i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 4 7 ( 2 0 2 2 ) 7 5 4 7 e7 5 6 8 Available online at www.sciencedirect.com ScienceDirect journal homepage: www.elsevier.com/locate/he Modelling the hydrodynamics and kinetics of methane decomposition in catalytic liquid metal bubble reactors for hydrogen production Lionel J.J. Catalan*, Ebrahim Rezaei Department of Chemical Engineering, Lakehead University, 955 Oliver Road, Thunder Bay, Ontario, P7J 5E1 Canada highlights graphical abstract Faster superficial gas velocity increases gas-liquid interfacial area and gas holdup. Larger diameter tubes lower the total melt volume in multitubular reactor designs. Liquid metal volume increases faster than exponentially with decreasing temperature. 14.8 m3 of Ni0$27Bi0.73 melt at 1050 C produces 10,000 Nm3.h1 of H2 at 80% CH4 conversion. Molten CueBi alloy results in smaller reactors (shorter tubes) than NieBi alloy. article info abstract Article history: Bubble reactors using molten metal alloys (e.g, nickel-bismuth and copper-bismuth) with Received 16 August 2021 strong catalytic activity for methane decomposition are an emerging technology to lower Received in revised form the carbon intensity of hydrogen production. Methane decomposition occurs non- 15 November 2021 catalytically inside the bubbles and catalytically at the gas-liquid interface. The reactor Accepted 9 December 2021 performance is therefore affected by the hydrodynamics of bubble flow in molten metal, Available online 11 January 2022 which determines the evolution of the bubble size distribution and of the gas holdup along the reactor height. A reactor model is first developed to rigorously account for the coupling Keywords: of hydrodynamics with catalytic and non-catalytic reaction kinetics. The model is then Hydrogen validated with previously reported experimental data on methane decomposition at Catalytic methane decomposition several temperatures in bubble columns containing a molten nickel-bismuth alloy. Next, Liquid metal the model is applied to optimize the design of multitubular catalytic bubble reactors at Bubble reactor industrial scales. This involves minimizing the total liquid metal volume for various tube Hydrodynamics diameters, melt temperatures, and percent methane conversions at a specified hydrogen Kinetics production rate. For example, an optimized reactor consisting of 891 tubes, each measuring 0.10 m in diameter and 2.11 m in height, filled with molten Ni0$27Bi0.73 at 1050 C and fed * Corresponding author. E-mail address: lionel.catalan@lakeheadu.ca (L.J.J. Catalan). https://doi.org/10.1016/j.ijhydene.2021.12.089 0360-3199/© 2021 Hydrogen Energy Publications LLC. Published by Elsevier Ltd. All rights reserved. 7548 i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 4 7 ( 2 0 2 2 ) 7 5 4 7 e7 5 6 8 with pure methane at 17.8 bar, may produce 10,000 Nm3.h1 of hydrogen with a methane conversion of 80% and a pressure drop of 1.6 bar. The tubes could be heated in a fired heater by burning either a fraction of the produced hydrogen, which would prevent CO2 generation, or other less expensive fuels. © 2021 Hydrogen Energy Publications LLC. Published by Elsevier Ltd. All rights reserved. Introduction Methane decomposition, also called methane pyrolysis (Eq. (1)), is an alternative to the traditional hydrogen production process. The latter combines steam methane reforming (SMR, Eq. (2)) with the water gas shift reaction (WGS, Eq. (3)). The main advantage of methane decomposition is the lack of direct CO2 emissions and the production of solid carbon, which may be sold or sequestrated. In contrast, the combination of SMR with WGS, (Eq. (4)), generates 1 mol of CO2 for every 4 mol of H2. A downside of methane decomposition is the low molar ratio of H2 production to CH4 consumption, which is only 2:1 compared to 4:1 for SMR-WGS. This results in a significant economic hurdle for methane decomposition that may be overcome with the ability to sell the carbon byproduct, government incentives, large taxes on CO2 emissions that would make SMR-WGS less profitable, or a combination of the above. The carbon by-product should not be used as a fuel since its combustion would produce CO2 and therefore negate the environmental benefit of methane decomposition. CH4 )/ 2H2 þ C(s) DH 298 ¼ 75 kJ mol1 (1) CH4 þ H2O )/ 3H2 þ CO DH 298 ¼ 206 kJ mol1 (2) CO þ H2O )/ H2 þ CO2 DH 298 ¼ -41 kJ mol1 (3) CH4 þ 2H2O )/ 4H2 þ CO2 DH 298 ¼ 165 kJ mol1 (4) Methane decomposition is one among many technological approaches to reduce CO2 emissions in H2 production processes [1]. Nonetheless, a recent life-cycle based study comparing the relative costs of CO2 mitigation for twelve different H2 production technologies using fossil fuels, nuclear energy, and renewable resources found that methane decomposition may be the most cost-effective abatement solution in the short term [2]. The formation of elemental carbon during methane decomposition creates a practical challenge for stable continuous reactor operation. When the reaction is carried out in the presence of solid catalyst pellets in a packed bed, the pressure drop increases rapidly and the catalyst quickly deactivates as carbon deposits on its surface [1,3]. On the other hand, non-catalytic methane decomposition in void tubular reactors causes buildup of hard carbon deposits on the reactor walls that eventually block the gas flow [4]. Fluidized bed reactors (FLBR) have been studied for their ability to mitigate these operational problems [5,6]. Qian et al. [7] found that an FLBR using 40 wt% Fe/Al2O3 catalyst maintained a low pressure drop and catalytic activity even though carbon was produced with a yield as high as 10.7 g carbon/g catalyst. Attrition of the produced carbon on the Fe catalyst surface in FLBRs can help maintain the conversion rate of methane over time [8]. Spent Fe/Al2O3 catalyst could be successfully regenerated multiple times by CO2 oxidation, and the regenerated catalyst not only required a reduced activation time but also improved the methane conversion [7]. Fe based catalysts are advantageous for methane decomposition because of their low cost, environmental friendliness, and easy activation. Recently, several industrial wastes or by-products containing iron were tested as catalysts for methane decomposition with encouraging results [9]. Liquid metal bubble reactors (LMBRs) provide an effective solution to the carbon deposition problem in laboratory studies. The gas injected at the bottom of the reactor rises as bubbles through the molten metal. Methane decomposition occurs inside the bubbles and at the gas-liquid interface if the molten metal has catalytic activity. Because carbon is much less dense than molten metal, it rises to the surface of the melt. In Catalan and Rezaei [10], we included a review of experimental studies of methane decomposition in LMBRs up to 2019. All studies concurred that most of the carbon formed a powder layer above the melt. Deposition of carbon on the reactor walls appears to be limited to very thin films. Upham et al. [11] observed that carbon deposited slowly on the walls of the reactor via precipitation from the saturated melt; however, the carbon-on-carbon deposition decreased with time such that the thickness of the deposited carbon layer remained less than 10 mm after one week of continuous operation. These observations are consistent with those of Geibler et al. [12] who reported that the carbon layer that had deposited on the reactor wall was only around 10 mm thick after 15 days of operation. Various methods have been envisioned in the literature for continuously removing the carbon floating on the surface of the melt. One method involves skimming the carbon mechanically from the surface, which is similar to the common practice for removing slag material from melts in metallurgical processes [11]. Alternatively, the fine powder of carbon could be entrained in the gas stream exiting the reactor and separated using cyclones, filters, or electrostatic precipitators [11,13]. To the authors’ knowledge, none of these methods has yet been demonstrated experimentally. Considering the high cost of molten metals, the carbon/metal separation should be very effective to limit the loss of metal and the economic burden of metal make-up. The contamination of carbon by metal may be substantially reduced by installing a molten salt layer floating on the liquid metal [14]. i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 4 7 ( 2 0 2 2 ) 7 5 4 7 e7 5 6 8 In the past few years, methane decomposition in LMBRs has been increasingly tested in catalytic molten metal alloys and salts such as Ni-Bi [11,14,15], Cu-Bi [16], Ga [17], MnCl2eKCl [18], NaCleKCleFe [19], and various alkali halide salts [20,21]. Additions of catalysts containing La, Ni, Co and Mn as particle suspensions in molten salts have also been investigated [22]. The main advantage of a catalytic melt is the potential to reduce the reactor size and to operate at lower temperatures ( < 1100 C) than with non-catalytic systems (e.g, molten Sn or Fe), while still achieving industrially relevant H2 production rates (several thousand Nm.3 h1) at relatively high CH4 conversions (>60%). Lower temperatures are desirable because they provide more flexibility for selecting materials of construction [23] and they reduce the costs of operation. Our previous work [10] centered on the design of noncatalytic LMBRs where methane decomposition occurred in the gas throughout the bubble but not specifically at the gasliquid interface. This required knowing the non-catalytic reaction rate in the gas phase, which in turn depends on reactant and product concentrations as well as on the gas volume in the melt. The gas volume per unit volume of melt (i.e., the gas holdup) is a complex function of the gas volumetric flow rate, pressure, temperature, composition, and of the liquid metal properties (density, viscosity, and surface tension). The gas holdup increases along the reactor height not only because the hydrostatic pressure decreases but also because each CH4 molecule that reacts produces two molecules of H2 (Eq. (1)). Hence, the reaction kinetics and hydrodynamics of LMBRs influence each other. We were the first to develop a coupled hydrodynamic and kinetic model of LMBRs for noncatalytic decomposition of methane [10]. Prior to this, other researchers had assumed an arbitrary gas holdup of 25% either in the entire melt [11,24] or at the melt bottom [13,25]. In our earlier study [10], we predicted the gas holdup using a drift flux correlation that had been experimentally corroborated in gas-molten metal mixtures [26,27], and we found that relatively large superficial gas velocities (>0.10 m.s1) were required to achieve gas holdups of about 10 vol percent or more. We developed thermodynamically consistent kinetic equations for non-catalytic methane decomposition based on the experimental data of Keipi et al. [28] obtained in a tubular reactor for the temperature range of 900e1450 K. Next, we applied our coupled hydrodynamic and kinetic model to optimize the diameter, height, and inlet pressure of a noncatalytic LMBR containing molten Sn by minimizing the volume of melt required to produce 200 kt.a1 of H2 at different temperatures and CH4 conversions. Since methane decomposition is endothermic (DHo298 ¼ 75 kJ/mol), heat