You are on page 1of 11

Journal of Natural Gas Science and Engineering 1 (2009) 205215

Contents lists available at ScienceDirect

Journal of Natural Gas Science and Engineering


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

Hydrogen production from methane reforming: Thermodynamic assessment and autothermal reactor design
C.N. Avila-Neto, S.C. Dantas, F.A. Silva, T.V. Franco, L.L. Romanielo, C.E. Hori, A.J. Assis*
Federal University of Uberlandia, School of Chemical Engineering, Av. Joao Naves de Avila, 2121 Bl 1K, Campus Santa Monica, CEP 38408-100, Uberlandia, MG, Brazil

a r t i c l e i n f o
Article history: Received 7 October 2009 Received in revised form 4 December 2009 Accepted 5 December 2009 Available online 12 January 2010 Keywords: Methane reforming Hydrogen production Chemical equilibrium Modeling and simulation Process optimization Scilab software

a b s t r a c t
In this study, a comparative thermodynamic analysis of methane reforming reactions is conducted using an in-house code. Equilibrium compositions are calculated by two distinct methods: (1) evaluation of equilibrium constants and (2) Lagrange multipliers. Both methods result in systems of non-linear algebraic equations, solved numerically using the Scilab (www.scilab.org) function fsolve. Effects of temperature, pressure, steam to carbon ratio (S/C) (steam reforming), CH4/CO2 ratio (dry reforming), oxygen to carbon ratio (O/C) (oxidative reforming) and steam to oxygen to carbon ratio (S/O/C) (autothermal reforming) on the reaction products are evaluated. Comparisons between experimental and simulated data, published in the literature, are used to validate the simulated results. We also present and validate a small-scale reactor model for the autothermal reforming of methane (ATR). Using this model, the reactor design is performed and key operational parameters are investigated in order to increase both H2 yield and H2/CO selectivity. The reactor model considers a mass balance equation for each component, and the set of ordinary differential equations is integrated using the Scilab function ode. This ATR reactor model is able to describe the inuence of temperature on methane conversion proles, aiming to maximize hydrogen production. The experimental results and the model presented good agreement for methane conversion in all studied temperature range. Through simulated data of methane conversions, hydrogen yields and H2/CO selectivity, it is observed that the best reaction temperature to maximize the yield of hydrogen for the ATR reaction is situated between 723 and 773 K. Inside these bounds, 50% of methane is converted into products. Also, the experimental data indicates that the Ni catalyst activity is not compromised. 2009 Elsevier B.V. All rights reserved.

1. Introduction In recent years, hydrogen has been attracting great interest as a clean fuel for combustion engines and fuel cells (Ayabe et al., 2003). Among all the potential sources of hydrogen, natural gas, which has methane as its main component, has been considered a good option because it is clean, abundant and it can be easily converted to hydrogen (Dias and Assaf, 2004). Actually, the main route to produce hydrogen from methane is catalytic reforming, which includes steam reforming (SRM, Eqs. (1)(3)), dry reforming (DRM, Eq. (4)), oxidative reforming (ORM, Eqs. (5) and (6)) and autothermal reforming (ATR, the coupling of steam and oxidative reforming) (Rostrup-Nielsen, 2002). CH4 DH2 O/COD3H2 DHo [ 206kJ=mol 298 (1)

CODH2 O/CO2 DH2 DHo [ L41kJ=mol 298 CH4 D2H2 O/CO2 D4H2 DHo [ 165kJ=mol 298 CH4 DCO2 42COD2H2 DHo [ 247kJ=mol 298 CH4 D1=2O2 /COD2H2 DHo [ L36kJ=mol 298 CH4 D2O2 4CO2 D2H2 O DHo [ L803kJ=mol 298

(2) (3) (4) (5) (6)

* Corresponding author. Fax: 55 (34) 3239 4189. E-mail addresses: ajassis@ufu.br, adilsonjassis@gmail.com (A.J. Assis). 1875-5100/$ see front matter 2009 Elsevier B.V. All rights reserved. doi:10.1016/j.jngse.2009.12.003

Steam reforming of methane (Eq. (1)) is the main industrial route to produce hydrogen and synthesis gas (a mixture of hydrogen and carbon monoxide) (al-Qahtani, 1997; Pedernera et al., 2007). This reaction produces a H2/CO ratio equal to three, which is high when compared to other reforming processes. In order to decrease the amount of carbon monoxide present in the synthesis gas, the former is further processed in a watergas shift reactor where carbon monoxide is converted into hydrogen and

206

C.N. Avila-Neto et al. / Journal of Natural Gas Science and Engineering 1 (2009) 205215

carbon dioxide by reaction with steam (Eq. (2)) (Seo et al., 2002). It is known that both steam reforming of methane and the watergas shift reaction are sufcient to represent the thermodynamic equilibrium of a steam reforming of methane process. Nevertheless, Xu and Froment (Xu and Froment, 1989) demonstrated that the sum of both reactions (Eq. (3)) is also necessary to describe experimental kinetic data. Considerable attention has also been paid to dry reforming of methane (Eq. (4)). It has the advantage of consuming both methane and carbon dioxide, two major undesirable greenhouse gases, and producing a synthesis gas with a H2/CO ratio near one, which can be used for adjusting H2/CO ratio in steam reforming, suitable for FischerTropsch reactions and methanol production (OConnor et al., 2006; Yin et al., 2007). Nevertheless, the employment of CO2 to produce synthesis gas augments the risk of carbon formation (alQahtani, 1997; Laosiripojana et al., 2005), since the process yields large amounts of CO and H2 is consumed substantially through the reverse of the watergas shift reaction (al-Qahtani, 1997). Dry reforming of methane may also be processed with gases produced from anaerobic decomposition of organic materials, such as biogas, which contains CO2 and CH4 in adequate ratios for this reaction. This process is of particular interest, since it allows taking advantage on the CO2 as oxidizer for the reaction, and thus, diminishing the concentration costs of this species (Benito et al., 2007). The two previous reforming systems (steam and dry reforming) are extremely endothermic and require high operational temperatures in order to obtain reasonable equilibrium conversions. This leads to high operational costs. Besides, the necessity to operate in such severe conditions can result on catalyst deactivation by sintering or coke deposition (Eq. (7)). CH4 /Cs D2H2 DHo [ 74kJ=mol 298 (7)

