Next Article in Journal
Three-Dimensional Numerical Simulation of Effective Thermal Conductivity and Fractal Dimension of Non-Aqueous Phase Liquid-Contaminated Soils at Mesoscopic Scale
Previous Article in Journal
Optimizing Hydrophobicity of Cu@Zn Foam Catalysts for Efficient CO2 Electroreduction in a Microchannel Reactor
Previous Article in Special Issue
Optimal Maintenance Strategy Selection for Oil and Gas Industry Equipment Using a Combined Analytical Hierarchy Process–Technique for Order of Preference by Similarity to an Ideal Solution: A Case Study in the Oil and Gas Industry
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Reduced-Order Modeling (ROM) of a Segmented Plug-Flow Reactor (PFR) for Hydrogen Separation in Integrated Gasification Combined Cycles (IGCC)

College of Engineering, University of Buraimi, Al Buraimi 512, Oman
Processes 2025, 13(5), 1455; https://doi.org/10.3390/pr13051455
Submission received: 24 March 2025 / Revised: 7 May 2025 / Accepted: 7 May 2025 / Published: 9 May 2025

Abstract

:
In an integrated gasification combined cycle (IGCC), a gasification process produces a gas stream from a solid fuel, such as coal or biomass. This gas (syngas or synthesis gas) resulting from the gasification process contains carbon monoxide, molecular hydrogen, and carbon dioxide (other gaseous components may also be present depending on the gasified solid fuel and the gasifying agent). Separating hydrogen from this syngas stream has advantages. One of the methods to separate hydrogen from syngas is selective permeation through a palladium-based metal membrane. This separation process is complicated as it depends nonlinearly on various variables. Thus, it is desirable to develop a simplified reduced-order model (ROM) that can rapidly estimate the separation performance under various operational conditions, as a preliminary stage of computer-aided engineering (CAE) in chemical processes and sustainable industrial operations. To fill this gap, we present here a proposed reduced-order model (ROM) procedure for a one-dimensional steady plug-flow reactor (PFR) and use it to investigate the performance of a membrane reactor (MR), for hydrogen separation from syngas that may be produced in an integrated gasification combined cycle (IGCC). In the proposed model, syngas (a feed stream) enters the membrane reactor from one side into a retentate zone, while nitrogen (a sweep stream) enters the membrane reactor from the opposite side into a neighbor permeate zone. The two zones are separated by permeable palladium membrane surfaces that are selectively permeable to hydrogen. After analyzing the hydrogen permeation profile in a base case (300 °C uniform temperature, 40 atm absolute retentate pressure, and 20 atm absolute permeate pressure), the temperature of the module, the retentate-side pressure, and the permeate-side pressure are varied individually and their influence on the permeation performance is investigated. In all the simulation cases, fixed targets of 95% hydrogen recovery and 40% mole-fraction of hydrogen at the permeate exit are demanded. The module length is allowed to change in order to satisfy these targets. Other dependent permeation-performance variables that are investigated include the logarithmic mean pressure-square-root difference, the hydrogen apparent permeance, and the efficiency factor of the hydrogen permeation. The contributions of our study are linked to the fields of membrane applications, hydrogen production, gasification, analytical modeling, and numerical analysis. In addition to the proposed reduced-order model for hydrogen separation, we present various linear and nonlinear regression models derived from the obtained results. This work gives general insights into hydrogen permeation via palladium membranes in a hydrogen membrane reactor (MR). For example, the temperature is the most effective factor to improve the permeation performance. Increasing the absolute retentate pressure from the base value of 40 atm to 120 atm results in a proportional gain in the permeated hydrogen mass flux, with about 0.05 kg/m2.h gained per 1 atm increase in the retentate pressure, while decreasing the absolute permeate pressure from the base value of 20 bar to 0.2 bar causes the hydrogen mass flux to increase exponentially from 1.15 kg/m2.h. to 5.11 kg/m2.h. This study is linked with the United Nations Sustainable Development Goal (SDG) numbers 7, 9, 11, and 13.