needs to be provided to the molten metal, and we discussed various potential heating methods. Our LMBRs were designed with isothermal conditions since the turbulent flow of gas rising at high superficial velocities in the melt was assumed to cause sufficient mixing and liquid recirculation to equalize the temperature [10]. Unlike our earlier work, the present article focusses on the design of catalytic LMBRs. Kinetic data for methane decomposition in two catalytic molten alloys (Ni0$27Bi0.73 and Cu0$45Bi0.55) have been recently published [11,16]. Coupling kinetics and hydrodynamics in catalytic LMBRs poses 7549 significant challenges because the catalytic reaction occurs at the interface between the gas bubbles and the liquid metal; hence, the catalytic reaction rate is proportional to the interfacial area, which in turn depends on the size distribution of bubbles. Previous attempts to model catalytic LMBRs have made simple assumptions about bubble sizes. Upham et al. [11] assumed an average bubble diameter of 1 cm throughout the melt. In Farmer et al. [25], the bubbles were assumed to be 1.5 cm in diameter at the bottom of the melt and to change in size as they travelled upward due to the decrease in hydrostatic pressure, the stoichiometry of the methane decomposition reaction, and the diffusion of H2 across the gas-molten metal interfacea. Von Wald et al. [29] assumed that the bubble size at any height was proportional to its value at the bottom of the melt and to a factor accounting for the CH4 conversion and the stoichiometry of the reaction. The method to calculate the bubble radius at the bottom of the reactor was not reported. By contrast, the present article relies on a dimensionless correlation developed and experimentally corroborated by Akita and Yoshida [30] to calculate the specific gas-liquid interfacial area at any height in LMBRs as a function of the superficial gas velocity, the diameter of the column, and the properties (surface tension, viscosity and density) of the molten metal. Changes in CH4 conversion and hydrostatic pressure with height are reflected in the superficial gas velocity. This approach leads to a coupled hydrodynamic and kinetic model of catalytic LMBRs that does not require making any assumption about the bubble sizes. The rest of this article has the following organization. The Theory section (Section: Theory) derives the material balance and pressure equations for the coupled model. It includes subsections on the kinetics of catalytic and non-catalytic methane decomposition in molten metals (Subsections: Kinetics of catalytic methane decomposition in molten metalseExpansion of the differential material balance equation), the specific gas-liquid interfacial area (Subsection: Specific gas-liquid interfacial area), the gas holdup (Subsection: Gas holdup in molten metals), and the reactive surface area (Subsection: Reactive surface area in catalytic molten metals). Next, the Applications section (Section: Applications) starts by validating the coupled model with previously reported experimental data obtained in laboratory-sized catalytic LMBRs (Subsection: Experimental validation of the catalytic liquid metal bubble reactor model). It then proceeds to scaleup of an industrially-sized multitubular reactor (Subsection: Reactor scale-up) and to design optimization by minimizing the melt volume for a given hydrogen production rate (Subsection: Reactor design optimization). The Discussion section (Section: Discussion) begins by examining the effect of practical considerations (e.g. limits on maximum pressure) on designs (Subsection: Variations from optimum designs). Next, the designs of the present article are compared to previous a Farmer et al. [25] modelled an LMBR surrounded by a membrane that was selectively permeable to H2. This created the conditions for net diffusion of H2 from the gas inside the bubbles toward the molten metal. In the absence of such a membrane, the net diffusion of H2 through the gas-liquid interface is nil when the reactor operates at steady state. 7550 i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 4 7 ( 2 0 2 2 ) 7 5 4 7 e7 5 6 8 catalytic LMBR designs from literature (Subsection: Comparison to previous catalytic LMBR design from the literature) and to steam methane reformer designs (Subsection: Comparison to steam methane reformer). In summary, this article differs from our previous work [10] in that it accounts for catalytic methane decomposition at the gas-molten metal interface. Furthermore, it improves on previous attempts to model catalytic methane decomposition in LMBRs [11,25,29] by developing a new model that calculates the interfacial area and gas holdup as a function of the LMBR operating conditions, so that methane conversion is successfully predicted without making ad hoc assumptions about the bubble size. Theory Methane decomposition in catalytic LMBRs is determined by the interplay between reaction kinetics and the hydrodynamics of gas rising in the melt. Because the catalytic reaction occurs at the gas-liquid interface, the reactor model must be able to predict the interfacial area. The model must also calculate the gas holdup to account for non-catalytic methane decomposition inside the gas bubbles. Moreover, both the interfacial area and the gas holdup depend on methane conversion because the reaction generates 2 mol of hydrogen for each mole of methane consumed (Eq. (1)). As gas rises in the reactor, the increasing number of gas moles and the decreasing hydrostatic pressure affect the gas holdup and the size distribution of bubbles. In this section, we develop a coupled hydrodynamic and kinetic model that will form the basis for designing catalytic LMBRs that can achieve desired H2 production rates and CH4 conversions. Fig. 1 shows a simplified diagram of an LMBR. The material balance equation for the gas phase in a differential volume dV (m3) consisting of a horizontal slice of the melt with length dL (m) is: n_CH4 ;b dXCH4 ¼ ðRc þ Rn Þ a dV ¼ ðRc þ Rn Þ a pD2 dL 4 (5) where Rc and Rn , both having units of mol CH4.m3.s1, are respectively the catalytic and non-catalytic reaction rates per unit volume of gas, a is the gas holdup (m3 gas.m3 reactor), D is the reactor inner diameter (m), n_CH4 ;b is the inlet mole flow rate (mol.s1) of methane at the bottom of the reactor, and dXCH4 is the change in methane conversion (dimensionless) occurring in the differential volume.b The pressure change (Pa) of the gas as it rises through the differential volume element is: dP ¼ rl ð1 aÞ þ rg a g dL ¼ ðrl aDrÞg dL (6) where rl is the liquid metal density (kg.m3), rg is the gas density (kg.m3), Dr ¼ rl rg , and g is the gravitational acceleration. Eqs. (5) and (6) constitute a system of coupled differential equations that must be solved numerically between the bottom (L ¼ 0, XCH4 ¼ 0, P ¼ Pb ) and the top of the melt (L ¼ Lt ) to find the methane conversion and the pressure at the outlet of the reactor. This requires relating the reaction rates Rc and Rn , as well as the gas holdup a, to XCH4 and P. These relationships are developed below. Kinetics of catalytic methane decomposition in molten metals The catalytic decomposition of methane is assumed to be proportional to the interfacial area between the gas and the liquid [11,25]. Hence, the catalytic reaction rate per unit volume of gas, Rc , is related to the catalytic reaction rate per unit of interfacial area, (mol CH4.m2.s1), and the specific interfacial area, (m2 (interfacial area).m3 (gas)), by: Rc ¼ ag rc (7) The value of ag depends on the bubble size distribution, which is related to the hydrodynamics of the rising gas bubbles in the melt. The calculation of ag is addressed in Section: Specific gas-liquid interfacial area. The rest of the present section deals with rc . The experiments of Upham et al. [11] and Palmer et al. [16] in Ni0$27Bi0.73 and Cu0$45Bi0.55 melts, respectively, have shown that the forward catalytic rate, rc;f (mol.m2.s1), is first-order with respect to CH4: rc;f ¼ kc;f CCH4 (8) where kc;f is the forward rate coefficient of the catalytic reaction (m.s1) and CCH4 is the concentration of methane in the b Fig. 1 e Simplified graphical representation of an LMBR. For convenience, all symbols and units are defined in the Nomenclature section at the end of the article. 7551 i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 4 7 ( 2 0 2 2 ) 7 5 4 7 e7 5 6 8 gas (mol.m3). The net catalytic reaction rate is the difference between rc;f and the backward catalytic rate, rc;b (mol.m2.s1): kc;f CH2 2 CH2 2 ¼ kc;f CCH4 1 rc ¼ rc;f rc;b ¼ kc;f CCH4 KC CCH4 $KC (9) where CH2 is the concentration of hydrogen in the gas (mol.m3) and Kc is the concentration-based equilibrium constant (mol.m3). The backward reaction term in Eq. (9) ensures that rc tends toward zero when the methane decomposition reaction approaches thermodynamic equilibrium [31]. From Le Chatelier's principle, the CH4 conversion at equilibrium increases at higher temperature because the reaction is endothermic and decreases at lower pressure because methane decomposition causes a net increase in the number of gas moles (Eq. (1)). The Arrhenius equation for kc;f is: kc;f ¼ koc;f exp Eac;f ! (10) RT 91204:6 K ¼ exp 13:2714 RT (12) Kinetics of non-catalytic methane decomposition Non-catalytic methane decomposition occurs in the gas phase inside the bubbles. The net non-catalytic reaction rate ; Rn (mol CH4.m3.s1), per unit volume of gas is given by Ref. [10]: Rn ¼Rn;f Rn;b ¼kn;f CCH4 n kn;f CH2 2 CH2 2 CCH4 n1 ¼kn;f CCH4 n 1 KC CCH4 $KC (13) where Rn;f and Rn;b are respectively the forward and backward non-catalytic reaction rates, kn;f is the forward rate constant (m3(n1).mol(1n).s1), and n is the order of the forward reaction with respect to methane. The Arrhenius equation for kn;f is: Ean;f ! The experimental values of the forward pre-exponential factor, koc;f (m.s1), and forward activation energy, Eac;f kn;f ¼ kon;f exp (J.mol1), are given in Table 1 for Ni0$27Bi0.73 and Cu0$45Bi0.55 melts [11,16]. Experimental results reported in Parkinson et al. [20] indicate that catalytic methane decomposition is very likely reaction limited, not mass transfer limited, in the range of relevant ag values. Hence, the net catalytic reaction rate (Eq. (9)) does not need to be corrected by an effectiveness factor that would otherwise account for mass transfer limitation. This is explained in more detail in Section: 1. Evidence of reaction limited kinetics at the surface of bubbles in the Supplementary Material. The concentration-based equilibrium constant KC is related to K, the unitless reaction equilibrium constant, by Ref. [31]: Table 2 lists the numerical values of kon;f , Ean;f and n determined in Catalan and Rezaei [10] from experimental data reported in Keipi et al. [28]. o Pni P iðgÞ KC ¼ K RT (11) where Po is the standard pressure (105 Pa), R is the universal gas constant (8.314 m3.Pa.K1.mol1), T is the temperature (K), P and ni is the sum of the stoichiometric coefficients of the Expansion of the differential material balance equation Substituting the equations for the catalytic and non-catalytic rates of methane decomposition, Eqs. (7), (9) and (13), in the differential material balance equation, Eq. (5), yields: CH2 2 CH2 2 þ kn;f CCH4 n 1 ag kc;f CCH4 1 CCH4 $KC CCH4 $KC pD2 a dL ¼ n_CH4 ;b dXCH4 4 iðgÞ 900e1200 