Finally, since there is still a limited number of studies in the literature referring to the reactor design and to the optimization of the operational conditions for the ATR reaction, the last objective is to present and to validate a reactor model in small scale for this reaction and to compare the results with the thermodynamics calculations. In addition, aiming to maximize the production of H2, the effects of reactor temperature on methane conversions, H2 yield and selectivity to H2 formation are investigated. Both analyses are conducted using in-house codes developed in the open-source software Scilab (INRIA ENPC,). Although we know the existence of a free and open-source software such as Cantera (Goodwin, 2004) that is a suite of object-oriented software tools for problems involving chemical kinetics, thermodynamics, and/or transport processes, we decided to use Scilab due to its much more mature development status, with matrices as the main data type, and the possibility to use the present developed codes in further studies regarding reactor optimization and automatic control in a straightforward way. 2. Modeling 2.1. Thermodynamic equilibrium Equilibrium compositions may be calculated by two distinct methods: (i) Evaluation of Equilibrium Constants (EEC) and (ii) Lagrange Multipliers (LM). As mentioned by (Smith et al., 2000), the equilibrium state has two distinctive characteristics for the given temperature and pressure: (i) the total Gibbs energy (Gt) is at a minimum and (ii) its differential should be zero. Each of these criteria may serve as a criterion of equilibrium. Thus, we can write an expression for Gt as a function of the extent of the reaction (x) and look for the value of x which minimizes Gt, or we may differentiate the expression, and equal it to zero, and solve for x. The former method is employed in the Lagrange multipliers procedure while the second methodology leads to the method of equilibrium constants. The results from these two methods should be ideally the same if the reaction system is well posed and the correlations for specic heat capacities, fugacity coefcients, virial coefcients, etc., are the same in both methods. In this study, while the LM method was applied to all reforming systems described in the previous section, the EEC method was only applied to steam and dry methane reforming. Both methods result in systems of nonlinear algebraic equations, solved numerically. 2.1.1. Evaluation of equilibrium constants method (EEC) An independent multi-reaction system is organized and each reaction is associated with a reaction coordinate (xj) and with an equilibrium constant (Kj). The equilibrium constant for each reaction in gas phase is described by Smith et al. (2000):

Therefore, in order to reduce these costs, oxidative reforming of methane has been investigated as an alternative. The partial oxidation of methane (Eq. (5)) is a slightly exothermic reaction that produces a H2/CO ratio around two, which is more adequate for FischerTropsch synthesis (Corbo and Migliardini, 2007). However, increasing the amount of oxygen in the feed may lead to total oxidation of methane (Eq. (6)), which produces CO2 and H2O. The coupling of steam reforming and partial oxidation of methane has as the main advantage the possibility to use the heat generated in the exothermic reaction as a source of energy for the endothermic reaction (Li et al., 2004; Mukainakano et al., 2007). This process is called autothermal reforming of methane. Since this system presents higher energy efciency and a more satisfactory H2/CO ratio for H2 production, this reaction can be considered an important alternative route for hydrogen production (Dias and Assaf, 2004). The conversion of methane and the selectivity of the reactions to hydrogen or synthesis gas depend on the values of several variables such as temperature, pressure, and reagents feed ratio, among others. In this way, although the thermodynamic equilibrium of reform reactions is fairly discussed in the literature, there is a lack of studies that present and compare all reforming reactions in the same basis. Additionally, the set of linearly independent reactions that describe thermodynamically each reforming system, as well as the conditions in which each reaction occurs is not clearly described in the literature. Therefore, this work has three main objectives. The rst one is to conduct a comparative thermodynamic analysis for the methane reforming process and to assess the inuence of key operational variables on chemical equilibrium. The second one is to determine the linearly independent reactions that describe in a satisfactory way the composition of the reformate of each reaction system.

Y ni b i P=P o n Kj a
i

(8)

The second term of Eq. (8) represents the product of the activities, b i , of all species in the mixture. The activities supply the a connection between the equilibrium state of interest and the standard states of the individual species. By denition, the activities are related to the fugacities of each species in the mixture, bi , by: f

b i hbi =fio a f

(9)

All methane reforming reactions occur in gaseous state and consequently the fugacity of species i in this type of reactional system is the fugacity of species i in a mixture of gases. Following Smith et al. (2000), the standard state of a gas is the state of pure

C.N. Avila-Neto et al. / Journal of Natural Gas Science and Engineering 1 (2009) 205215

207

ideal gas in the standard state pressure, Po, of 1 bar. Thus, for reactions occurring in gas phase, as the fugacity of an ideal gas is equal to its pressure, fio P o for each species i. The fugacity reects in its turn the non-idealities of the mixture and it is a function of temperature, pressure and composition. For species in gas phase, the fugacity is related to the fugacity coefcient of species i in b mixture, f i , and to the mole fraction of species i, yi, by the expression:

system solving. Equilibrium constants method suffers from difculties in convergence when the equilibrium constants are quite different from each other. Overall speaking, each method has its advantages and disadvantages and these are somehow dependent on each specic problem considered. 2.1.3. Simulation method Almost all methane reforming reactions occur at low pressures. Consequently, it could be assumed that the fugacity coefcient of b species i in mixture, f i , is one. Nevertheless, since pressure effect is b also studied in this work, this hypothesis is not employed and f i is estimated by a generalization of the virial equation of state truncated at the second virial coefcient Smith et al. (2000). The second virial coefcient is calculated by the correlation of Pitzer and Curl which was modied by Meng et al. (2004).

b fyP b fi i i

(10)

Therefore, substituting Eqs. (9) and (10) in Eq. (8), one can obtain an expression to calculate the thermodynamic equilibrium as a function of T and P by means of the EEC method.

Y
i

b yi f i

n i

n P=P o Kj

(11)

Equilibrium constants (Kj) are useful to evaluate if a determined operational condition (T, P, composition) favors a certain reaction. This evaluation is made through the analysis of the equilibrium constant of a reaction compared to the constants of the other reactions considered in the system. 2.1.2. Lagrange multipliers method (LM) The system composition in thermodynamic equilibrium can also be found by solving N equilibrium equations (Eqs. (12) and (16)), one per molecule, w mass balance equations (Eq. (13)), one per element, and two restriction equations (Eqs. (14) and (15)), representing the mass balance for the gas phase and the global mass balance, respectively.

2 3   P4 1X X b ln f i yk yj 2dki dkj 5 B RT ii 2 j
k

(17)

DGfi RTln bi =fio f


X
i

X
k

lk aik 0 ; i 1; 2; .; N 1

(12)

In both methods, the variations of the standard enthalpies and Gibbs free energies of formation of the reaction system (DH j e DG j, respectively) are functions of temperature and must be corrected with the specic heat capacities of the gases, CP. This is a key parameter in equilibrium calculation and some effort must be paid in its estimation. The input parameters of the simulator are: mole fraction of the reactants, temperature and pressure in the reactor input and operational temperature and pressure. The solution of the equations is made through implemented computational codes in the open-source platform for numerical computation Scilab. Numeric calculations are made through the function fsolve, which uses a modication of the Powell hybrid method with a tolerance of 1010. 2.2. Autothermal reforming of methane reactor modeling

ni aik Mk ; k 1; 2; .; w

(13)
2.2.1. Chemical reaction In this study, ATR is dened as a combination of total oxidation and steam reforming of methane. In the reactor, several chemical reactions occur, and their rates are strongly dependent on the reactions conditions. In order to reduce the number of reactions and to keep the model as simple as possible, only the reactions with signicant reaction rates were considered. According to the literature Hoang et al. (2006), the reactions that prevail in the kinetics of autothermal reforming of methane are: steam reforming (Eq. (1)) (RRM,I), watergas shift (Eq. (2)) (RRM,II), the sum of these two previous reactions (Eq. (3)) (RRM,III) and total oxidation of methane (Eq. (6)) (RRM,IV), which is much more exothermic than the partial oxidation. Hence, the model takes into account four reactions and six gas species: methane, oxygen, carbon dioxide, water, carbon monoxide and hydrogen. Nitrogen is present in the inlet air and it is considered as an inert, which affects only the gas properties and the equilibrium conversions. 2.2.2. Reaction kinetic model There are a large number of kinetic models for steam reforming and watergas shift reactions in the literature. The model proposed by Xu and Froment for nickel catalysts is considered to be more general and it has been the most used in the literature Hoang et al. (2006). For methane total combustion, the rate equation proposed by Trimm and Lam (1980) was adopted in this work (See Eqs. (8)(24)).