1. Introduction

1.1. General Overview

Gasification is a thermochemical process in which a carbonaceous solid fuel (such as coal or biomass) is converted into a synthesis gas, which is known as “syngas” [1,2,3], using a gasification agent, also called “gasifying reagent” or “gasification medium” (such as steam, oxygen, carbon dioxide, or a mixture of them) under a controlled gasification environment (such as the residence time, heating rate, pressure, and temperature) within a facility called a “gasifier” [4,5,6,7,8]. In other words, gasification is a partial oxidation (substoichiometric) process consisting of physical processes, such as pyrolysis, and chemical reactions, such as gasification using steam [9,10]. Syngas is a mixture of gases, primarily carbon monoxide (CO) and hydrogen (H2), with possible additional impurities such as carbon dioxide (CO2), methane (CH4), water vapor (H2O), and nitrogen (N2) [11]. Syngas is also produced from natural gas or light crude oil fractions through steam reforming [12]. The molar ratio of molecular hydrogen (H2) to carbon monoxide increases as the ratio of hydrogen atoms to carbon atoms in the feedstock increases. Thus, ideal steam reforming of one mole of carbon gives a molar ratio of H2:CO equal to 1:1, given the following reaction:
C + H 2 O H 2 + C O
which has one mole of hydrogen and one mole of carbon monoxide as the products. On the other hand, the ideal steam reforming of one mole of methane (as an approximation for natural gas) gives a molar ratio of H2:CO equal to 3:1, given the following reaction:
C H 4 + H 2 O 3 H 2 + C O
which has three moles of hydrogen and one mole of carbon monoxide as the products [13].
Syngas has multiple uses beyond being a gaseous fuel that can be used in gas turbines or boilers [14,15]. For example, syngas is also a feedstock for the production of methanol, ammonia, synthetic gasoline, and hydrogen [16,17] in alignment with SDG 9 “Industry, innovation and infrastructure” [18,19,20,21,22,23,24]. Syngas-based hydrogen, particularly, is of special environmental importance, as it enables the production of electricity through fuel cells without harmful carbon dioxide emissions [25,26]. This is aligned with the United Nation’s seventh Sustainable Development Goal (SDG) “Goal 7: Affordable and Clean Energy” [27,28,29,30,31,32]. Syngas-based hydrogen also can be used as an alternative fuel for sustainable mobility [33,34,35] in alignment with SDG 11 “Sustainable Cities and Communities” [36,37,38,39,40,41], leading to reduced greenhouse gas (GHG) emissions in the transportation sector and thus mitigating global warming effects [42,43,44] in alignment with SDG 13 “Climate Action” [45,46,47], and also leading to improved air quality in the built environment [48,49] in alignment with SDG 11 “Sustainable Cities and Communities” [50,51,52,53], since the combustion of hydrogen does not lead to carbon dioxide or toxic products [54,55] (but undesired nitrogen oxides or NOx “NO and NO2” are still possible due to the oxidizing air [56]). In the power sector, nitrogen-diluted syngas-based hydrogen can be used in hydrogen-fueled gas turbines, with no harmful carbon dioxide emissions [57,58]. This dilution allows for reducing the combustion temperatures, and thus reducing the NOx emissions [59]. A nitrogen-diluted hydrogen flame with a fuel mixture having a chemical composition of 60% nitrogen and 40% hydrogen (mole fractions) was experimentally studied, and it was shown that this nitrogen dilution is effective in adjusting the flame length [60]. Separating hydrogen from syngas increases the concentration of carbon dioxide in the remaining content [61], thereby facilitating the carbon capture process afterward [62,63], which is similar to how using pure oxygen (rather than air) as an oxidizer in oxy-fuel combustion [64,65,66,67] causes the combustion products to be rich in carbon dioxide and this makes its capture easier [68,69]. The impurity carbon dioxide can be captured from the syngas, preventing its release into the atmosphere [70,71]. The captured carbon dioxide may be stored underground and it may be utilized commercially to achieve economic gains in different applications [72,73,74]. This carbon capture scenario is classified as a pre-combustion capture because carbon dioxide is segregated from the fuel stream (syngas in this case) before it is sent to the combustion facility [75,76].