C, K is approximated very accurately by Ref. [10]: Table 1 e Kinetic parameters for catalytic methane decomposition in Ni0·27Bi0.73 and Cu0·45Bi0.55 from Refs. [11,16], respectively. (15) The concentrations of CH4 and H2 in the gas phase are related to the methane conversion, XCH4 , as follows [31]: CCH4 ¼ iðgÞ gas components (CH4 and H2) taking part in methane P decomposition: ni ¼ 2 1 ¼ 1. In the temperature range of (14) RT CH2 ¼ Tb n_CH4 ;b ð1 XCH4 Þ P T V_ g;b ð1 þ εXCH4 Þ Pb (16) n_CH4 ;b qH2 þ 2XCH4 P Tb Pb T V_ g;b ð1 þ εXCH4 Þ (17) where Pb and Tb are the pressure (Pa) and temperature (K) at the bottom of the melt, ε is the mole fraction of CH4 in the reactor feed, qH2 is the mole ratio of H2 to CH4 in the reactor feed, and V_ g;b is the volumetric flow rate of gas at the bottom of the melt (m3.s1). With the ideal gas approximation, V_ g;b is given by: Parameter Units Value in Ni0$27Bi0.73 Value in Cu0$45Bi0.55 koc;f Eac;f a m.s1 J.mol1 7.88 104 208,000 4.32 105a 222,000 Palmer et al. [16] calculated ko ¼ 3.7 108 ± 2.0 108 s1 in Rc ¼ ko expðEac;f =RTÞ CCH4 by assuming a bubble diameter Table 2 e Kinetic parameters for non-catalytic methane decomposition. db ¼ 0.007 m. They kept the methane conversion below 10% to minimize the backward reaction in their experiments. Therefore, koc;f is related to their pre-exponential factor ko by koc;f ¼ ko = ag Parameter with ag ¼ 6=db : n kon;f Ean;f Units 3(n1) m .mol J.mol1 unitless Value (1n) .s 1 1.4676 1010 284,948 1.0809 7552 i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 4 7 ( 2 0 2 2 ) 7 5 4 7 e7 5 6 8 n_CH4 ;b þ n_H2 ;b þ n_I;b RTb V_ g;b ¼ Pb (18) In Eq. (18), n_H2 ;b and n_I;b are the inlet flow rates (mol.s1) of H2 and inert components (e.g., argon) at the bottom of the reactor. Substitution of Eq. (18) in Eqs. (16) and (17) gives: P n_CH4 ;b ð1 XCH4 Þ CCH4 ¼ n_CH4 ;b þ n_H2 ;b þ n_I;b ð1 þ εXCH4 Þ RT CH2 ¼ n_CH4 ;b qH2 þ 2XCH4 P n_CH4 ;b þ n_H2 ;b þ n_I;b ð1 þ εXCH4 Þ RT (19) (20) Substituting Eqs. (19) and (20) in Eq. (15) and further rearrangement yields the expanded differential material balance equation of the LMBR: ( ag kc;f ð1 XCH4 Þ P n_CH4 ;b þ n_H2 ;b þ n_I;b ð1 þ εXCH4 Þ RT #n ) ð1 XCH4 Þ P þ kn;f , n_CH4 ;b þ n_H2 ;b þ n_I;b ð1 þ εXCH4 Þ RT ( ) 2 n_CH4 ;b qH2 þ 2XCH4 P 1 n_CH4 ;b þ n_H2 ;b þ n_I;b ð1 þ εXCH4 Þð1 XCH4 Þ,KC RT dXCH4 a pD2 ¼ dL 4 " (21) Before Eq. (21) can be solved, the terms ag and a must be written in terms of XCH4 and P. This is done in the next two sections. Specific gas-liquid interfacial area For a distribution of bubble sizes, the specific interfacial area ag (m2 (interfacial area).m3 (gas)) is related to the volume- Nevertheless, these measurement techniques and correlations for initial bubbles sizes (as they leave the sparger orifice or nozzle) are not relevant for measuring or predicting bubble sizes in the main part of bubble columns. This is because bubbles interact with each other as they rise, especially at conditions of high superficial gas velocities relevant to industrial operation. Consequently, the size of the majority of bubbles throughout most of the column depends mostly on the balance between the rates of bubble coalescence and breakup. These rates, in turn, depend on the superficial gas velocity and the liquid properties [30]. To the authors’ knowledge, no published study has reported bubble size measurements in molten metal at conditions of high superficial gas velocities in regions where bubble coalescence and break up are taking place. The present study uses the experimentally-based correlation of Akita and Yoshida [30] to predict dvs in LMBRs (Eq. (23)). These researchers used a photographic technique to measure bubble sizes in square columns having cross sections of 7.7 cm 7.7 cm, 15 cm 15 cm, and 30 cm 30 cm. The experiments consisted of injecting air or oxygen in water, 30e100% glycol solutions, methanol, or carbon tetrachloride. The bubble sizes were measured at a height of 150 cm from the bottom of the column in the region of bubble coalescence and breakup. They were found to be independent of the device used to inject the gas in the column (perforated plate, porous gas plate, or single orifice gas sparger). dvs ¼ 26ðNBo Þ0:50 ðNGa Þ0:12 ðNFr Þ0:12 D (23) The dimensionless Bond number (NBo ), Galilei number (NGa ), and Froude number (NFr ) are defined as follows: surface mean bubble diameter dvs (m), also called the Sauter mean diameter, by: NBo ≡ gD2 rl s (24) 6 ag ¼ dvs NGa ≡ gD3 n2l (25) (22) In other words, dvs is the diameter of a bubble having the same volume/surface area ratio as the entire distribution. Optical and photographic techniques are not applicable to the measurement of bubble sizes in opaque liquids such as molten metals. Techniques based on differential pressure [32] and acoustical [33,34] measurements have been reported for determining the size of bubbles as they detach from a submerged orifice in various liquid metals (tin, lead, copper, mercury, and silver). Sun et al. [32] found that helium bubble diameters experimentally measured in molten tin at temperatures between 400 and 700 C were in relatively good agreement (average deviations of only 4.97%) with the empirical correlation of Jamialahmadi et al. [35]. Interestingly, this correlation was based on experiments carried out with water, wateralcohol, and waterglycerol mixtures. This indicates that under the circumstances of the experiments carried out by Sun et al. [32], the detaching bubble sizes in molten metal could be acceptably predicted by correlations developed with fluids having different properties (particularly much lower density and surface tension). NFr ≡ jg ðgDÞ1=2 (26) where nl is the liquid kinematic viscosity (m2.s1), s is the surface tension (N.m1), and jg is the superficial gas velocity (m.s1). Akita and Yoshida [30] found that their correlation was in agreement with experimental data from an earlier study [36] where wider variations of system properties, column size, and superficial gas velocity were tested. Combining the results of their two articles, the dvs correlation was experimentally validated for the ranges NBo; exp ¼ 7.98 102 e 4.85 104, NGa; exp ¼ 6.25 106 e 1.79 1012, and NFr ¼ 8 104 e 1.35 101. The superficial gas velocity is given by: jg ¼ 4V_ g pD2 (27) where V_ g is the actual volume flow rate of the gas (m3.s1), which depends on the local pressure and methane conversion in the reactor: i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 4 7 ( 2 0 2 2 ) 7 5 4 7 e7 5 6 8 Pb T V_ g ¼ V_ g;b ð1 þ εXCH4 Þ Tb P (28) Substituting V_ g;b from Eq. (18) in Eq. (28) gives: RT V_ g ¼ n_CH4 ;b þ n_H2 ;b þ n_I;b ð1 þ εXCH4 Þ P (29) Next, Eq. (29) is substituted in Eq. (27) to yield: jg ¼ 4 n_CH4 ;b þ n_H2 ;b þ n_I;b ð1 þ εXCH4 Þ RT 2 P pD (30) Eq. (30), in combination with Eqs. (22)e(26), is used to relate ag to XCH4 and P. The dvs correlation of Eq. (23) applies to the parts of the column where bubble sizes are determined by the balance between the rates of bubble coalescence and breakup. This correlation, however, does not apply to the initial bubbles in the first few centimeters above the injector, where the local volume-surface mean bubble diameter d*vs (m) depends on the orifice diameter do (m) and the gas velocity through the orifice uo (m.s1). For initial bubbles, Akita and Yoshida [30] found the following correlation: d*vs uo ¼ 1:88 pffiffiffiffiffiffiffiffi do gdo !1=3 (31) Andreini et al. [33] also found that the size of the initial bubbles in liquid tin, lead, and copper depended on do and uo . Nevertheless, for reactor design, which is the focus of the present article, Eq. (23) applies unless the reactor is very short. Predicting dvs requires estimating liquid metal properties (density, viscosity, surface tension, and liquidus temperature). There is only a very limited amount of experimental data for the physical properties of NieBi and CueBi alloys in the 7553 literature. Therefore, several models were utilized to estimate the properties of these alloys as a function of temperature from the properties of their constituent pure liquid metals. This is explained in detail in Section: 2. Estimation of properties of liquid metal alloys in the Supplementary Material. Fig. 2A shows dvs predicted by Eq. (23) as a function of jg for a base case consisting of liquid Ni0$27Bi0.73 in a 0.18-m diameter column at 1000 C, and for variations on this base case. Increasing jg causes dvs to progressively decrease. Reducing the column diameter by a factor of 3 down to 0.06 m results in dvs increasing by approximately 40%. Raising the temperature to 1200 C has a minimal effect on dvs . Changing the liquid alloy to Cu0$45Bi0.55 increases dvs by approximately 11%. For simplicity, dvs will be called the mean bubble diameter from now on in this article. It is important to remember, however, that dvs is a measure of the specific gas-liquid interfacial area through Eq. (22) rather than an actual bubble diameter. Gas holdup in molten metals Kataoka and Ishii [26] developed a correlation for gas holdup that was experimentally corroborated in molten metal baths [27]. If the liquid metal is not flowing in and out of the reactor, a is given by: a¼ jþ jg g ¼ þ þ C0 jg þ Vgj C0 jg þ Vgj (32) where Vgj is the void-fraction weighted mean drift velocity (m.s1) and C0 is the distribution parameter which depends on the column geometry. For a round column, qffiffiffiffiffiffiffiffiffiffiffi C0 ¼ 1:2 0:2 rg rl (33) Fig. 2 e Effects of superficial gas velocity, reactor diameter, temperature, alloy type, pressure, and gas composition on (A) the volume-surface mean bubble diameter and (B) the reactive surface area per unit volume of reactor. The base case corresponds to D ¼ 0.18 m, T ¼ 1000 C, Alloy ¼ Ni0·27Bi0.73, P ¼ 15 bar, and Gas ¼ CH4. The legend indicates the parameters that change from the base case. 7554 i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 4 7 ( 2 0 2 2 ) 7 5 4 7 e7 5 6 8 The dimensionless superficial gas velocity, jþ g , and the þ , are defined by: dimensionless drift velocity, Vgj jþ g ≡ jg þ ≡ Vgj (34) 1=4 sgDr r2l Vgj sgDr r2l (35) 1=4 þ depends on the viscosity The method for evaluating Vgj number, Nm , and the dimensionless hydraulic equivalent diameter of the vessel, D*H , which are defined as follows: Nm ≡ ml pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi rl s s=ðgDrÞ (36) 1=2 Cu0$45Bi0.55, this situation occurs when the column diameter is larger than 65 mm, the temperature is above 800 C, the pressure ranges from 1 to 80 bar, and the gas is a mixture of CH4 and H2. In Catalan and Rezaei [10], we reported graphically how the superficial gas velocity, pressure, temperature, and phase compositions affect the gas holdup in a molten metal bath. In order to use the above equations to calculate a in LMBRs, jg and rg must be expressed as functions of XCH4 and P. This has already been done for jg in Eq. (30). The equation relating rg to XCH4 and P is developed next. In accordance with the ideal gas approximationc, the gas density is given by: rg ¼ D D*H ≡ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi s=ðgDrÞ (37) PMg RT (41) where Mg (kg.mol1) is the molar mass of the gas: CH2 MH2 þ CCH4 MCH4 þ CI MI CH2 þ CCH4 þ CI where ml is the liquid dynamic viscosity (Pa.s). When jþ g [0:5, a significant amount of liquid recirculation Mg ¼ þ occurs in the melt, and Vgj is