X
i

ni nG ; i 1; 2; .; N 1

(14)

X
i

ni nT ; i 1; 2; .; N

(15)

Eq. (12) is valid only for species in gaseous state. Hence, it represents in this work the mass balance for CH4, H2O, CO, CO2, O2 and H2 (however, as will be shown, O2 is not considered in the SRM and DRM systems). Since the vapor pressure of solid carbon is practically zero for low temperatures and pressures, it has a tendency to precipitate. Besides, as the pressure effect on the activity coefcient of a solid is very small, an insignicant error is introduced by the hypothesis o that the fC =fC ratio is one. Hence, for solid carbon, Eq. (12) may be re-written as:

DGo;CS lCS 0 f

(16)

Calculating the thermodynamic equilibrium with the Lagrange multipliers method has the advantage of being independent of the explicit reactions taking place while the equilibrium constants method is very useful to describe the extent of the reactions over the entire range of operational conditions. Additionally, there are numerical differences between the two methods. This is because in Lagrange multipliers method, lagrangeans have no physical meaning. Therefore, it can be very difcult to provide a good guess for these variables, which can be very important in non-linear

kj ko exp Ej =RT j Kiad Kiad;o exp DHi =RT

(18) (19)

208

C.N. Avila-Neto et al. / Journal of Natural Gas Science and Engineering 1 (2009) 205215 Table 2 Vant Hoff parameters for adsorption of different species for ATR. Kiad;o (bar1)

  2 rRM;I kI =p2;5 pCH4 pH2 O p3 2 pCO =KIeq 1=U H H2 eq rRM;II kII =pH2 pCO pH2 O pH2 pCO2 =KII 1=U2   eq rRM;III kIII =p3;5 pCH4 p2 2 O p4 2 pCO2 =KIII 1=U2 H H H2 rRM;IV  kIV;a pCH4 pO2 1
C KCH4 pCH4 C KO2 pO2

(20) (21) (22)

DHi (m3

bar1 mol1) 0.38280 0.70650 0.82900 0.88680

KiC;o (bar1) 1.26 101 7.78 107

DHiC;o (m3

bar1 mol1) 0.27300 0.92800

CH4 CO H2 H2O CH4 (C) O2 (C)

6.65 8.23 6.12 1.77

10 105 109 105 (bar)

2  (23)

C Combustion. Data taken from the work of Halabi et al. (2008).

kIV;b pCH4 pO2 1


C KCH4 pCH4

C KO2 pO2

volume. In this condition, it is reasonable to adopt the following simplifying assumptions: i. interfacial mass transfer resistance and intra particle diffusion limitations can be neglected; ii. pressure drop is negligible; iii. the reactor can be considered as isothermal. With the assumptions adopted previously, the molar ow along the axial direction, for each component, Fi (mol/s), can be described by the following mass balance equation (Halabi et al., 2008):

U 1 KCO pCO KH2 pH2 KCH4 pCH4 KH2 O pH2 O =pH2

(24)

This kinetic model was developed for supported platinum catalysts and the corresponding adsorption parameters were adjusted for nickel. Ni catalyst was assumed to be in the reduced state, which implies that total combustion and reforming occur in parallel (Smet et al., 2001). Hence, the rate corresponding to the steam methane reforming reaction set (RRM,I, RRM,II and RRM,III) and to the total oxidation of methane (RRM,IV) are represented by Eqs. (20)(24). The equilibrium constants are calculated by means of the equations available in Table 1. To compute the reaction rate constants, the parameters of Arrhenius equation (Eq. (18)) are also available in Table 1, and the adsorption constants of Eq. (19), in Table 2. The rate of consumption/formation of each species in the gas phase is determined by summing up the reaction rates of each species in all reactions, as shown below in Eqs. (5)(30).

X dFi rij rb S dz j

(31)

rCH4 rRM;I rRM;III rRM;IV rH2 O rRM;I rRM;II 2rR;III 2rRM;IV rCO rRM;I rRM;II rCO2 rRM;II rRM;III rRM;IV rO2 2rRM;IV rH2 3rRM;I rRM;II 4rRM;III

(25) (26) (27) (28) (29) (30)

where i denotes the gas species; j represents the reaction index; z is the reactor length (0 to L); density of catalyst bed is rb, with a value of 1.87 106 g/m3; S (m2) is the reactor transversal area and rij are the reaction rates. The model is constituted by a set of ordinary differential equations (ODEs), non-linear, of initial value in length. The initial condition is given by Fi Fi0, and the inlet composition in molar ratio is 16.7% of CH4, 1.7% of O2, 41.6% of H2O and 40% of N2. The ODE system was integrated numerically using the function ode of the open-source software Scilab. 3. Results and discussion 3.1. Steam reforming of methane (SRM) According to Seo et al. (2002), the species in thermodynamic equilibrium for steam reforming of methane are CH4, H2O, CO, CO2, H2, C(s) and the radicals H, O, OH, HO2, HCO, CH and CH2. However, theirs simulations indicate that the concentrations of radicals can be neglected when compared to the concentrations of other products and thus, they were not considered in this work. Fig. 1a presents the yield of all species that constitute the reformate product of steam reforming in chemical equilibrium as a function of temperature with a pressure of 1 atm and a steam to carbon ratio (S/C) of 2. At 373 K, CH4 does not react and the species in chemical equilibrium are only CH4 and H2O in the proportion that they were fed to the system. As the temperature is incremented, the rst reaction to occur is methane decomposition (RSRM,III), generating C(s) and H2. The other two linearly independent reactions, steam reforming of methane (RSRM,I) and watergas shift (RSRM,II), start to occur around 600 K (Fig. 1b). It can be seen that a maximum is formed for reaction RSRM,III approximately at 823 K. This is the temperature in which the highest deposition of coke occurs, with a value around 0.7 mol. Above this temperature, while CH4 decomposition fades rapidly, steam reforming of methane and watergas shift reactions start to prevail. There is a maximum of CO2 formation (0.22 mol) around 973 K. However, as the reaction RSRM,II decreases smoothly as the temperature attends

2.2.3. Reactor model A one-dimensional model is proposed to represent a xed bed reactor, with nickel based catalyst, in small scale, operating in steady state condition. The reactor considered with 4 mm length is operated with 120 mL/min of inlet ow and 1 cm3 of catalyst

Table 1 Equilibrium constants and Arrhenius kinetic parameters for ATR. Reaction, j I II III IV,a IV,b Equilibrium constant, Kjeq KIeq exp26;830 30:114 T (bar2) eq eq eq KII KI KIII (bar2)
eq KIII exp4400 4:036 T

ko (mol1 kgcat x s1) j 1.17 1015(bar0.5) 2.83 10


14

Ej (m3 bar1 mol1) 2.401 2.439 0.6713 0.8600 0.8600

(bar

0.5

5.43 105 (bar1) 8.11 105 (bar2) 6.82 105 (bar2)

Data taken from the work of Halabi et al. (2008).