Hydrogen separation from a hydrogen-containing gas mixture can be accomplished by various methods, including membrane separation [77,78]. Palladium-based (Pd-based) membranes allow the production of high-purity hydrogen [79,80,81], which is important for the effective operation of proton exchange membrane (PEM) fuel cells [82,83,84,85,86]. Palladium-based membranes are known for selectively separating hydrogen from a gas stream [87], which undergoes a dehydrogenation process as hydrogen is removed and is made to pass across the membrane. Compared to pressure swing adsorption (PSA), where adsorbing the impurities is used for hydrogen separation, membrane-based hydrogen separation allows for higher hydrogen recovery [88]. The term “hydrogen recovery” means the fraction of hydrogen mass in a gaseous mixture that is extracted (separated) from that mixture acting as the source of hydrogen [89,90,91,92,93,94]. Thus, if all the hydrogen content in a mixture of gases is extracted, the hydrogen recovery reaches its highest value of 100% or 1.0. Pressure swing adsorption is a process used for separating a gas (such as hydrogen) from a mixture of gases, or for purifying a gas mixture by removing present gas impurities in it [95,96,97,98,99,100]. Although this is the same purpose as the palladium membrane in the current study, the mechanism is very different. In PSA, selective adsorption–desorption takes place. When PSA is used in separating or purifying hydrogen, a mixture of gases containing hydrogen is pressurized, and then it passes through a reactor with an adsorbent material in the form of solid particles, which selectively attract the molecules of the impurities (the gases other than hydrogen) from the mixture onto their surface, as an adsorption stage, leaving hydrogen atoms at a high level of purity. Then, the pressure is decreased, which causes bound molecules of the impurities (the adsorbed non-hydrogen gases) to be released from the adsorbent surface, as a desorption stage [101]. Compared to cryogenic distillation, where impurities are condensed at low temperatures for purifying hydrogen; membrane purification is less energy-intensive [102]. The cryogenic separation of hydrogen uses a refrigeration system to cool a gaseous mixture containing hydrogen [103]. Because the constituent gases of the mixture typically condense and change from a gaseous state into a liquid state at different temperatures (at different boiling points), non-hydrogen gases in the mixture condense first, leaving only the hydrogen in a purified gaseous form [104,105,106,107,108,109,110]. Among all elements, hydrogen has the second lowest boiling point of any pure substance, which is about 20.3 K (−252.85 °C) at normal atmospheric pressures [111,112,113,114], while helium has the lowest one, which is about 4.2 K (−268.95 °C) at a normal atmospheric pressure [115,116]. Similarly, hydrogen has the second lowest melting point of any element, while helium has the lowest melting point [117,118,119,120,121,122,123]. Membrane-based hydrogen purification here refers to a pressure-driven process, where the selective permeation of hydrogen through the palladium membrane is stimulated by a difference in the partial pressure of hydrogen as it is transferred from the retentate side (before the hydrogen separation, where hydrogen has a higher partial pressure) to the permeate side (after the hydrogen separation, where hydrogen has a lower partial pressure). It may be worth mentioning that the carbon capture process (such as pre-combustion capture as demonstrated in the current study, or post-combustion carbon capture and oxy-fuel carbon capture) is a retrofitting action to mitigate carbon dioxide (CO2) emissions from existing energy systems or other industrial facilities [124,125]. For better environmental aspects, a total switch to renewable energy sources (particularly solar energy [126,127] and wind energy [128,129]) and harvested energy [130,131] is preferred because, in such scenarios, no carbon dioxide is produced [132,133,134]. However, such an energy transition toward renewable and sustainable resources may not be readily easy for existing conventional energy plants, whereas carbon capture strategies may remain suitable solutions for low-emission energy and process systems [135].

