P. Steele, et al., A ther-
modynamic analysis of hydrogen production by steam reforming of glycerol, Inter-
Program: THALES, Investing in knowledge society through the Euro- national Journal of Hydrogen Energy 32 (2007) 2875–2880.
pean Social Fund, grant number MIS 379333. [31] C.K. Cheng, S.Y. Foo, A.A. Adesina, Thermodynamic analysis of glycerol-steam
reforming in the presence of CO2 or H2 as carbon gasifying agent, International
Journal of Hydrogen Energy 37 (2012) 10101–10110.
References [32] G. Yang, H. Yu, F. Peng, H. Wang, J. Yang, D. Xie, Thermodynamic analysis of hydro-
gen generation via oxidative steam reforming of glycerol, Renewable Energy 36
[1] F. Ma, M. Hanna, Biodiesel production: a review, Bioresource Technology 70 (1999) (2011) 2120–2127.
1–15. [33] F.J. Gutiérrez Ortiz, P. Ollero, A. Serrera, A. Sanz, Thermodynamic study of the super-
[2] M.F. Milazzo, F. Spina, S. Cavallaro, J.C.J. Bart, Sustainable soy biodiesel, Renewable & critical water reforming of glycerol, International Journal of Hydrogen Energy 36
Sustainable Energy Reviews 27 (2013) 806–852. (2011) 8994–9013.
[3] J.C.J. Bart, N. Palmeri, S. Cavallaro, Biodiesel Science and Technology: From Soil to Oil, [34] X. Wang, N. Wang, M. Li, S. Li, S. Wang, X. Ma, Hydrogen production by glycerol
1st ed. Woodhead Publishing Series in Energy, 2010. steam reforming with in situ hydrogen separation: a thermodynamic investigation,
[4] C.A.G. Quispe, C.J.R. Coronado, J.A. Carvalho Jr., Glycerol: production, consumption, International Journal of Hydrogen Energy 35 (2010) 10252–10256.
prices, characterization and new trends in combustion, Renewable & Sustainable [35] S. Authayanun, A. Arpornwichanop, Y. Patcharavorachot, W. Wiyaratn, S.
Energy Reviews 27 (2013) 475–493. Assabumrungrat, Hydrogen production from glycerol steam reforming for low-
[5] M. Pagliaro, R. Ciriminna, H. Kimura, M. Rossi, C. Della Pina, From glycerol to value- and high-temperature PEMFCs, International Journal of Hydrogen Energy 36
added products, Angewandte Chemie International Edition in English 46 (2007) (2011) 267–275.
4434–4440. [36] Y. Li, W. Wang, B. Chen, Y. Cao, Thermodynamic analysis of hydrogen production via
[6] M. Stelmachowski, Utilization of glycerol, a by-product of the transestrification pro- glycerol steam reforming with CO2 adsorption, International Journal of Hydrogen
cess of vegetable oils: a review, Review-Literature and Arts of the Americas 18 Energy 35 (2010) 7768–7777.
(2011) 9–30. [37] N. Luo, F. Cao, X. Zhao, T. Xiao, D. Fang, Thermodynamic analysis of aqueous-
[7] P.D. Vaidya, A.E. Rodrigues, Glycerol reforming for hydrogen production: a review, reforming of polylols for hydrogen generation, Fuel 86 (2007) 1727–1736.
Chemical Engineering and Technology 32 (2009) 1463–1469. [38] N. Luo, X. Zhao, F. Cao, T. Xiao, D. Fang, Thermodynamic study on hydrogen gener-
[8] N.H. Tran, G.S.K. Kannangara, Conversion of glycerol to hydrogen rich gas, Chemical ation from different glycerol reforming processes, Energy & Fuels 21 (2007)
Society Reviews 42 (2013) 9454–9479. 3505–3512.