C.N. Avila-Neto et al. / Journal of Natural Gas Science and Engineering 1 (2009) 205215

209

a
Yield (mol)

4.0 3.2 2.4 1.6 CH4 0.8 C(s) H2

100 80 60 40 20 CO2 0

3.5 3.0
CH4 conversion (%)

H2

2.5

Yield (mol)

H2O

2.0 1.5 1.0 0.5 0.0 0.0 0.5 1.0 1.5 2.0

H2O C(s) CO CO2


2.5 3.0

CO

b
Reaction coordinate (mol)

0.0

0.8 0.6 0.4 0.2 0.0 373

RSRM,I

Steam to carbon ratio (mol/mol)


Fig. 3. SRM equilibrium yields as a function of S/C. (T 1273 K, P 1 atm).

RSRM,III RSRM,II

523

673 823 973 Temperature (K)

1123

1273

Fig. 1. SRM equilibrium results as a function of temperature. (a) Products yield. (b) Reaction coordinates. (P 1 atm, S/C 2).

its maximum value (1273 K), CO2 yield decreases only 0.08 mol from 973 to 1273 K. On the other hand, H2 yield is proportional to CH4 conversion, and it grows very fast in the temperature interval studied. From 1123 to 1273 K, H2 yield practically does not vary and it attends a value of 3.11 mol per mole of CH4 in the feed. Furthermore, H2/CO selectivity has a value of 3.8 at these conditions. To verify if the equilibrium results obtained by simulation are representative, kinetic and equilibrium CH4 conversion data are compared in Fig. 2. In this gure, two types of equilibrium curves can be observed: while the rst one (with C(s)) was obtained by considering the formation of solid carbon in the equilibrium products, the second one (without C(s)) was obtained without this consideration. From the equilibrium curves and experimental points, it is clearly observed that CH4 conversion increases when carbon is deposited. Although Schadel et al. (2009) did not conduct tests to verify coke deposition, one can suppose that, since the conversion points did not trespass de equilibrium barrier without C(s), coke has not been deposited in a large scale. The same is not

true for the data obtained by Rakass et al. (2006). The authors noted the presence of extensive coke deposits for all samples that have been exposed to the methane-rich fuel mixtures at temperatures from 50 to 70050  C. Consequently and as expected, the experimental points are comprised between the two equilibrium curves. In the absence of water, since CH4 is the only species in the system, it can only be decomposed in H2 and C(s). As shown in Fig. 3, it is possible to convert almost 100% of CH4 at the temperature of 1273 K and S/C 0 in 2 mol of H2 and 1 mol of C(s). Since reactions RSRM,I and RSRM,III produce more H2 than RSRM,II, adding any amount of water to the system always enhances the production of H2 and diminishes coke deposition. At the conditions mentioned above, the feed of 3 mol of H2O per mole of CH4 enables the production of 3.25 mol of H2 with a H2/CO selectivity equal to 4.4 and no coke deposition. On the other hand, CO2 production is increased, attending a value of 0.25 mol at 1273 K (Fig. 3). At the temperature of 1273 K, the pressure effect on H2 yield is negligible (see Fig. 4). It can be seen in the gure that, for temperatures lower than 1123 K, increasing the pressure tends to decrease H2 yield. Furthermore, this effect is more acute in the interval between 373 and 873 K. The fact is that the reaction occurs with an augmentation in the volume and thus, increasing the pressure shifts the equilibrium towards the reactants. As an example, raising the pressure from 0.1 to 1 atm, decreases H2 yield from 1.74 to 0.58 mol. Table A.1 shows a comparison between simulated results and literature data for the SRM reaction system Lutz et al. (2003) employed the software CHEMKIN with the program for chemical equilibrium calculations Stanjan, with S/C 2 and P 10 atm. The mean relative error between the results of this work (LM method) and those of Lutz et al. (2003) is 11.8%. Seo et al. (2002) used the

100

3.5

CH4 conversion (%)

80

3.0
60 40 20 0 373

Yield (mol)

With C(s)

Without C(s)

2.5 2.0 1.5 1.0 0.1 atm 1 atm 2 atm 3 atm 4 atm 5 atm 523 673 823 973 1123 1273

523

673

823

973

1123

1273

0.5 0.0 373

Temperature (K)
Fig. 2. CH4 conversion for SRM. (.) Thermodynamic equilibrium; (-) Unsupported Ni powder catalysts (Rakass et al. (2006)); (D) Rh-based monolithic honeycomb catalyst (Schadel et al. (2009)). (P 1 atm, S/C 2).

Temperature (K)
Fig. 4. H2 yield for SRM at various pressures as a function of temperature. (S/C 2).

210

C.N. Avila-Neto et al. / Journal of Natural Gas Science and Engineering 1 (2009) 205215

a
Yield (mol)

2.0 1.6 1.2 0.8 0.4 CH4 CO2 H2O C(s) H2 CO

100 95

100 80

CH4 conversion (%)

CH4 conversion (%)

With C(s)
60

90 85 80 75

Without C(s)
40 20 0 373

b
Reaction coordinate (mol)

1.8 0.0 1.4 1.0 0.6 0.2 -0.2 -0.6

373

523

673

823

973

1123

1273

Temperature (K)
DMR,III RDRM,III

523

673

823

973

1123

1273

Temperature (K)
R RDMR,I RDRM,I
Fig. 6. CH4 conversion for DRM. (.) Equilibrium; (-) 4% Rh/a-Al2O3 (Donazzi et al., 2008); (,) Ca0.2La0.8Ni0.3Al0.7O2.9 and (D) Sr0.8La0.2Ni0.3Al0.7O2.6 (Khalesi et al. (2008). (P 1 atm, CO2/CH4 feed ratio 1).

RDMR,II RDRM,II

-1.0 373

523

673 823 973 Temperature (K)

1123

1273

Fig. 5. DRM equilibrium results as a function of temperature. (a) Products yield. (b) Reaction coordinates. (P 1 atm, CO2/CH4 feed ratio 1).

software Aspen Plus with S/C 1 and P 1 bar. This time, the mean relative error reached a value of 128.9%. Nevertheless, with the purpose to analyze whether coke deposition inuences or not the results signicantly, the same analyses were conducted discarding the presence of coke in the products. Comparing the results of this work with those of Lutz et al. (2003), the errors were 4.2% (LM method) and 3.8% (EEC method), whereas for the results of Seo et al. (2002), the errors reached 30.5% (LM method) and 30.6% (EEC method). In this way, it can be supposed that neither of the authors considered the formation of coke in the products. As both studies presented their results by graphs, it was necessary to use the (Software ScanIt) to obtain the data. 3.2. Dry reforming of methane (DRM) The thermodynamic equilibrium analysis of dry reforming of methane is carried out considering the same species as in the case of steam reforming of methane. Fig. 5a presents the yield of all species that constitute the reformate product of DRM in chemical equilibrium as a function of temperature with a pressure of 1 atm and a carbon dioxide to methane ratio (CO2/CH4) of 1. Fig. 5b shows the reaction coordinates of the linearly independent reactions that represent the chemical equilibrium at the conditions mentioned above. At the temperature of 373 K, 87.5% of CH4 is converted by reactions RDRM,I and RDRM,III (Eqs. (4) and (7), respectively) in equal amounts of H2O and C(s). As the temperature is increased, the previous reactions fade gradually and the reverse of the watergas shift reaction (reaction RDRM,II) starts to raise. At about 973 K, the reverse of the watergas shift reaction shifts its direction towards the formation of CO2 and H2. At 1273 K, reaction RDRM,II is the main responsible for the production of H2, and the reformate is constituted of 0.15 mol of H2O, 1.67 mol of CO, 1.84 mol of H2 and 0.24 mol of C(s). Therefore, H2/CO selectivity at these conditions attends the value of 1.1, that is, 2.7 less then in the steam reforming system. As it can be seen by the experimental data taken from the work of Donazzi et al. (2008), the points at lower temperatures appear