given by either Eqs. (38a), (38b), or In Eq. (42), MH2 , MCH4 and MI are respectively the molar masses of H2 (0.002016 kg.mol1), CH4 (0.01604 kg.mol1), and an inert component (if present). The inert component concentration is given by Ref. [31]: (38c): þ ¼ 0:0019D*H Vgj 0:809 rg rl 0:157 Nm 0:562 for Nm 2:2 103 and D*H 30 (38a) þ ¼ 0:030 rg rl Vgj þ ¼ 0:92 rg rl Vgj 0:157 0:157 (43) (38b) (38c) Mg ¼ Conversely, when jþ g ≪0:5, liquid recirculation is less and the experimental data are adequately predicted by the drift flux correlations of Ishii [37] for bubbly and churn-turbulent flow: pffiffiffi þ ¼ 2ð1 aÞ1:75 for bubbly flow Vgj (39a) pffiffiffi þ Vgj ¼ 2 for churn turbulent flow (39b) qH2 þ 2XCH4 MH2 þ ð1 XCH4 ÞMCH4 þ qI MI 1 þ qH2 þ XCH4 þ qI (44) Then, substituting Eq. (44) in Eq. (41) provides the desired relationship between rg ; XCH4 , and P: rg ¼ qH2 þ 2XCH4 MH2 þ ð1 XCH4 ÞMCH4 þ qI MI P RT 1 þ qH2 þ XCH4 þ qI (45) Reactive surface area in catalytic molten metals Eqs. (38) and (39) can be combined to account for the whole range of jþ g as follows [27]: o n þ þ þ þ Vgj ¼ Vgj ðEq:39Þexp Ajþ ðEq:38Þ 1 exp Ajþ Vgj g g P n_CH4 ;b qI n_CH4 ;b þ n_H2 ;b þ n_I;b Rð1 þ εXCH4 Þ RT where qI is the mole ratio of inert component to methane in the reactor feed. Substitution of Eqs. (19), (20) and (43) in Eq. (42) gives: Nm 0:562 for Nm 2:2 103 and D*H 30 for Nm 2:2 103 and D*H 30 CI ¼ (42) (40) þ with A ¼ 1.39 so that expðAjþ g Þ ¼ 0.5 at jg ¼ 0.5. Hibiki et al. [27] used neutron radiography to visualize nitrogen bubbles and measure the gas holdup in a PbeBi melt contained in a rectangular tank for a range of nitrogen flow rates and liquid metal heights. Their measurements of gas holdups in the N2ePbeBi system and those of Saito et al. [38] in the N2eGa system were adequately predicted by Eq. (40). The relative error on gas holdup was approximately ±30% for 102 jþ g 2 and 0 a 0:32. The reactive surface area per unit volume of reactor is the product ag a, which has units of m2 (interfacial area).m3 (reactor). Fig. 2B shows that ag a increases considerably with jg at low values of jg and more moderately at larger jg values. Reducing the column diameter or pressure and increasing the temperature cause ag a to decrease. Changing the gas from CH4 to H2 decreases ag a due to the lower density of H2. The reactive surface area of Cu0$45Bi0.55 is lower than that of Ni0$27Bi0.73 by approximately 15% when used in a bubble column under same operating conditions. The product ag a depends on longitudinal position along the reactor because both ag and a depend on L. The reactive surface area for the entire column, Ar (m2), is the integral of ag a from the bottom to the top of the melt: þ It is noteworthy that Vgj becomes independent of the col- umn diameter when D*H 30 (Eqs. (38b) and (38c)) because the gas flow is no longer affected by the vessel walls except in close proximity to the walls. For molten Ni0$27Bi0.73 or c The error introduced by the ideal gas approximation is inconsequential compared to the relative error of ±30% on gas holdup predictions. i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 4 7 ( 2 0 2 2 ) 7 5 4 7 e7 5 6 8 Ar ¼ ð Vt 0 pD2 ag adV ¼ 4 ð Lt 0 6 a dL dvs (46) Applications The system of Eqs. (5) and (6) was numerically solved in MATLAB 2019b using the function ode23s with given reactor dimensions and temperature, as well as inlet feed flowrate and composition, to obtain the profiles of CH4 conversion, gas holdup, pressure, superficial gas velocity, and mean bubble diameter along the height of the reactor. Reactive surface areas (Eq. (46)) were calculated using the trapezoidal numerical integration method in MATALB 2019b. Experimental validation of the catalytic liquid metal bubble reactor model Upham et al. [11] reported methane decomposition experiments in a 0.03-m diameter bubble column filled with molten Ni0$27Bi0.73 at 1040 and 1065 C. The depth of the melt above the injection point was varied up to 1.15 m. The feed gas consisted of an 80/20 mol% mixture of methane/argon with a standard feed flow rate 10 cm3 (std).min1 and a pressure of 200 kPa. The composition of the reactor outlet was measured with a mass spectrometer to calculate the methane conversion. The experimental methane conversions reported in Ref. [11] are compared with the predictions of our catalytic LMBR model in Fig. 3A. The agreement between the catalytic LMBR model predictions and the experimental data is excellent at both temperatures. It is crucial to emphasize that this agreement was obtained without the need to adjust any parameter. This is in contrast to previous models in Refs. [11,25] where the bubble size had to be adjusted so that the predicted methane conversions would match the experimental measurements. The model presented here is a significant advance because it calculates the specific gas-liquid interfacial area and the gas holdup from the superficial gas velocity and the properties of the liquid and gas phases. This allows using the model to design LMBRs at larger scales with an increased degree of confidence (Sections: Reactor scale-up and Reactor design optimization). Fig. 3A also shows that when only the non-catalytic methane decomposition reaction is included in the model, the predicted CH4 conversions are much reduced at all reactor heights. This demonstrates the benefit of catalytic molten metals for decreasing the required reactor size for a given CH4 conversion. Fig. 3B shows the profiles of the pressure and superficial gas velocity along the catalytic melt at 1040 C. The pressure decreases almost linearly from bottom to top, as expected from Eq. (6). At any height, the pressure gradient is given by ðrl aDrÞg. Since the gas holdup remains 1% (see Fig. 3C) and the density of the liquid alloy is constant, the pressure gradient is nearly constant along the reactor height. The superficial gas velocity increases almost linearly with respect to reactor height in the bottom half of the melt and faster than linearly in the top half. This is consistent with Eq. 7555 (30) which shows that jg is proportional to ð1 þXCH4 Þ=P. As the height increases, XCH4 increases (Fig. 3A) and P decreases (Fig. 3B). Fig. 3C shows the model predictions for the mean bubble diameter, gas holdup, and reactive surface area versus reactor height at 1040 C for the 1.15-m long bubble reactor in Ref. [11]. As bubbles rise, dvs decreases and a increases from the bottom (dvs;b ¼ 0.809 cm, ab ¼ 0.00280) to the top (dvs;t ¼ 0.697 cm, at ¼ 0.00917). This is because the superficial gas velocity jg increases with height (Fig. 3B). The reactive surface area increases faster than linearly with respect to height. This is consistent with Eq. (46) but contrasts with the assumption made by Upham et al. [11] that the reactive surface area is a linear function of height. According to Eq. (46), the slope of the curve relating Ar to height (dAr =dLÞ is proportional to 6a=dvs . As height increases, the slope increases because a increases and dvs decreases (Fig. 3C). Rahimi et al. [14] carried out methane decomposition experiments in a 1000-mm long, 22-mm inner diameter bubble column reactor containing a 660-mm layer of molten Ni0$27Bi0.73 under a 260-mm layer of molten salt (NaBr). The purpose of the salt layer was to reduce metal losses and carbon contamination by the metal. The experiments investigated the effects of melt temperature and volume flow rate of the inlet gas (pure methane) on CH4 conversion. Our coupled catalytic LMBR model was also able to reproduce adequately the experimental results of Rahimi et al. [14], as shown in Fig. 4. In particular, the decreasing trend of CH4 conversion with increasing inlet flow rate was reproduced by accounting for the non-catalytic methane conversion in the injection tube submerged in the melt. More details on the interpretation of Fig. 4 are provided in Section: 3. Comparison with the experimental data of Rahimi et al. in the Supplementary Material. Reactor scale-up Industrial scale catalytic LMBRs should be able to produce several thousand Nm3.h1 of hydrogen. A practical implementation may consist of several vertical tubes filled with catalytic liquid metal and externally heated by burners in a furnace (Fig. 5). The burners could be fed with natural gas or a fraction of the produced hydrogen. Burners may be located in the bottom of the furnace as illustrated in Fig. 5, or they may be placed in rows on the sidewalls of the furnace. The tops of the tubes connect to a production manifold where the carbon layer could be continuously removed (e.g, skimmed mechanically). Fig. 6A shows the profiles of predicted CH4 conversion and superficial gas velocity in a single tube measuring 0.10 m in internal diameter and 2.11 m in length that could be part of a multitubular reactor for industrial use. The tube is filled with Ni0$27Bi0.73, and the conditions inside the tube are isothermal at 1050 C. Pure CH4 is fed at the bottom of the tube at a rate of 0.0858 mol.s1 and a pressure of 17.8 bar. The CH4 conversion reaches 80.0% at the top of the tube. The superficial gas velocity increases with height: it starts at 6.75 cm s1 at the bottom of the tube and reaches 13.4 cm.s1 at the top. The Froude number, which is proportional to the superficial gas velocity for a given tube diameter (Eq. (26)), is purposefully limited to 0.135 at the top of the tube, since this is the 7556 i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 4 7 ( 2 0 2 2 ) 7 5 4 7 e7 5 6 8 Fig. 3 e Profiles of (A) predicted and experimental methane conversions at 1040 and 1065 C, (B) predicted superficial gas velocity and pressure at 1040 C, and (C) predicted mean bubble diameter, gas holdup, and reactive surface area at 1040 C versus reactor height in a bubble column filled with Ni0·27Bi0.73. The melt was 1.15 m in length and 0.03 m in diameter. The feed was 10 cm3 (std).min¡1 of 80 mol% methane and 20 mol% argon mixture at 200 kPa. Markers are experimental data from Upham et al. [11]. Lines represent model predictions. maximum NFr value at which the correlation for dvs has been experimentally validated [30,36]. Fig. 6B shows the predicted pressure and gas holdup for the same tube. The pressure decreases almost linearly with height down to 16.2 bar at the tube outlet, which corresponds to a pressure drop of 1.6 bar. The gas holdup increases from 11.2% to 13.1% from the bottom to the top of the tube, which is well below the maximum a values of 30e32% for the dvs and a correlations [27,30]. The gas holdups in the scaled up tube are much larger than in the laboratory sized column where they remained below 1% (Fig. 3C). This is due to the higher superficial gas velocities in the scaled up tube. Fig. 6C shows the profiles of the predicted mean bubble diameter and reactive surface area. The mean bubble diameter decreases from 3.16 mm at the bottom to 2.91 mm at the top of the tube, which is considerably smaller than in the laboratory sized column of Upham et al. [11] where dvs ranged from 7 to 8 mm (Fig. 3C). This is due to a combination of a larger tube diameter and higher superficial gas velocity in the scaled-up case, and is consistent with the trends shown in Fig. 2A. The reactive surface area increases slightly faster than linearly with respect to height because the local reactive surface area per unit volume of reactor, 6a=dvs , rises from 213 m2.m3 