1.2. The Objectives of the Study

This work considers metal membranes that are unsupported (bulk), relatively thick, dense (non-porous) palladium foils for hydrogen separation. Catalytic reactors with dense metallic membranes for hydrogen production have shown a potential to overcome limitations on the yield that are associated with chemical equilibrium [136]. The mechanism of permeation through the membrane involves the dissociation of molecular hydrogen at the surface of the palladium membrane as hydrogen atoms, which then diffuse into the palladium membrane [137]. An isothermal hydrogen separation module (thus, having a uniform temperature) is adopted here. The spatial variation in the chemical composition of the permeate stream and retentate stream is modeled numerically as a plug-flow reactor. The permeation performance is examined under a reference set of representative conditions, as well as when each of the three control (design) variables is changed from the reference (base) case. These control variables are as follows: (1) the reactor temperature; (2) the retentate-side pressure; (3) the permeate-side pressure. Our study is based on simplified system-level computational modeling using principles of membrane-based hydrogen permeation. The interested reader can find some information about modeling a continuous plug-flow reactor in Appendix A.

1.3. Novelty and Added Knowledge

The contributions of this work include the following: (1) the detailed step-by-step numerical model algorithm for membrane-based hydrogen permeation; (2) the regression models that qualitatively show the impact of three operational conditions on membrane-based hydrogen permeation; (3) the contrasting profiles of hydrogen recovery and mole fractions in an isothermal membrane reactor at a base case and three other extreme cases. Although reduced-order modeling (ROM) is a new concept, and it has been used in several fields of engineering and physics [138,139,140,141,142]; our study proposes a detailed reduced-order model for plug-flow reactors (PFRs) when viewed as membrane reactors for mass transfer. This unique model is helpful in computer-aided design (CAD) and computer-aided engineering (CAE) in important applications under chemical engineering (CAE) and industrial processes, as an early-stage analysis without formidable three-dimensional computational fluid dynamics (CFD) models [143,144,145,146,147] and expensive experimental testing [148,149]. For high-fidelity analysis, these sophisticated CFD models or experimental prototypes may be utilized while guided by the finding of our ROM. In addition, all the results and regression models presented here are novel, thereby expanding our knowledge about palladium-based hydrogen production.

1.4. The Structure of the Manuscript

In the next section, the research method is described. Then, details about the geometric and inlet parameters are provided. Most of these settings remain fixed throughout the entire study. Then, the modeling procedure is described for the hydrogen permeation in the reactor. After this, five quantitative scalar quantities are introduced as criteria for evaluating the overall permeation performance. This is followed by a presentation of results for a base case, combined with a discretization-sensitivity analysis that confirms the adequacy of the spatial resolution utilized. Then, the influence of three control variables is explored, with some regression models relating the value of each control variable (as an independent or explanatory variable) to the five permeation metrics of the hydrogen permeation (as dependent or response variables). Finally, concluding remarks are provided.

2. Research Method

The research conducted here is computational in principle, and it relies on the modeling of hypothetical units for the permeation of hydrogen from a feedstock (feed) gas stream to a destination stream called permeate.

2.1. Connection to Real-Life Application

The feed stream is representative of a realistic syngas flow in an integrated gasification combined cycle (IGCC), in which coal is converted into syngas (mainly carbon monoxide “CO”, molecular hydrogen “H2”, and carbon dioxide “CO2”) [150,151,152,153,154,155,156,157,158,159]. In the IGCC power plant, this syngas drives gas turbines as the first method of electric power generation [160,161,162,163,164,165,166,167,168]. Then, the heat content in the exhaust gases is partially recovered to produce steam as a working fluid for steam turbines that form the second method of electric power generation [169,170,171,172,173,174,175,176,177]. An IGCC power plant has an efficiency (with respect to the lower heating value, LHV) of about 48%, which means it has a higher efficiency (10–15% increase) and lower harmful emissions (10–15% reduction) compared to a conventional steam power plant operating by burning coal as a fuel [178,179,180,181].

2.2. Computer Modeling Tool

We performed the computational modeling through spreadsheets (Microsoft Excel software program, Version (2019)), where the mathematical equations governing the hydrogen permeation along the membrane length were implemented as dependent formulas. The built-in tool “Goal Seek” within the Microsoft Excel desktop application (under the Microsoft Windows operating system) was utilized for solving the nonlinear algebraic equation relating the unknown membrane length to the specified hydrogen recovery target (95%) [182,183]. Obtaining a solution using this nonlinear solver tool is nearly immediate on a personal computer. This is a big advantage compared to multi-dimensional detailed computational fluid dynamics (CFD) models, which not only take more time and require more computing power, but also require advanced skills and specialized software for setting the model and running a computer solver [184,185,186,187,188,189,190]. While such detailed CFD models allow for more realistic solutions (such as three-dimensional flow fields), for a preliminary design stage, a quick and convenient tool may be advantageous for exploring the approximated impact of the quantities of interest, particularly when only system-level operational parameters are desired. Such reduced-order modeling allows for rapid preliminary design, benchmarking, optimization and control, parametric sensitivity analysis, and investigating the validity of wide ranges of operating conditions [191,192,193,194].