much below the equilibrium barrier calculated considering formation of solid carbon (Fig. 6). This is probably due to the fact that, at the conditions employed to estimate the DRM equilibrium, the catalysts tested by the authors are not selective to coke. However, if formation of solid carbon is not considered in the equilibrium equations, since the formation of CO is not propitiated at lower temperatures and CH4 cannot decompose, CH4 does not react by means of reaction RDRM,I, and the equilibrium conversion curve appears little above the experimental points. As for the experimental results of Khalesi et al. (2008), all points appear between the two equilibrium curves, suggesting the formation of coke in the surface of the catalysts. Moreover, the authors indicate that the Ca-based perovskites deposit less coke than the Sr-based perovskites. One could suggest that, the closer the experimental points are to the equilibrium curve calculated with formation of solid carbon, more susceptible the catalyst would be to coke deposition. Indeed, analyzing Fig. 6, experimental data related to the Sr-based catalyst are closer to the equilibrium curve calculated with formation of coke. The effect of CO2/CH4 ratio in the feed for a temperature of 1273 K and pressure of 1 atm is shown in Fig. 7. The maximum yield of H2 is achieved when no CO2 is fed to the system, that is, for the condition in which only methane decomposition occurs. However, raising the CH4/CO2 feed ratio from 0 to 1 diminishes H2 yield in only 0.16 mol (or 8%). In doing so, while carbon deposition decreases from 1 to about 0.23 mol (or 77%), CO production is affected in an opposite way, augmenting its value from 0 to almost

3.0 2.5

Yield (mol)

2.0 1.5 1.0 0.5 0.0 0.0

H2

CO

CO2 C(s)
0.5 1.0 1.5 2.0 2.5

H2O
3.0

CO2/CH4 feed ratio (mol/mol)


Fig. 7. DRM equilibrium yields as a function of CO2/CH4 feed ratio. (T 1273 K, P 1 atm).

C.N. Avila-Neto et al. / Journal of Natural Gas Science and Engineering 1 (2009) 205215

211

2.0

2.0 1.6

100 90

Yield (mol)

H2 Yield (mol)
1.2 0.8 0.4

1.0

0.5

0.1 atm 1 atm 2 atm 3 atm 4 atm 5 atm 523 673 823 973 1123 1273

H2O C(s) CH4 CO

80 70 60

0.0 373

CO2
50

Fig. 8. H2 yield for DRM at various pressures as a function of temperature. (CO2/CH4 feed ratio 1).

Reaction coordinates (mol)

Temperature (K)

0.0
0.6 0.4 0.2 0.0 -0.2 -0.4 -0.6 373 523 673

RORM,IV RORM,III RORM,II

1.7 mol (or 170%). Furthermore, while 100% of CH4 is converted, CO2 conversion goes up to 90%. Concerning the pressure effect (Fig. 8), it seems to affect the yield of H2 in the same way than in steam reforming. At the temperature of 1273 K, pressure is almost negligible, but its effect is more acute in the interval from 523 to 1123 K. As in SRM, the DRM reaction system also occurs with an augmentation in the volume. Table A.2 shows the validation of the simulated results of this work by comparison with CH4 equilibrium conversion data of Akpan et al. (2007). The mean relative error between the results of this work (LM method) and those of Akpan et al. (2007) is 6.32%. A simulation considering the formation of coke was not included for this reaction, but it does not differ a lot from the results attained in the steam reforming equilibrium calculations. 3.3. Oxidative reforming of methane (ORM) For the oxidative reforming system, the following species were considered: CH4, H2O, CO, CO2, H2, O2 and C(s). Despite Zhu et al. (2001) consider the formation of others hydrocarbons such as C2H6, C2H4, C2H2, CH3OH, HCHO and HCOOH, the authors obtained a small yield for these species (<107%) and, for this reason, they were not considered in this work. As suggested by Mattos et al. (2002), two stages are involved in the ORM processes. In the rst stage, all oxygen is consumed leading to the formation of CO2 and H2O by total oxidation of methane (RORM,III) (Eq. (6)). Latter, in the second stage, synthesis gas is produced by steam (RORM,I) and dry (RORM,II) methane reforming reactions (Eqs. (1) and (4)). Thus, for the case where the oxygen to carbon ratio (O/C) is 0.5 and the operational pressure is 1 atm, the products yield and the reaction coordinates for oxidative reforming can be found in Fig. 9a,b, respectively. In all temperature range, the 0.5 mol of O2 fed is completely consumed by the total oxidation reaction. Synthesis gas production by reaction RORM,II also does not vary signicantly and produces around 0.25 mol of CO and H2 in all temperature range. At the temperature of 373 K, for each mole of H2O formed, 0.5 mol of coke are deposited by CH4 decomposition (RORM,IV). Nevertheless, at this temperature, CH4 conversion attends only 50%. Raising the temperature from 373 to 800 K increases coke deposition from 0.5 to 0.75 mol. Actually, this is the temperature where reaction RORM,IV produces the highest amount of carbon. Above 800 K, CH4 decomposition starts to fade since coke deposition starts to decrease rapidly. Above 673 K, CH4 consumption by the steam reforming of methane reaction (RORM,II) augments rapidly, attending the values of 0.8 and 1.9 mol for CO and H2 yields, respectively, at 1273 K. It is expected that, for higher

RORM,I
823 973 1123 1273

Temperature (K)
Fig. 9. ORM equilibrium results as a function of temperature. (a) Products yield. (b) Reaction coordinates. (P 1 atm, O/C 0.5).

temperatures, H2/CO selectivity attends the value of 2, justifying the complete consumption of CH4 by the partial oxidation reaction, which is a linear combination of reactions RORM,I, RORM,II and RORM,III. Fig. 10 shows some experimental results taken from literature together with the CH4 equilibrium curves considering and not considering formation of solid carbon. Despite the fact that all catalysts present good resistance to coke deposition, the data obtained by Dal Santo et al. (2008) and Lanza et al. (2008) surpassed the equilibrium barrier at intermediate temperatures. Nevertheless, these experimental values do not exceed more than 10% of the equilibrium data. On the other hand, CH4 conversion data taken from the work of Hong and Wang (2009) are below the equilibrium curve that did not consider coke deposition. The previous analysis suggested that the partial oxidation reaction is propitiated at higher temperatures. It is now necessary

100 80 60 40 20 0 573

CH4 conversion (%)

With C(s)

Without C(s)

723

873

1023

1173

Temperature (K)
Fig. 10. CH4 conversion for ORM. (.) Equilibrium; (-) Rh/Al2O3 (Dal Santo et al. (2008)); (,) Pt50Ru50 supported catalyst Lanza et al. (2008); (:) Ni catalyst and (D) Ni nanowire catalyst (Hong and Wang, 2009). (P 1 atm, O/C 0.5).