at the bottom to 270 m2.m3 at the top of the tube. The average time tr (s) taken by bubbles to rise from the bottom of the melt to a height L is: tr ¼ ðL 1 dL ¼ 0 vg ðL a 0 jg dL (47) i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 4 7 ( 2 0 2 2 ) 7 5 4 7 e7 5 6 8 7557 close to the reactor inlet. The average residence time of the gas in the tube is tR ¼ 2.39 s. Reactor design optimization The optimization presented here is based on a H2 production rate of 10,000 Nm3.h1, which corresponds to a small-scale hydrogen plant by recent standards [39]. The feed is pure methane and the selectivity to H2 is assumed to be 100% as this value was nearly achieved under laboratory conditions [11]. The capital cost of the reactor depends greatly on the total volume of the melt given by: Fig. 4 e Predicted and experimental methane conversions versus inlet gas flow rate and melt temperature in a bubble column measuring 22 mm in inner diameter, 1000 mm in length, and containing layers of 660 mm of Ni0·27Bi0.73 and 260 mm of NaBr. The feed was pure methane. Markers are experimental data from Rahimi et al. [14]. Lines represent model predictions. Vmelt ¼ Nt pD2 Lt 4 (49) where Nt is the number of tubes. The other important contributors to the capital cost are the melt composition, the tube material of construction, and the pressure inside the tubes. Molten Cu0$45Bi0.55 is cheaper than Ni0$27Bi0.73 as the price of Cu is roughly half that of Ni [40]. The tube material of construction is selected based on the melt composition, temperature, and pressure. Higher pressures require a larger tube wall thickness, thus increasing the cost of tubes. _ H2 ðtotalÞ, was held The total outlet flow rate of hydrogen, n constant at 122.4 mol s1, which is equivalent to 10,000 Nm3.h1. Because methane decomposition produces 2 mol of H2 for each mole of CH4 that reacts, the total feed flow rate of _ CH4 ðtotalÞ, is given by: methane, n _ CH4 ðtotalÞ ¼ n _ H2 ðtotalÞ n 2 XCH4 ;t (50) where XCH4 ;t is the methane conversion at the top of the tubes. The number of tubes and the inlet flow rate of methane in each tube, n_CH4 ;b , are related by: Nt ¼ _ CH4 ðtotalÞ n n_CH4 ;b (51) Combining Eqs. (49)e(51) yields: Vmelt ¼ Fig. 5 e Conceptual design of a multi-tubular LMBR heated by burners. where vg ¼ jg =a represents the local average rise velocity (m.s1) of the gas through the melt. The average residence time of the gas in the tube is obtained by setting the upper bound of the integral to L ¼ Lt in Eq. (47). ð Lt _ H2 ðtotalÞ pD2 Lt n 8 n_CH4 ;b XCH4 ;t (52) Several optimization problems were solved for specified tube diameters, melt temperatures, CH4 conversions, and melt compositions. - Tube diameters: 0.075, 0.1, 0.125, 0.15, and 0.2 m Melt temperatures: 950, 1000, 1050, 1100, and 1150 C CH4 conversions: 60, 70, 80, and 90% Melts: Ni0$27Bi0.73 and Cu0$45Bi0.55 (48) For each problem, the optimization procedure consisted of finding the tube length, Lt , the feed molar flow rate per tube, _ CH4 ;b , and the feed pressure, Pb , that minimized Vmelt (Eq. (52)) n Fig. 6D shows how vg and tr increase as a function of the reactor height. The increase in gas rise velocity is more pronounced near the bottom of the tube because jg is proportional to 1 þ XCH4 (Eq. (30)), and methane decomposition is faster using the fmincon function with the sqp algorithm in MATLAB 2019b. The number of tubes was then calculated with Eq. (51). The optimizations were subject to the following constraints: tR ¼ 0 a dL jg 7558 i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 4 7 ( 2 0 2 2 ) 7 5 4 7 e7 5 6 8 Fig. 6 e Predicted profiles of (A) CH4 conversion and superficial gas velocity, (B) gas holdup and pressure, (C) reactive surface area and mean bubble diameter, and (D) average rise velocity and rise time of the gas in a single tube. Tube diameter ¼ 0.10 m, length ¼ 2.11 m, melt ¼ Ni0·27Bi0.73, temperature ¼ 1050 C, feed ¼ 0.0858 mol s¡1 of pure CH4 at 17.8 bar. 102 jþ g 2 (53a) at 0:30 (53b) NFr 0:135 (53c) 0:1 Lt ðmÞ 150 (53d) _ CH4 ;b mol:s1 150 0:001 n (53e) 1 Pb ðbarÞ 150 (53f) The upper limits on dimensionless superficial gas velocity (Eq. (53a)), gas holdup (Eq. (53b)), and Froude number (Eq. (53c)) correspond to the ranges of experimental validation of the correlations for a and dvs . The allowable ranges for tube length (Eq. (53d)), methane feed to each tube (Eq. (53e)), and bottom pressure (Eq. (53f)) were chosen to provide a wide search space for the optimization variables. Table 3 summarizes the optimized designs that examine the effects of tube diameter, temperature, methane conversion, and melt composition. These designs are easily adjustable for different rates of H2 production by only changing the number of tubes. For example, optimum designs for 100,000 Nm3.h1 of H2 would require ten times the number of tubes indicated in Table 3. All other parameters would stay the same. In all designs, the Froude number is at the upper limit of its allowed range (0.135) at the top of the tubes. This maximizes the superficial gas velocity, which in turn minimizes 7559 i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 4 7 ( 2 0 2 2 ) 7 5 4 7 e7 5 6 8 Table 3 e Optimal reactor designs (minimum melt volumes) for producing 10,000 Nm3.h¡1 of H2 in a multitubular catalytic LMBR fed by pure CH4. Design Melt composition Methane conversion (%) Temperature ( C) Tube inner diameter (m) Tube length (m) Number of tubes Melt volume (m3)a Methane feed rate per tube (mmol.s1) Inlet pressure (bar) Outlet pressure (bar) Bottom superficial gas velocity (cm.s1) Top superficial gas velocity (cm.s1) Bottom gas holdup (%) Top gas holdup (%) GHSV (Nm3 CH4.m3 catalyst.h1)b Gas residence time (s) Reactor duty (MW) Average heat flux (kW.m2) Bond number (x 103) Galilei number (x 1010) Inlet Froude number (x 102) Outlet Froude number (x 102) a b A B Ni0$27Bi0.73 80 80 1050 1050 0.075 0.10 2.15 2.11 1830 891 17.4 14.8 41.8 85.8 17.84 17.81 16.17 16.19 5.83 6.75 11.6 13.4 10.5 11.2 12.2 13.1 405 482 2.62 2.39 7.52 7.52 8.1 12.7 1.39 2.47 8.90 21.1 6.80 6.82 13.5 13.5 C D E F G H I J K L M 80 1050 0.125 2.07 510 13.0 150 17.79 16.20 7.57 15.0 11.9 13.9 554 2.22 7.52 18.1 3.85 41.2 6.84 13.5 80 1050 0.15 2.04 323 11.6 237 17.76 16.21 8.31 16.4 12.5 14.6 622 2.09 7.52 24.2 5.54 71.2 6.85 13.5 80 1050 0.20 1.98 157 9.8 487 17.72 16.23 9.78 18.9 13.4 15.9 750 1.90 7.52 38.5 9.86 169 6.98 13.5 80 950 0.10 9.42 1987 147 38.5 14.24 6.71 3.50 13.4 8.3 12.5 47.5 13.0 6.76 1.15 2.46 16.4 3.53 13.5 80 1000 0.10 4.49 1242 43.8 61.6 14.70 11.19 5.65 13.4 10.4 12.9 162 5.31 7.13 4.07 2.46 18.7 5.71 13.5 80 1100 0.10 1.02 675 5.39 113 22.97 22.19 7.18 13.4 11.6 13.2 1324 1.14 7.91 36.7 2.47 23.6 7.25 13.5 80 1150 0.10 0.51 526 2.11 146 29.90 29.51 7.50 13.4 11.8 13.4 3393 0.57 8.31 98.5 2.47 26.1 7.57 13.5 60 1050 0.10 1.03 333 2.70 306 52.03 51.26 8.23 13.4 14.0 15.9 3636 1.35 8.03 74.5 2.47 21.1 8.31 13.5 70 1050 0.10 1.46 518 5.94 169 31.13 30.02 7.58 13.4 12.7 14.5 1393 1.78 7.74 32.6 2.47 21.1 7.66 13.5 90 1050 0.10 3.25 2073 52.9 32.8 9.10 6.53 5.05 13.4 9.1 11.3 117 3.44 7.35 3.5 2.47 21.1 5.10 13.5 Cu0$45Bi0.55 80 1050 0.10 1.54 886 10.7 86.3 17.44 16.28 6.93 13.4 11.6 13.4 667 1.77 7.52 17.5 2.18 14.8 7.00 13.5 Includes the volume of gas bubbles dispersed in the liquid metal. The catalyst volume is the liquid metal volume excluding the gas phase. the mean bubble size and maximizes the gas holdup (while keeping it below its allowed maximum of 30%). The net effect is to maximize the reactive surface area, as shown by Eq. (46). A potential source of error when predicting dvs in liquid metals with Eq. (23) is that Akita and Yoshida [30] developed this dimensionless number correlation with liquids having significantly lower densities and surface tensions than those of NieBi and CueBi alloys. Nevertheless, the catalytic LMBR designs in Table 3 correspond to ranges of the Bond, Galilei, and Froude numbers (NBo ¼ 1.39 103 e 9.86 103, NGa ¼ 8.90 1010 e 1.69 1012 and NFr ¼ 3.53 102 e 1.35 101) that are contained within the experimental ranges reported by Akita and Yoshida [30,36] (NBo; exp ¼ 7.98 102 e 4.85 104, NGa; exp ¼ 6.25 106 e 1.79 1012 and NFr; exp ¼ 8 104 e 1.35 101). Effect of tube diameter The first set of optimizations (Designs A e E) was carried out for five different tube diameters between 0.075 and 0.20 m using Ni0$27Bi0.73 at a constant temperature of 1050 C and a CH4 conversion of 80%. The melt volume (Eq. (52)), which includes all the tubes, decreases from 17.4 to 9.8 m3 as the tube diameter increases from 0.075 to 0.20 m. Fig. 7A shows this trend graphically. Meanwhile, the required number of tubes decreases from 1830 to 157 and the tube length decreases from 2.15 to 1.98 m. Fig. 7A provides additional insights on the effect of tube diameter on reactor performance. As the tube diameter increases, the reactive surface area comprising all the tubes, NtAr, where Ar is given by Eq. (46), stays nearly constant. Therefore, the main reason why the melt volume grows with decreasing tube diameter is that the ratio of reactive surface area to melt volume increases. This ratio is expressed mathematically as Nt Ar =Vmelt and is obtained by combining Eqs. (46) and (49): Nt Ar 1 ¼ Vmelt Lt ð Lt 0 6 a dL dvs (54) Eq. (54) shows that Nt Ar =Vmelt is simply the average of the ratio 6a=dvs taken over the length of a tube. Fig. 7B reveals that the average mean bubble diameter decreases with increasing tube diameter, while the average gas holdup increases. These trends, combined with Eq. (54), explain that the ratio of reactive surface area to melt volume increases with tube diameter. Fig. 7C shows that the top pressure stays nearly constant at 16.2 bar as the tube diameter varies. In order for the catalytic and non-catalytic reaction rates to remain above zero at the top of the tubes, the outlet pressure must stay well below the pressure at which the equilibrium conversion is 80%. This pressure, which is denoted Pressure @XCH4, eq on Fig. 7C, is 20.7 bar at 1050 C. The bottom pressure is the sum of the top pressure and the hydrostatic pressure exerted by the melt, which is itself proportional to the tube length. Because the tube length decreases slightly as the tube diameter increases, the bottom pressure also decreases slightly from 17.8 to 17.7 bar. The above results favor larger diameter tubes as they reduce the melt volume, number of tubes, and pressure drop; however, material and heat transfer limitations must also be considered in the selection of tube diameter. The average heat flux from the burners through the tube wall, qav (W.m2), depends on tube diameter and must be kept in a reasonable range. 