[9] J. Holladay, J. Hu, D. King, Y. Wang, An overview of hydrogen production technologies, [39] S. Adhikari, S. Fernando, A. Haryanto, A comparative thermodynamic and experi-
Catalysis Today 139 (2009) 244–260. mental analysis on hydrogen production by steam reforming of glycerin, Energy &
[10] R. Cortright, R. Davda, J. Dumesic, Hydrogen from catalytic reforming of biomass- Fuels 21 (2007) 2306–2310.
derived hydrocarbons in liquid water, Nature 418 (2002) 54–57. [40] H. Chen, Y. Ding, N.T. Cong, B. Dou, V. Dupont, M. Ghadiri, P.T. Williams, A compar-
[11] S. Adhikari, S.D. Fernando, A. Haryanto, Hydrogen production from glycerol: an ative study on hydrogen production from steam-glycerol reforming: thermodynam-
update, Energy Conversion and Management 50 (2009) 2600–2604. ics and experimental, Renewable Energy 36 (2011) 779–788.
[12] M.L. Barbelli, F. Pompeo, G.F. Santori, N.N. Nichio, Pt catalyst supported on α-Al2O3 [41] X. Wang, M. Li, S. Li, H. Wang, S. Wang, X. Ma, Hydrogen production by glycerol steam
modified with CeO2 and ZrO2 for aqueous-phase-reforming of glycerol, Catalysis reforming with/without calcium oxide sorbent: a comparative study of thermodynamic
Today 213 (2013) 58–64. and experimental work, Fuel Processing Technology 91 (2010) 1812–1818.
[13] D.Ö. Özgür, B.Z. Uysal, Hydrogen production by aqueous phase catalytic reforming of [42] F.C.L. Han, Hydrogen production from aqueous phase reforming of glycerol: thermo-
glycerine, Biomass and Bioenergy 35 (2011) 822–826. dynamic analysis and experimental validation, Mater. Renew. Energy Environ.
[14] P.V. Tuza, R.L. Manfro, N.F.P. Ribeiro, M.M.V.M. Souza, Production of renewable (ICMREE), 2011 Int. Conf., vol. 1, 2011, pp. 255–258.
hydrogen by aqueous-phase reforming of glycerol over NieCu catalysts derived [43] D.-Y. Peng, D.B. Robinson, A new two-constant equation of state, Industrial and
from hydrotalcite precursors, Renewable Energy 50 (2013) 408–414. Engineering Chemistry Fundamentals 15 (1976) 59–64.
[15] B. Meryemoglu, B. Kaya, S. Irmak, A. Hesenov, O. Erbatur, Comparison of batch [44] E. Neau, O. Hernández-Garduza, J. Escandell, C. Nicolas, I. Raspo, The Soave, Twu and
aqueous-phase reforming of glycerol and lignocellulosic biomass hydrolysate, Fuel Boston–Mathias alpha functions in cubic equations of state, Fluid Phase Equilibria
(2012) 2–5. 276 (2009) 87–93.
A. Seretis, P. Tsiakaras / Fuel Processing Technology 134 (2015) 107–115 115

[45] R.R. Davda, J.W. Shabaker, G.W. Huber, R.D. Cortright, J.A. Dumesic, A review of [47] R. Perry, D. Green, J. Maloney, Perry's Chemical Engineers' Handbook, 8th ed.
catalytic issues and process conditions for renewable hydrogen and alkanes by McGraw-Hill, New York, 2008.
aqueous-phase reforming of oxygenated hydrocarbons over supported metal [48] K. Faungnawakij, R. Kikuchi, K. Eguchi, Thermodynamic analysis of carbon formation
catalysts, Applied Catalysis B: Environmental 56 (2005) 171–186. boundary and reforming performance for steam reforming of dimethyl ether,
[46] J. Xie, D. Su, X. Yin, C. Wu, J. Zhu, Thermodynamic analysis of aqueous phase Journal of Power Sources 164 (2007) 73–79.
reforming of three model compounds in bio-oil for hydrogen production, Interna- [49] Y.-C. Lin, Catalytic valorization of glycerol to hydrogen and syngas, International
tional Journal of Hydrogen Energy 36 (2011) 15561–15572. Journal of Hydrogen Energy 38 (2013) 2678–2700.

You might also like