CH4 conversion (%)

1.5

212

C.N. Avila-Neto et al. / Journal of Natural Gas Science and Engineering 1 (2009) 205215

2.0

100

1.5

CH4 conversion (%)

H2 Yield (mol) H2O


1.0

80

With C(s)
60 40 20 0 473

Without C(s)

C(s)

CO

CO2

0.5

Reactor Model Equilibrium Experimental 573 673 773 873 973 1073

0.0 0.0

0.5

1.0

1.5

2.0

Oxygen to carbon ratio (mol/mol)


Fig. 11. ORM equilibrium yields as a function of O/C. (T 1273 K, P 1 atm).

Temperature (K)
Fig. 13. CH4 conversions for ATR as a function of temperature. Experimental data were obtained by Ayabe et al. (2003) during the heating process of ATR over a Ni/Al2O3 catalyst. (Reaction conditions: CH4, 16.7%; O2, 1.7%; H2O, 41.6%; N2 (balance); S/C 2.5; SV 7200 h1).

to stipulate the relationship between the yield of H2 and the oxygen to carbon ratio. Fig. 11 shows this relationship for a temperature of 1273 K and pressure of 1 atm. As expected, CH4 decomposition occurs at the beginning of the abscissa axis, generating H2 and C(s). Partial oxidation attends its highest occurrence at O/C 0.5, producing 0.8 mol of CO and 1.9 mol of H2. For higher O/C ratios, CH4 decomposition almost stops and steam reforming decreases proportionally to the augmentation in O/C. Therefore, H2 is mainly produced from dry reforming. At O/C 2, total oxidation dominates and generates 1 mol of CO2 and 2 mol of H2O. As to the pressure effect, the results for the H2 yield and coke deposition for O/C 0.5 can be seen in Fig. 12a,b, respectively. The curves of ORM resemble the curves of SRM. Increasing the pressure tends to decrease the yield of H2. However, the pressure effect is more pronounced in the interval from 373 to 873 K. Coke deposition is directly proportional to the pressure, but as the yield of H2, the effect is more acute for intermediate temperatures. At 673 K, while for 0.1 atm 1 mol of coke is deposited, for 5 atm the deposition attends only 0.3 mol.

To validate the simulations, the results were compared with the equilibrium composition data reported by Zhu et al. (2001) (see Table A.3). However, since the authors did not consider coke deposition, another program was constructed to generate the results discarding the presence of C(s), so it was possible to compare data of the same nature. Thus, the average relative error between the results of this work (LM method without coke) and those of Zhu et al. (2001) is 28.7%. 3.4. Autothermal reforming of methane (ATR) Ayabe et al. (2003) analyzed the activity of 10% Ni/Al2O3 catalyst for ATR reaction as a function of temperature (light-off curve). The authors observed that there is a large difference in methane conversions (i) when the reaction begins at low temperatures and the temperature is increased gradually, compared to the inverse procedure, (ii) starting the reaction at higher temperatures and gradually reducing it. The results show a strong hysteresis of conversion behavior in relation to temperature. Fig. 13 presents a comparison among simulated, experimental and thermodynamic methane conversions at the reactor exit (or in z L), with initial conditions given in Section 2.2.3. The experimental data for the chemical equilibrium shown in the gure are with respect to the cooling process and are in line with the simulated ones. However, it appears that, at lower temperatures, the reactor model, based on kinetic expressions, tend to follow the heating process (not shown), where the conversion is much lower when compared to the cooling process. It is also possible to observe in Fig. 13 that high methane conversions are favored at higher temperatures. The composition proles of ATR in chemical equilibrium (not considering coke deposition) and calculated by the reactor model are shown in Fig. 14a (light-off curve) for the same conditions mentioned previously. At lower temperatures, less than 10% of CH4 is converted by total oxidation, forming CO2 and H2O. Steam reforming and watergas shift increases in the same degree when augmenting the temperature until 850 K. At higher temperatures, CO2 mole fraction starts to decrease as a result of the reduction of the second reaction. When attending 1273 K, steam reforming occurs isolated and generates CO and H2. If coke deposition is considered, it attends a maximum of 1.5 mol at about 750 K and starts to decrease rapidly as the temperature is incremented. However, when compared to the oxidative reforming system, the amount of coke deposited at 1273 K is lower (see Figs. 9a and 14b). This is because feeding water to the system tends to shift the

a
H2 yield (mol)

2.0

1.5

1.0

0.5

0.1 atm 1 atm 2 atm 3 atm 4 atm 5 atm


0.1 atm 1 atm 2 atm 3 atm 4 atm 5 atm

b
Coke deposition (mol)

0.0
0.8 0.6 0.4 0.2 0.0 373

523

673

823

973

1123

1273

Temperature (K)
Fig. 12. H2 yield for ORM at various pressures as a function of temperature. (O/C 0.5).

C.N. Avila-Neto et al. / Journal of Natural Gas Science and Engineering 1 (2009) 205215

213

a
Mole fraction (%)

50 40 30 20 10

simulations. Thus, it can be assumed that the authors did not consider coke deposition.
H2

3.5. Methane reforming: an overview For all practical purposes, especially considering industrial production of hydrogen through methane reforming, the best criterion of process optimization is to minimize the energy used per kilogram of hydrogen produced. The cost of the energy used will be directly affected by process temperature and pressure. The shown results (e.g., Figs. 4, 8, and 12) are for the same temperature range (3731273 K) and the patterns of H2 yield are similar for all reforms along this temperature range. It can be seen also that high yields are obtained at pressures around 1 atm. Therefore, the effect of high or low pressure costs can be neglected as a rst approximation. Based on that, the ratio energy/(H2 produced), can be minimized through the maximization of H2 yield (mol) since the energy used will be approximately the same for equal temperatures. Summarizing, maximizing H2 yield can be used as a rst approximate criterion to compare all methane reforming. Based on that, it can be seen that steam methane reforming process presents the best results concerning hydrogen production, since it yields around 3 mol, compared to 1.75 mol in DRM and ORM.

CH4 O2

H2O CO CO2
2.0

b
Mole fraction (%)

0
40 30 20 10 0 473

C(s) H2O CO CO2


873 1073

1.2 0.8 0.4 0.0 1273

CH4
673

Temperature (K)
Fig. 14. Composition and coke deposition in chemical equilibrium and calculated by the reactor model for ATR. (a) Without coke deposition. (b) With coke deposition. () Reactor model. (.) Equilibrium. (Reaction conditions: CH4, 16.7%; O2, 1.7%; H2O, 41.6%; N2 (balance); S/C 2.5; SV 7200 h1).