7560 i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 4 7 ( 2 0 2 2 ) 7 5 4 7 e7 5 6 8 Fig. 7 e Effect of tube diameter on (A) melt volume, reactive surface area, and ratio of reactive surface area to melt volume, (B) average mean bubble diameter and gas holdup, and (C) pressures and tube length. Melt ¼ Ni0·27Bi0.73, T ¼ 1050 C, H2 production ¼ 10,000 Nm3.h¡1, XCH4;t ¼ 0.80. QR qav ¼ pNt DL (55) In Eq. (55), QR is the reactor duty (W), which was determined with the process simulation software Aspen Plus. In our calculations of QR and qav reported in Table 3, the temperature of the methane entering the bottom of the tube was assumed to be 800 C and D was taken as the inner tube diameter. Table 3 shows that qav based on the inner tube surface increases with increasing tube diameter but stays within the normal range of steam methane reformers ( 150 kW.m2) [39,41] even for diameters as high as 0.20 m. Effect of temperature The second set of optimizations (Designs FeI) was carried out for four additional temperatures between 950 C and 1150 C with molten Ni0$27Bi0.73 in tubes measuring 0.10 m in diameter at a CH4 conversion of 80% (Table 3). Fig. 8A shows that the total reactive surface area decreases nearly exponentially with increasing temperature. This is consistent with the exponential increase of catalytic reaction rates with temperature implied by Arrhenius law. On the other hand, the melt volume decreases somewhat faster than exponentially with respect to temperature, from 147 m3 at 950 C to 2.11 m3 at 1150 C. This is because the ratio of reactive surface area to melt volume, which is itself related to the ratio of gas holdup to mean bubble diameter by Eq. (54), increases with temperature. Indeed, Fig. 8B shows that the average gas holdup increases and the average mean bubble diameter decreases with increasing temperature. Fig. 8C shows that the top pressure increases with temperature as the pressure at which the equilibrium conversion is 80% (Pressure @XCH4, eq) increases from 10.5 bar at 950 C to i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 4 7 ( 2 0 2 2 ) 7 5 4 7 e7 5 6 8 7561 Fig. 8 e Effect of temperature on (A) melt volume, reactive surface area, and ratio of reactive surface area to melt volume, (B) average mean bubble diameter and gas holdups, and (C) pressures and tube length. Melt ¼ Ni0·27Bi0.73, tube diameter ¼ 0.10 m, H2 production ¼ 10,000 Nm3.h¡1, XCH4;t ¼ 0.80. 37.1 bar at 1150 C. The increase in temperature causes the optimum tube length to drop considerably from 9.4 to 0.51 m, and the number of tubes to decrease from 1987 to 526 tubes. The optimum bottom pressure, which is the sum of the top pressure and the hydrostatic pressure exerted by the melt, increases with increasing temperature. The pressure drop, i.e. the difference between the bottom and top pressures, decreases with increasing temperature. Effect of methane conversion The third set of optimizations (Designs J e L) was carried out at three additional CH4 conversions (60%, 70% and 90%) with molten Ni0$27Bi0.73 in tubes measuring 0.10 m in diameter at a constant temperature of 1050 C (Table 3). Higher CH4 conversions require larger reactive surface areas and, therefore, larger melt volumes and smaller gas hourly space velocities (GHSV) (Fig. 9A). The melt volume at 60% conversion is relatively small (2.70 m3) but increases to 52.9 m3 at 90% conversion. As conversion increases from 60 to 90%, the pressure that corresponds to the equilibrium of the methane decomposition reaction (denoted Pressure @XCH4, eq on Fig. 9B) decreases from 65.4 to 8.6 bar. Hence, the optimum pressure at the top of the tubes must also decrease in order to stay below the equilibrium pressure and, thus, allow for a positive reaction rate. The pressure drop in the tubes (difference between bottom and top pressures) depends on the height of the tubes though the hydrostatic pressure exerted by the melt. Because the tube length increases with methane conversion, the pressure drop also increases. The reactor duty decreases from 8.03 MW to 7.35 MW as the CH4 conversion increases from 60% to 90%. This is because less methane needs to be heated from 800 C to the reaction temperature of 1050 C as conversion increases. The heat consumed by the reaction is the same at all CH4 conversions since the hydrogen production rate is fixed at 10,000 Nm3.h1. The inner area of the tubes (i.e., the area for heat transfer to the melt) is proportional to the number of tubes and their 7562 i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 4 7 ( 2 0 2 2 ) 7 5 4 7 e7 5 6 8 Fig. 9 e Effect of CH4 conversion on (A) melt volume, reactive surface area, and GHSV, (B) pressures and tube length, and (C) average heat flux and inner tube area. Melt ¼ Ni0·27Bi0.73, T ¼ 1050 C, tube diameter ¼ 0.10 m, H2 production ¼ 10,000 Nm3.h¡1. length, both of which increase with methane conversion. As a result, the average heat flux decreases from 74.5 kW.m2 to 3.5 kW.m2 as the CH4 conversion increases from 60% to 90% (Fig. 9C). Effect of catalytic molten metal alloy composition The effect of catalyst composition can be determined by comparing Designs B and M, which use different molten metal alloys but the same temperature (1050 C), tube diameter (0.10 m), and CH4 conversion (80%). The required melt volume is smaller for Design M (10.7 m3 of Cu0$45Bi0.55) than for Design B (14.8 m3 of Ni0$27Bi0.73), which represents a volume reduction of 38%. This is consistent with the experimental observation of Palmer et al. [16] that Cu0$45Bi0.55 is a somewhat more active catalyst than Ni0$27Bi0.73. The lower melt volume for Cu0$45Bi0.55 vs. Ni0.27Bi0.73 results in shorter tubes (1.54 m vs 2.11 m) but nearly the same number of tubes (886 tubes vs 891 tubes). The inlet pressure is somewhat lower for Cu0$45Bi0.55 (17.4 bar vs 17.8 bar). Given that copper costs about half the price of nickel per tonne of metal [40], Cu0$45Bi0.55 is expected to result in a cheaper initial catalyst load for the reactor. Nevertheless, other factors such as potential differences in the quality of the produced carbon, tendencies of Ni and Cu to leave the reactor with the carbon product, and melt compatibility with tube construction materials should be investigated. Discussion Variations from optimum designs The optimized reactor designs listed in Table 3 minimize the melt volumes but sometimes lead to large internal pressures that may cause excessive creep rates in metal tubes subjected to very high temperatures. The combination of internal pressure, tube diameter, and tube wall thickness determines the hoop stress. At very high temperatures, excessive hoop stresses lead to creep damage that limits the tube lifetime i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 4 7 ( 2 0 2 2 ) 7 5 4 7 e7 5 6 8 [39,41,42]. Special alloys containing high concentrations of Ni and Cr (e.g., HP-40 grade Centralloy G4852 Micro R [43]) are used for steam methane reformer tubes exposed to temperatures up to 1050 C [44]. The lifetime of a tube exposed to a given temperature and hoop stress can be estimated by means of a Larson-Miller plot specific to the tube material of construction, as shown in Section: 4. Influence of creep on tube life in the Supplementary Material. Our analysis is based on a design life of 100,000 h, a maximum tube temperature of 1050 C, and HP-40 grade Centralloy G4852 Micro R as the tube material. When the tube diameter is set at 0.10 m and the maximum internal pressure is 52 bar (Design J), the tubes require a wall thickness of 55 mm. This is much thicker than tubes used in steam methane reformer, which typically have wall thicknesses between 10 and 25 mm [42]. The internal pressure should be below 33 bar to keep the wall thickness under 25 mm. This prediction does not account for how the lifetime of the tubes may be affected by chemical compatibility issues between the molten metal alloy and the tube material. This important aspect is outside the scope of this article. In the case of Design J (Ni0$27Bi0.73, T ¼ 1050 C, D ¼ 0.10 m, CH4 conversion ¼ 60%), it is worth finding what penalty must be paid in terms of increased melt volume if the maximum internal pressure is limited to 30 bar in order to bring the tube wall thickness to a reasonable value (21 mm). Hence, Design J was re-run with the additional constraint Pb 30 bar, resulting in a new design named J’. The two designs are compared in Table 4. Table 4 e Effect of constraining the pressure to Pb 30 bar on optimal reactor design (minimum melt volumes) for producing 10,000 Nm3.h¡1 of hydrogen with a CH4 conversion of 60% in a multitubular reactor fed by pure methane. Design Melt composition Temperature ( C) Methane conversion (%) Tube inner diameter (m) Tube length (m) Number of tubes Melt volume (m3)a Methane feed rate per tube (mmol.s1) Inlet pressure (bar) Outlet pressure (bar) Bottom superficial gas velocity (cm.s1) Top superficial gas velocity (cm.s1) Bottom gas holdup (%) Top gas holdup (%) GHSV (Nm3 CH4.m3 catalyst.h1)b Gas residence time (s) Reactor duty (MW) Average heat flux (kW.m2) Bond number (x 103) Galilei number (x 1010) Inlet Froude number (x 102) Outlet Froude number (x 102) a b J Ni0$27Bi0.73 1050 60 0.10 1.03 333 2.70 306 52.03 51.26 8.23 13.4 14.0 15.9 3636 1.35 8.03 74.5 2.47 21.1 8.31 13.5 J0 0.81 582 3.70 175 30.00 29.39 8.19 13.4 13.1 14.9 2618 1.02 8.03 54.3 2.47 21.1 8.27 13.5 Includes the volume of gas bubbles dispersed in the liquid metal. The catalyst volume is the liquid metal volume excluding the gas phase. 7563 Unsurprisingly, the new optimum inlet pressure is the maximum allowed, i.e., 30 bar. The melt volume increased from 2.70 m3 (Design J) to 3.70 m3 (Design J’). The number of tubes increased from 333 to 582, while the tube length decreased from 1.03 to 0.81 m. The pressure drop decreased from 0.77 to 0.61 bar. Hence, the main penalty for limiting the maximum pressure is a higher reactor volume and catalyst load (þ37% in this particular example). Comparison to previous catalytic LMBR design from the literature Von Wald et al. [29] investigated the optimal dimensions of a catalytic LMBR to produce 10.4 kt.a1 (i.e., 13,400 Nm3.h1) of hydrogen as part of a methane pyrolysis energy system which also included heat exchange and separation equipment. The optimization minimized the H2 price required to achieve a net present value of zero at a specified internal rate of return. The reactor consisted of a single bubble column containing Ni0$27Bi0.73. The reactor inlet pressure, temperature, and methane conversion were fixed at 30 barg, 1100 C, and 90%, respectively. The optimum column radius and height were determined to be approximately 2.2 m and 1 m, respectively. When the coupled catalytic LMBR model developed in the present article is applied to the column dimensions determined by Von Wald et al., the resulting methane conversion is only 78.7%, which is well below the specified 90%. In order to reach 90% methane conversion with a column measuring 2.2m in radius, our model predicts that the height should extend to 24.5 m, which is considerably larger than the height of 1 m calculated by Von Wald et al. These discrepancies are explained by differences in assumptions between the two approaches. Most importantly, our kinetic model is constrained by thermodynamic equilibrium (Eqs. (9) and (13)). By contrast, Von Wald et al. included no equilibrium constraint and stated that the continuous removal of carbon from the system may allow the maximum mole fraction of H2 to exceed its equilibrium value. In Section: 5. Chemical equilibrium equation of the Supplementary Material, we show that the chemical equilibrium of methane decomposition is independent of the concentration of carbon and, therefore, is not affected by carbon removal. Von Wald et al. also contended that improvements in bubble column design made by Farmer et al. [25] can allow H2 mole fractions to exceed equilibrium values. Farmer et al. modelled a reactor containing a membrane around the melt to selectively remove nearly all H2 before the bubbles left the melt. Selective H2 removal shifts the equilibrium of the CH4 decomposition reaction to the right to enable the CH4 conversion to reach nearly 100%. In contrast, non-membrane LMBRs, such as the reactor modelled by Von Wald et al., do not selectively remove the products of the reaction along their length and therefore necessitate the inclusion of thermodynamic equilibrium limitations to be realistically sized. Another reason for differences in predicted heights between our model and that of Von Wald et al. is related to the reactor pressure drop. At 1100 C, the pressure corresponding to 90% equilibrium conversion is 11.5 bar. Since the inlet pressure at the bottom of the column is 31 bar (30 barg), the CH4 conversion cannot reach 90% until such a height where the pressure drops below 11.5 bar. Indeed, our model predicts 7564 i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 4 7 ( 2 0 2 2 ) 7 5 4 7 e7 5 6 8 that the outlet pressure is 11.2 bar at the height of 24.5 m where the methane conversion reaches 90%d. By contrast, Von Wald et al. assume that the pressure stays constant in the reactor. Moreover, our model accounts for the rate of noncatalytic methane decomposition, unlike that of Von Wald et al. Besides the thermodynamic equilibrium constraint, the reactor pressure drop, and accounting for the non-catalytic methane decomposition, there are also significant differences between the two models regarding the treatment of hydrodynamics. Von Wald et al. assume that the changes in bubble diameter and gas holdup with height only depend on CH4 conversion according to the following two equations: dv ¼ dvb ð1 þ XCH4 Þ1=3 (56) av ¼ avb ð1 þ XCH4 Þ (57) where dv and av are respectively the bubble diameter and the gas holdup in their model, and the subscript b refers to the bottom of the column. Eqs. (56) and (57) are tantamount to assuming that the mean bubble volume and the gas holdup are only affected by the stoichiometry of the methane decomposition reaction, and they neglect the effect of decreasing gas pressure with height. By contrast, our model calculates the mean bubble diameter and gas holdup at any height using respectively Eq. (23) and Eq. (32), which have been validated experimentally. Fig. 10 shows that Eq. (56) in the model of Von Wald et al. and Eq. (23) in our model predict opposing trends for the change of bubble diameter versus height. Moreover, Eq. (32) in our model predicts substantially smaller increases in gas holdup with respect to height than Eq. (57) from Von Wald et al. [29]. Comparison to steam methane reformer In this section, the operational characteristics of a multitubular catalytic LMBR are compared to those of a steam methane reformer (SMR). Both designs use a large number of externally fired tubes. The SMR tubes are packed with porous catalyst pellets. A typical modern reformer for hydrogen production operates with a molar H2O/CH4 ratio of 1.8e2.5, a pressure of 30 bar, and a syngas outlet temperature of 920 C. The produced syngas contains 71.1 mol% H2, 17.6 mol% CO, 4.6 mol% CO2 and 6.7 mol% CH4 on a water free basis [44]. The reformer is followed by a water gas shift (WGS) reaction system where nearly all the CO reacts with H2O to produce more H2 and CO2 (Eq. (3)). As a result, a typical syngas composition at the exit of the shift reactor is 75.4 mol% H2, 18.9 mol% CO2 and 5.7 mol% CH4 on a water free basis. By contrast, methane decomposition produces no CO2, and the gas outlet of the catalytic LMBR consists of a mixture of H2 and unreacted CH4. When pure methane is fed to a catalytic LMBR, the stoichiometry of methane decomposition (Eq. (1)) indicates that the mole fraction of H2 in the outlet gas, yH2 , is related to the methane conversion by: yH2 ¼ 2XCH4;t 1 þ XCH4 ;t (58) Eq. (58) implies that a methane conversion of 60.5% is required in the catalytic LMBR to obtain yH2 ¼ 0.754, i.e., the same hydrogen mole fraction produced by a modern SMR followed by WGS. If a higher hydrogen purity is required, it can be increased to 99.999 mol% by using a pressure swing adsorption (PSA) unit downstream of the LMBR or WGS. The PSA off-gas may then be used as fuel in the burners of the LMBR or SMR. Table 5 e Comparison of operating conditions for producing hydrogen by SMR in tubular catalytic reformers and by methane decomposition in a catalytic LMBR. Design H2 production rate (Nm3/h) P (bar) Outlet gas temperature ( C) Tube diameter (m) Tube heated length (m) Number of tubes Fig. 10 e Comparisons of mean bubble diameter and gas holdup vs height calculated by our model and that of Von Wald et al. [29]. Tube diameter ¼ 0.10 m, length ¼ 2.1 m, melt ¼ Ni0·27Bi0.73, temperature ¼ 1050 C, feed ¼ 0.0858 mol s¡1 of pure CH4 at 17.8 bar. d Although the column diameter of 4.4 m results in Bond and Galilei numbers that are well above their validated ranges for prediction of mean bubble diameters, our model still reliably demonstrates the effect of thermodynamic equilibrium on the required reactor height. Reactor volume (m3) GHSV (Nm3 CH4.m3 catalyst.h1) Pressure drop (bar) Average heat fluxc (kW.m2) a b c Modern tubular reformer LMBR (Case J0 )a 2500e300,000 [39] 30 [39] 700e950 [39,44,45] 0.07e0.20 [39,41] 6e13 [39,41] 12 to more than 800 [39,41,46] 30b [39] 1345b [39] 100,000 30 1050 0.10 0.81 5820 0.6e0.7 [46] 45e150 [39,41] 0.61 54.3 37 2618 The number of tubes and the reactor volume of Case J0 in Table 4 were multiplied by a factor of 10 to account for the tenfold increase in the hydrogen production rate. Based on 100,000 Nm3/h hydrogen, reformer outlet gas at 875 C and 31 bar, 0.10-m inner tube diameter, 13-m long tubes, and 294 tubes [39]. Based on m2 of inner tube surface. i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 4 7 ( 2 0 2 2 ) 7 5 4 7 e7 5 6 8 Table 5 compares the operating conditions of a catalytic LMBR producing 100,000 Nm3.h1 hydrogen with 60% CH4 conversion to those of modern tubular steam methane reformers. The LMBR gas pressure, tube diameter, pressure drop, and average heat flux fall within the range of SMR operating conditions. In contrast, the LMBR uses a higher outlet gas temperature, a shorter tube length, and a larger number of tubes. Considering the high density of the molten metal alloy, the lower tube length is beneficial for limiting the pressure drop in the LMBR. For the given hydrogen production rate, the volumes of the catalytic LMBR and tubular reformer are remarkably similar (37 vs 30 m3). The GHSV of the catalytic LMBR is nearly double that of the tubular reformer because methane decomposition (Eq. (1)) consumes twice as much methane compared to SMR þ WGS (Eq. (4)) to produce a given amount of hydrogen. 7565 with higher temperature and lower methane conversion. The pressure drop, which is nearly proportional to tube length, decreased with increasing temperature and decreasing methane conversion. Changing the liquid metal composition from Ni0$27Bi0.73 to Cu0$45Bi0.55 resulted in a smaller melt volume for a given application because the kinetics of catalytic methane decomposition is faster with the copper-based catalyst. This is encouraging since the price of copper is roughly half that of nickel on a mass basis. All the multitubular LMBR designs presented in this article (Tables 3 and 4) can be easily adapted to different H2 production rates by changing the number of tubes proportionally. The optimum values of the tube length, bottom pressure, and top pressure are independent of the H2 production rate. Declaration of competing interest Conclusions The catalytic LMBR model developed in this article was successfully validated at various temperatures with previously reported experimental data for methane decomposition in laboratory-sized bubble columns filled with molten Ni0$27Bi0.73. The model was applied to design industrial multitubular LMBRs producing 10,000 Nm3.h1 of hydrogen. The designs were optimized by minimizing the volume of the molten metal for various combinations of tube diameters, melt temperatures, methane conversions and molten metal compositions. The optimization always resulted in the Froude number reaching its maximum allowed value of 0.135 at the top of the tubes. For a given tube diameter, maximizing the Froude number is equivalent to maximizing the superficial gas velocity. Large superficial gas velocities increase the specific gas-liquid interfacial area and the gas holdup, thus favouring both the catalytic and non-catalytic reaction rates. Mean bubble sizes were smaller and gas holdups were higher in industrial applications compared to laboratory LMBRs due to faster superficial gas velocities and larger tube diameters. For a given melt temperature and methane conversion, the lowest melt volume was achieved for the largest tube diameter (20 cm) because the ratio of reactive area to melt volume increased with tube diameter. Larger tube diameters resulted in slightly shorter tube lengths and a much lower number of tubes. Nevertheless, the maximum tube diameter is limited by material and heat transfer considerations in practical applications. The minimum melt volume dropped somewhat faster than exponentially as a function of melt temperature for a given tube diameter and methane conversion. Both the optimum tube length and number of tubes decreased greatly with increasing temperature. Specifying a larger methane conversion increased the melt volume considerably for a given tube diameter and melt temperature. Both the optimum tube length and number of tubes increased with increasing methane conversion. The outlet pressure at the top of the tubes always stayed below the pressure corresponding to the equilibrium CH4 conversion in order to keep the reaction rate positive. The optimum inlet pressure at the bottom of the tubes increased The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper. Acknowledgements The authors would like to thank the Natural Sciences and Engineering Research Council of Canada (Discovery Grant Program) for providing funding for this work. Appendix A. Supplementary data Supplementary data to this article can be found online at https://doi.org/10.1016/j.ijhydene.2021.12.089. Nomenclature Ar ag Ci C0 D DH do dvs d*vs Eac;f Ean;f g reactive surface area for the entire column or tube (m2) specific interfacial area per unit of gas volume (m2 (interface).m3 (gas)) concentration of component i in the gas phase (mol.m3) distribution parameter (unitless) reactor or tube inner diameter (m) dimensionless hydraulic equivalent diameter of the reactor (m) orifice diameter (m) volume-surface mean bubble diameter in most of the reactor (m) volume-surface mean bubble diameter just above the injector (m) activation energy of the forward catalytic reaction (J.mol1) activation energy of the forward non-catalytic reaction (J.mol1) gravitational acceleration (9.81 m.s2) 7566 jg kc;f koc;f kn;f kon;f K KC L M n n_ NBo NFr NGa Nm Nt P qav QR rc R Rc Rn T tr tR uo V Vgj Vmelt V_ vg X i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 4 7 ( 2 0 2 2 ) 7 5 4 7 e7 5 6 8 superficial gas velocity (m.s1) forward rate constant of the catalytic reaction (m.s1) forward pre-exponential factor of the catalytic reaction (m.s1) forward rate constant of the non-catalytic reaction (m3(n1).mol(1n).s1) forward pre-exponential factor of the non-catalytic reaction (m3(n1).mol(1n).s1) reaction equilibrium constant based on fugacity ratios (unitless) reaction