Coke deposition (mol)

H2

1.6

4. Conclusions A fairly complete and systematic thermodynamics analysis was carried out for steam, dry, oxidative and autothermal reforming of methane. The equilibrium data taken from the literature were compared with the results obtained with the Lagranges multipliers method and the equilibrium constants evaluation method. Although the results shown in this manuscript were already mentioned in other publications, only few of them analyzed the deposition of coke and dened the operational range in which it happens with more probability. In Figs. 13, for example, we can see clearly that intermediate temperatures and low steam to carbon ratios in the feed favors the deposition of carbon particles. The analyses enabled to validate the simulated results of this work and to verify the best operational conditions for each methane reforming reaction. It was shown that the steam reforming of methane process generates the highest yield for H2 (3.36 mol per mol of CH4 fed) with full CH4 conversion and almost no coke deposition (0.05 mol). These values were attended with a temperature of 1120 K, atmospheric pressure and a steam to carbon ratio of 4. We also determined the linearly independent reactions that describe in a satisfactory way the reformate composition of each reaction system. With this analysis, it was possible to predict in which operational range each reaction prevails or not over the others. The proposed mathematical model for ATR has shown a good adjustment to the experimental data of Ayabe et al. (2003), presenting a good agreement for methane conversions and hydrogen yields over the whole range of temperature studied. The simulated data showed that the best reaction temperature to maximize the yield of hydrogen is comprised between 723 and 773 K, where more than 50% of methane is converted into products. Acknowledgements This work has been supported by Brazilian funding agencies, CAPES and CNPq (Grant # 475934/2006-7) as well as, Rede Bra sileira de Hidrogenio (CTPETRO/FINEP/PETROBRAS)

consumption of methane towards the steam reforming of methane reaction. Besides, the introduction of an inert (in this case N2) decreases the consumption of methane by the methane decomposition reaction. In Fig. 14 for some species the results based on the reactor model for ATR (solid lines) are slightly above chemical equilibrium (dotted lines), at low and at high temperatures. This can be explained by the fact that the reactor model whose kinetic expressions were taken from Trimm and Lam (1980) had its constants adjusted for temperatures between 773 K and 873 K. In our analysis, we have broadened this range interval in order to have a better understanding of each species behavior along temperature. Through the results of methane conversion and composition proles, it is possible to establish that, for a maximum in H2 production, associated with smaller operational costs, the best reaction conditions situate between 723 and 773 K. At these temperatures, it is possible to obtain a high yield of H2, an elevated H2/CO selectivity and a methane conversion around 50%. Besides, the catalyst properties are maintained with smaller chances of sintering. In addition, these conditions have the advantage of smaller energy costs, because the reaction temperature is low. Nevertheless, as demonstrated in Fig. 14b, one should pay attention to coke deposition, since it is thermodynamically propitiated at the temperature interval mentioned above. Accordingly to Fig. 12b, increasing the system pressure or the oxygen to carbon ratio could be a reasonable way to avoid this reaction. The validation of the equilibrium data is performed by comparison with the data reported by Ayabe et al. (2003), in the same operational conditions (Table A.4). The authors provide in graphics the conversion of CH4 and H2/(CO CO2) selectivity with four different types of feed. The average relative error was found to be less than 1%, not considering formation of coke in the

214

C.N. Avila-Neto et al. / Journal of Natural Gas Science and Engineering 1 (2009) 205215

Appendix

Table A.1 Comparison between simulated and literature equilibrium data for SRM. Operational conditions Compounds 773 K This work LM Lutz et al. (2003) CH4 H2O S/C 2 CO P 10 atm CO2 H2 Seo et al. (2002) S/C 1 P 1 bar CH4 H2O CO CO2 H2 0.267 0.537 0.002 0.037 0.157 0.320 0.249 0.019 0.071 0.341 EEC 0.260 0.524 0.002 0.041 0.173 0.317 0.245 0.019 0.072 0.347 0.260 0.524 0.004 0.038 0.174 Liter. 873 K This work LM 0.210 0.434 0.015 0.059 0.282 0.194 0.132 0.090 0.062 0.522 EEC 0.202 0.420 0.016 0.063 0.299 0.192 0.128 0.091 0.063 0.526 0.203 0.421 0.015 0.061 0.300 0.137 0.151 0.059 0.046 0.525 Liter. 973 K This work LM 0.133 0.322 0.055 0.065 0.425 0.079 0.054 0.185 0.025 0.657 EEC 0.126 0.309 0.056 0.068 0.441 0.078 0.053 0.185 0.025 0.659 0.126 0.314 0.061 0.065 0.434 0.055 0.064 0.155 0.023 0.652 Liter. 1073 K This work LM 0.058 0.230 0.114 0.051 0.547 0.027 0.019 0.229 0.007 0.718 EEC 0.054 0.222 0.115 0.053 0.556 0.026 0.019 0.229 0.007 0.719 0.050 0.222 0.115 0.050 0.563 0.018 0.018 0.224 0.004 0.716 Liter. 1173 K This work LM 0.015 0.185 0.154 0.036 0.610 0.009 0.007 0.243 0.002 0.729 EEC 0.013 0.181 0.154 0.038 0.614 0.009 0.007 0.243 0.002 0.739 0.015 0.184 0.153 0.038 0.61 0.004 0.004 0.242 0.000 0.739 Liter. 1273 K This work LM 0.003 0.175 0.169 0.029 0.624 0.004 0.003 0.247 0.001 0.745 EEC 0.003 0.173 0.168 0.030 0.626 0.004 0.003 0.247 0.001 0.745 0.004 0.176 0.168 0.027 0.625 0.000 0.000 0.246 0.000 0.748 Liter.

Table A.2 Comparison between simulated and literature equilibrium data for DRM. Operational conditions Conv. (%) 673 K This work LM Akpan et al. (2007) CH4:CO2:N2 2:2:1 P 1 atm CH4 3.96 EEC 3.95 3.38 Liter. 773 K This work LM 16.02 EEC 16.02 15.8 Liter. 873 K This work LM 43.01 EEC 43.10 45.1 Liter. 973 K This work LM 74.43 EEC 74.59 77.8 Liter. 1073 K This work LM 91.24 EEC 91.35 93.2 Liter.

Table A.3 Comparison between simulated and literature equilibrium data for ORM. Operational conditions Compounds O/C 0.5 This work LM Zhu et al. (2001) T 873 K P 1 atm CH4 H2O CO CO2 H2 0.214 0.107 0.155 0.107 0.417 0.156 0.078 0.200 0.078 0.488 Liter. O/C 1.0 This work LM 0.060 0.237 0.118 0.195 0.390 0.023 0.198 0.148 0.175 0.456 Liter. O/C 1.5 This work LM 0.005 0.409 0.062 0.270 0.254 0.000 0.401 0.063 0.268 0.268 Liter.

Table A.4 Comparison between simulated and literature equilibrium data for ATR. Operational conditions Ayabe et al. (2003) P 1 atm CH4 1 1 1 1 1 H2O 2.5 2.5 2.5 2.5 2.5 O2 0.1 0.1 0.3 0.5 1 N2 2.4 2.4 2.2 2 1.5 Conv. CH4 (%) Selectivity H2 CO CO2 Variable 573 K This work LM 11.53 2.26 1.10 0.69 0.30 11.8 2.27 1.11 0.67 0.32 Liter. 673 K This work LM 25.38 3.18 2.22 1.66 0.91 26.1 3.15 2.21 1.67 0.94 Liter. 773 K This work LM 49.92 3.44 2.84 2.38 1.55 50.1 3.40 2.83 2.37 1.55 Liter. 873 K This work LM 82.09 3.34 2.95 2.60 1.77 81.6 3.32 2.93 2.58 1.77 Liter. 973 K This work LM 98.13 3.20 2.88 2.55 1.72 97.2 3.20 2.88 2.54 1.73 Liter. 1073 K This work LM 99.85 3.13 2.80 2.47 1.65 99.7 3.11 2.79 2.46 1.65 Liter.