equilibrium constant based on concentrations (mol.m3) length (m) molar mass (kg.mol1) order of the forward non-catalytic reaction with respect to CH4 (unitless) mole flow rate (mol.s1) Bond number (dimensionless) Froude number (dimensionless) Galilei number (dimensionless) viscosity number (dimensionless) number of tubes in multitubular reactor pressure (Pa) average heat flux (W.m2) reactor duty (W) catalytic reaction rate per unit of interfacial area (mol CH4.m2.s1) universal gas constant (8.314 J.mol1.K1) catalytic reaction rate per unit volume of gas (mol CH4.m3.s1) non-catalytic reaction rate per unit volume of gas (mol CH4.m3.s1) temperature (K) average rise time of the gas to a height L in the melt (s) average residence time of the gas in the melt (s) gas velocity through orifice (m.s1) volume (m3) void-fraction weighted mean drift velocity (m.s1) total volume of melt (m3) volumetric flow rate (m3.s1) local average rise velocity of the gas (m.s1) conversion (fraction) Greek letters a gas holdup (fraction) ε mole fraction of CH4 in the reactor feed mole ratio of component i (H2 or inert) to CH4 in the qi reactor feed ml dynamic viscosity of liquid metal (Pa.s or kg.m1.s1) kinematic viscosity of liquid metal (m2.s1) nl stoichiometric coefficient of component i (unitless) ni Dr density difference between liquid metal and gas (kg.m3) r density (kg.m3) s surface tension of the liquid metal (N.m1) sh hoop stress (MPa) Subscripts b bottom c catalytic f forward g gas I inert component l liquid metal n non-catalytic t top w wall Superscripts þ dimensionless o standard references [1] Muradov N. Low to near-zero CO2 production of hydrogen from fossil fuels: status and perspectives. Int J Hydrogen Energy 2017;42:14058e88. https://doi.org/10.1016/ j.ijhydene.2017.04.101. [2] Parkinson B, Balcombe P, Speirs JF, Hawkes AD, Hellgardt K. Levelized cost of CO2 mitigation from hydrogen production routes. Energy Environ Sci 2019;12:19e40. https://doi.org/ 10.1039/c8ee02079e. [3] Abbas HF, Wan Daud WMA. Hydrogen production by methane decomposition: a review. Int J Hydrogen Energy 2010;35:1160e90. https://doi.org/10.1016/ j.ijhydene.2009.11.036. nades A, Ruiz E, Ferruelo EM, Herna ndez F, Cabanillas A, [4] Aba pez C, Gavela R, Barrera G, Martı́nez-Val JM, Rubio JA, Lo rrez D. Experimental Rubbia C, Salmieri D, Rodilla E, Gutie analysis of direct thermal methane cracking. Int J Hydrogen Energy 2011;36:12877e86. https://doi.org/10.1016/ j.ijhydene.2011.07.081. [5] Muradov N. Hydrogen via methane decomposition: an application for decarbonization of fossil fuels. Int J Hydrogen Energy 2001;26:1165e75. https://doi.org/10.1016/S03603199(01)00073-8. [6] Qian JX, Chen TW, Enakonda LR, Bin Liu D, Basset JM, Zhou L. Methane decomposition to pure hydrogen and carbon nano materials: state-of-the-art and future perspectives. Int J Hydrogen Energy 2020;45:15721e43. https://doi.org/10.1016/ j.ijhydene.2020.04.100. [7] Qian JX, Enakonda LR, Wang WJ, Gary D, Del-Gallo P, Basset JM, Bin Liu D, Zhou L. Optimization of a fluidized bed reactor for methane decomposition over Fe/Al2O3 catalysts: activity and regeneration studies. Int J Hydrogen Energy 2019;44:31700e11. https://doi.org/10.1016/ j.ijhydene.2019.10.058. [8] Jang HT, Cha WS. Hydrogen production by the thermocatalytic decomposition of methane in a fluidized bed reactor. Kor J Chem Eng 2007;24:374e7. https://doi.org/ 10.1007/s11814-007-5037-9. [9] Qian JX, Bin Liu D, Basset JM, Zhou L. Methane decomposition to produce hydrogen and carbon nanomaterials over costless, iron-containing catalysts. J Clean Prod 2021;320:128879. https://doi.org/10.1016/ j.jclepro.2021.128879. [10] Catalan LJJ, Rezaei E. Coupled hydrodynamic and kinetic model of liquid metal bubble reactor for hydrogen production by noncatalytic thermal decomposition of i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 4 7 ( 2 0 2 2 ) 7 5 4 7 e7 5 6 8 [11] [12] [13] [14] [15] [16] [17] [18] [19] [20] [21] [22] [23] methane. Int J Hydrogen Energy 2020;45:2486e503. https:// doi.org/10.1016/j.ijhydene.2019.11.143. Upham DC, Agarwal V, McFarland EW, Khechfe A, Gordon MJ, Metiu H, Snodgrass ZR. Catalytic molten metals for the direct conversion of methane to hydrogen and separable carbon. Science 2017;358:917e21. https://doi.org/ 10.1126/science.aao5023. nades A, Heinzel A, Mehravaran K, Müller G, Geißler T, Aba Rathnam RK, Rubbia C, Salmieri D, Stoppel L, Stückrad S, Weisenburger A, Wenninger H, Wetzel T. Hydrogen production via methane pyrolysis in a liquid metal bubble column reactor with a packed bed. Chem Eng J 2016;299:192e200. https://doi.org/10.1016/j.cej.2016.04.066. Parkinson B, Tabatabaei M, Upham DC, Ballinger B, Greig C, Smart S, McFarland E. Hydrogen production using methane: techno-economics of decarbonizing fuels and chemicals. Int J Hydrogen Energy 2018;43:2540e55. https://doi.org/10.1016/ j.ijhydene.2017.12.081. Rahimi N, Kang D, Gelinas J, Menon A, Gordon MJ, Metiu H, McFarland EW. Solid carbon production and recovery from high temperature methane pyrolysis in bubble columns containing molten metals and molten salts. Carbon 2019;151:181e91. https://doi.org/10.1016/ j.carbon.2019.05.041. Palmer C, Bunyan E, Gelinas J, Gordon MJ, Metiu H, McFarland EW. CO2-free hydrogen production by catalytic pyrolysis of hydrocarbon feedstocks in molten Ni-Bi. Energy Fuel 2020;34:16073e80. https://doi.org/10.1021/ acs.energyfuels.0c03080. Palmer C, Tarazkar M, Kristoffersen HH, Gelinas J, Gordon MJ, McFarland EW, Metiu H. Methane pyrolysis with a molten CueBi alloy catalyst. ACS Catal 2019;9:8337e45. https:// doi.org/10.1021/acscatal.9b01833. rez BJ, Medrano Jime nez JA, Bhardwaj R, Goetheer E, Leal Pe van Sint Annaland M, Gallucci F. Methane pyrolysis in a molten gallium bubble column reactor for sustainable hydrogen production: proof of concept & techno-economic assessment. Int J Hydrogen Energy 2021;46:4917e35. https:// doi.org/10.1016/j.ijhydene.2020.11.079. Kang D, Rahimi N, Gordon MJ, Metiu H, McFarland EW. Catalytic methane pyrolysis in molten MnCl2-KCl. Appl Catal B Environ 2019;254:659e66. https://doi.org/10.1016/ j.apcatb.2019.05.026. Kang D, Palmer C, Mannini D, Rahimi N, Gordon MJ, Metiu H, McFarland EW. Catalytic methane pyrolysis in molten alkali chloride salts containing iron. ACS Catal 2020;10:7032e42. https://doi.org/10.1021/acscatal.0c01262. Parkinson B, Patzschke CF, Nikolis D, Raman S, Hellgardt K. Molten salt bubble columns for low-carbon hydrogen from CH4 pyrolysis: mass transfer and carbon formation mechanisms. Chem Eng J 2021;417:127407. https://doi.org/ 10.1016/j.cej.2020.127407. Parkinson B, Patzschke CF, Nikolis D, Raman S, Dankworth DC, Hellgardt K. Methane pyrolysis in monovalent alkali halide salts: kinetics and pyrolytic carbon properties. Int J Hydrogen Energy 2021;46:6225e38. https:// doi.org/10.1016/j.ijhydene.2020.11.150. Patzschke CF, Parkinson B, Willis JJ, Nandi P, Love AM, Raman S, Hellgardt K. Co-Mn catalysts for H2 production via methane pyrolysis in molten salts. Chem Eng J 2021;414:128730. https://doi.org/10.1016/j.cej.2021.128730. U.S. Department of Energy (DOE). A BestPractices process heating technical brief materials selection considerations for [24] [25] [26] [27] [28] [29] [30] [31] [32] [33] [34] [35] [36] [37] [38] [39] [40] 7567 thermal process equipment energy efficiency and renewable energy. 2004. https://www.energy.gov/sites/prod/files/2014/ 05/f15/proc_heat_tech_brief.pdf. Parkinson B, Matthews JW, McConnaughy TB, Upham DC, McFarland EW. Techno-economic analysis of methane pyrolysis in molten metals: decarbonizing natural gas. Chem Eng Technol 2017;40:1022e30. https://doi.org/10.1002/ ceat.201600414. Farmer TC, McFarland EW, Doherty MF. Membrane bubble column reactor model for the production of hydrogen by methane pyrolysis. Int J Hydrogen Energy 2019;44:14721e31. https://doi.org/10.1016/j.ijhydene.2019.03.023. Kataoka I, Ishii M. Drift flux model for large diameter pipe and new correlation for pool void fraction. Int J Heat Mass Tran 1987;30:1927e39. https://doi.org/10.1016/0017-9310(87) 90251-1. Hibiki T, Saito Y, Mishima K, Tobita Y, Konishi K, Matsubayashi M. Study on flow characteristics in gas-molten metal mixture pool. Nucl Eng Des 2000;196:233e45. https:// doi.org/10.1016/S0029-5493(99)00293-9. Keipi T, Li T, Løvås T, Tolvanen H, Konttinen J. Methane thermal decomposition in regenerative heat exchanger reactor: experimental and modeling study. Energy 2017;135:823e32. https://doi.org/10.1016/ j.energy.2017.06.176. Von Wald GA, Masnadi MS, Upham DC, Brandt AR. Optimization-based technoeconomic analysis of moltenmedia methane pyrolysis for reducing industrial sector CO2 emissions. Sustain Energy Fuels 2020;4:4598e613. https:// doi.org/10.1039/d0se00427h. Akita K, Yoshida F. Bubble size, interfacial area, and liquidphase mass transfer coefficient in bubble columns. Ind Eng Chem Process Des Dev 1974;13:84e91. https://doi.org/ 10.1021/i260049a016. Fogler HS. Elements of chemical reaction engineering. 5th ed. Prentice Hall; 2016. Sun Z, Parkinson B, Agbede OO, Hellgardt K. Noninvasive differential pressure technique for bubble characterization in high-temperature opaque systems. Ind Eng Chem Res 2020;59:6236e46. https://doi.org/10.1021/acs.iecr.9b04466. Andreini RW, Foster RJ, Callen JS. Characterization of gas bubbles injected into molten metals under laminar flow conditions. Metall Trans B 1977;8B:625e31. Sano M, Mori K. Bubble formation from single nozzles in liquid metals. Trans Jpn Inst Met 1976;17:344e52. Jamialahmadi M, Zehtaban MR, Müller-Steinhagen H, Sarrafi A, Smith JM. Study of bubble formation under constant flow conditions. Chem Eng Res Des 2001;79:523e32. https://doi.org/10.1205/02638760152424299. Akita K, Yoshida F. Gas holdup and volumetric mass transfer coefficient in bubble columns. Ind Eng Chem Process Des Dev 1973;12:76e80. https://doi.org/10.1021/i260045a015. Ishii M. One-dimensional drift-flux model and constitutive equations for relative motion between phases in various two-phase flow regimes. Argonne National Laboratory; 1977. Saito M, Sawada T, Teraoka Y, Nezu A. Dispersion characteristics of gas-liquid two-phase pools. In: Proc. Sixth int. Conf. Nucl. Eng. ICONE-6, San Diego, CA, USA; 1998. Rostrup-Nielsen J, Christiansen LJ. Concepts in syngas manufacture. London: Imperial College Press; 2011. Daily metal prices (n.d.), https://www.dailymetalprice.com/ metalprices.php?c¼ni&u¼lb&d¼20. [Accessed 25 June 2021]. 7568 i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 4 7 ( 2 0 2 2 ) 7 5 4 7 e7 5 6 8 [41] Rostrup-Nielsen JR. Catalytic steam reforming. Berlin: Springer-Verlag; 1984. [42] Guglielmino E, Pino R, Servetto C, Sili A. Creep damage of high alloyed reformer tubes. In: Makhlouf ASH, Aliofkhazrei M, editors. Handb. Mater. Fail. Anal. With case stud. From chem. Concr. Power ind. Oxford, UK: Butterworth Heinemann; 2016. p. 69e122. [43] Schmidt þ Clemens Group. Centralloy ® G 4852 Micro R material data sheet. Germany: Lindlar; 2009. [44] Rostrup-Nielsen JR. Production of synthesis gas. Catal Today 1993;18:305e24. https://doi.org/10.1016/09205861(93)80059-A. [45] Fowles M, Carlsson M. Steam reforming of hydrocarbons for synthesis gas production. Top Catal 2021. https://doi.org/ 10.1007/s11244-021-01496-z. [46] Basin M, Gary D, Whittenberger W, Shah J. Increased steam methane reformer throughput. Www.Digitalrefining.Com; 2017. p. 1e5.