C.N. Avila-Neto et al. / Journal of Natural Gas Science and Engineering 1 (2009) 205215

215

References
al-Qahtani, H., 1997. Effect of ageing on a steam reforming catalyst. Chem. Eng. J. 66, 5156. Akpan, E., Sun, Y., Kumar, P., Ibrahim, H., Aboudheir, A., Idem, R., 2007. Kinetics, experimental and reactor modelling studies of the carbon dioxide reforming of methane (CDRM) over a new Ni/CeO2ZrO2 catalyst in a packed bed tubular reactor. Chem. Eng. Sci. 62, 40124024. Ayabe, S., Omoto, H., Utaka, T., Kikuchi, R., Sasaki, K., Teraoka, Y., Eguchi, K., 2003. Catalytic autothermal reforming of methane and propane over supported metal catalysts. Appl. Catal. A. 241, 261269. Benito, M., Garca, S., Ferreira-Aparicio, P., Garca Serrano, L., Daza, L., 2007. Development of biogas reforming NiLaAl catalysts for fuel cells. J. Power Sources 169, 177183. Corbo, P., Migliardini, F., 2007. Hydrogen production by catalytic partial oxidation of methane and propane on Ni and Pt catalysts. Int. J. Hydrogen Energy 32, 5566. Dal Santo, V., Mondelli, C., De Grandi, V., Gallo, A., Recchia, S., Sordelli, L., Psaro, R., 2008. Supported Rh catalysts for methane partial oxidation prepared by OMCVD of Rh(acac)(CO)2. Appl. Catal. A: Gen. 346, 126133. Dias, J.A.C., Assaf, J.M., 2004. Autothermal reforming of methane over Ni/gAl2O3 catalysts: the enhancement effect of small quantities of noble metals. J. Power Sources 130, 106110. Donazzi, A., Beretta, A., Groppi, G., Forzatti, P., 2008. Catalytic partial oxidation of methane over a 4% Rh/a-Al2O3 catalyst. Part II: role of CO2 reforming. J. Catal. 255, 259268. Goodwin, D., 2004. CANTERA: object-oriented software for reacting ows. Calif. Inst. Technol.. <http://www.cantera.org>. Halabi, M.H., Croon, M.H.J.M., Van Der Schaaf, J., Cobden, P.D., Schouten, J.C., 2008. Modeling and analysis of autothermal reforming of methane to hydrogen in a xed bed reformer. Chem. Eng. J. 137, 568578. Hoang, D.L., Chan, S.H., Ding, O.L., 2006. Hydrogen production for fuel cells by autothermal reforming of methane over sulde nickel catalyst on gamma alumina support. J. Power Sources 159, 12481257. Hong, X., Wang, Y., 2009. Partial oxidation of methane to syngas catalyzed by a nickel nanowire catalyst. J. Nat. Gas Chem. 18, 98103. INRIA ENPC. Scilab, On-line. <www.scilab.org>. Khalesi, A., Arandiyan, H.R., Parvari, M., 2008. Effects of lanthanum substitution by strontium and calcium in LaNiAl perovskite oxides in dry reforming of methane. Chin. J. Catal. 29 (10), 960968. Lanza, R., Canu, P., Jars, S.G., 2008. Partial oxidation of methane over PtRu bimetallic catalyst for syngas production. Appl. Catal. A. 348, 221228. Laosiripojana, N., Sutthisripok, W., Assabumrungrat, S., 2005. Synthesis gas production from dry reforming of methane over CeO2 doped Ni/Al2O3:

inuence of the doping ceria on the resistance toward carbon formation. Chem. Eng. J. 112, 1322. Li, B., Maruyama, K., Nurunnabi, M., Kunimori, K., Tomishige, K., 2004. Temperature proles of alumina-supported noble metal catalysts in autothermal reforming of methane. Appl. Catal. A. 275, 157172. Lutz, A.E., Bradshaw, R.W., Keller, J.O., Witmer, D.E., 2003. Thermodynamic analysis of hydrogen production by steam reforming. Int. J. Hydrogen Energy 28, 159167. Mattos, L.V., Oliveira, E.R., Resende, P.D., Noronha, F.B., Passos, F.B., 2002. Partial oxidation of methane on Pt/CeZrO2 catalysts. Catal. Today 77, 245256. Meng, L., Duan, Y.Y., Li, L., 2004. Correlations for second and third virial coefcients of pure uids. Fluid Phase Equilibria 226, 109120. Mukainakano, Y., Yoshida, K., Kado, S., Okumura, K., Kunimori, K., Tomishige, K., 2007. Catalytic performance and characterization of PtNi bimetallic catalysts for oxidative steam reforming of methane. Chem. Eng. Sci. 63, 48914901. OConnor, A.M., Schuurman, Y., Ross, J.R.H., Mirodatos, C., 2006. Transient studies of carbon dioxide reforming of methane over Pt/ZrO2 and Pt/Al2O3. Catal. Today 115, 191198. Pedernera, M.N., Pina, J., Borio, D.O., 2007. Kinetic evaluation of carbon formation in a membrane reactor for methane reforming. Chem. Eng. J. 134, 138144. Rakass, S., Oudghiri-Hassani, H., Rowntree, P., Abatzoglou, N., 2006. Steam reforming of methane over unsupported nickel catalysts. J. Power Sources 158, 485496. Rostrup-Nielsen, J.R., 2002. Syngas in perspective. Catal. Today 71, 243247. Schadel, B.T., Duisberg, M., Deutschmann, O., 2009. Steam reforming of methane, ethane, propane, butane, and natural gas over a rhodium-based catalyst. Catal. Today 142, 4251. Seo, Y.-S., Shirley, S.T., Kolaczkowski, S.T., 2002. Evaluation of thermodynamically favourable operating conditions for production of hydrogen in three different reforming technologies. J. Power Sources 108, 213225. Smet, C.R.H., Croon, M.H.J.M., Berger, R.J., Marin, G.B., Schouten, J.C., 2001. Design of adiabatic xed-bed reactors for the partial oxidation of methane to synthesis gas. Application to production of methanol and hydrogen-for-fuel-cells. Chem. Eng. Sci. 56, 48494861. ` Smith, J.M., Van Ness, H.C., Abbott, M.M., 2000. Introduao a termodinamica da engenharia qumica, fth ed. LTC, Rio de Janeiro. Software ScanIt. www.amsterchem.com/scanit.html>. Trimm, D.L., Lam, C.W., 1980. The combustion of methane on platinumalumina bre catalysts-I: kinetics and mechanism. Chem. Eng. Sci. 35, 14051413. Xu, J., Froment, G.F., 1989. Methane steam reforming, methanation and watergas shift: I. Intrinsic kinetics. AIChE J. 35, 8896. Yin, L., Wang, S., Lu, H., Ding, J., Mosto, R., Hao, Z., 2007. Simulation of effect of catalyst particle cluster on dry methane reforming in circulating uidized beds. Chem. Eng. J. 131, 123134. Zhu, J., Zhang, D., King, K.D., 2001. Reforming of CH4 by partial oxidation: thermodynamic and kinetic analyses. Fuel 80, 899905.

You might also like