2.3. Controllable Variables

The variables that are subject to change in this work as independent parameters are (1) the temperature of the membrane reactor (while keeping the retentate-side pressure and the permeate-side pressure at reference values of a base case), (2) the retentate-side pressure (while keeping the temperature and the permeate-side pressure at reference values of a base case), and (3) the permeate-side pressure (while keeping the temperature and the retentate-side pressure at reference values of a base case).
For each of the three scenarios listed above, the influence of the isolated design variable on the hydrogen permeation and fluid flow was investigated.

2.4. Simplifying Assumptions and Their Implications

Some assumptions were made in the present work, which reduces the interaction of factors and makes the interpretation of cause–effect dependence more evident. Such assumptions include the uniformity of the temperature and ignoring the effect of permeation on the pressures [195]. It is acknowledged that these assumptions limit the generality of the results and introduce deviations from real-world systems. However, the study focuses on trends and patterns of relational dependence between variables for a simplified configuration. Thus, the results presented here can be useful when regarded as broad guides rather than firm facts.
The assumption of constant temperature (isothermal environment) seems proper since the modeled dehydrogenation process does not involve internal heat sources, combustion, coolants, or external heat exchange [196].
The assumption of fixed total pressure in either the retentate zone or the permeate zone during the hydrogen transfer from the retentate zone to the permeate zone was relaxed given that it is the partial pressure of hydrogen (not the total pressure of the gas mixture) that effectively drives this mass transfer. The variations in the partial hydrogen pressure during the hydrogen transfer were captured in our simplified model through the variation in the hydrogen mole fraction.
The detailed steps of the numerical modeling procedure for the proposed palladium-based hydrogen separation are presented in Appendix B.

2.5. Flow Setup

Figure 1 is a sketch of the stream directions and the compositions of the feed (inlet of the retentate side, in the shell) and the sweep (inlet of the permeate side, inside the tubes). Instead of showing all eight tubes (which can make the sketch unclear), only the central tube is shown. The coordinate (x) is the longitudinal distance measured from the left edge of the membrane reactor. The feed syngas enters the shell from the left, and the sweep nitrogen enters the tubes from the right.
The counter-flow profile used here for the streams (where the two streams enter from the opposite ends of the membrane reactor) is preferred over the parallel or co-flow profile (when the two streams enter the membrane reactor from the same edge and exit from the same edge). This is due to a more even distribution of the driving force for the mass transfer of hydrogen from the retentate side to the permeate side. This advantage of counter-flow design over parallel flow (co-flow) design also exists in the field of heat exchangers, where heat transfer rather than mass transfer is sought [197,198,199].

3. General Model Settings

3.1. Fixing Common Parameters

This section illustrates various elements of the hydrogen separation modeling that remained unchanged for all simulation cases. They may be classified into two groups: (1) geometric parameters; (2) flow parameters. It should be noted that this work is not primarily intended to solve a particular problem or recommend a specific design. Instead, more value is provided by examining variations in permeation parameters when three operational variables are changed. Thus, any reasonable selection of settings was considered satisfactory here to establish a starting design point in the design space.

3.2. Underlying Geometry

All the imagined hydrogen separation modules investigated in this study share the same presumed geometric configuration except for the length. Such a proposed module is in the form of a shell-and-tube reactor, with eight cylindrical tubes placed inside a cylindrical shell. The sum of the cross-sectional areas of the tubes is 50% of the cross-sectional area of the outer enclosing cylinder, which forms the outer boundary of the shell. In the current work, the term “shell” means the part of the outer enclosing cylinder that remains after subtracting the tubes. Thus, it refers to the passage available for the gaseous stream to flow around the tubes but within the outer enclosing cylinder. This stream is referred to as the shell stream or the retentate. It is the fluid stream where the syngas feedstock enters and loses hydrogen during a dehydrogenation process to the other stream located inside the tubes. The retentate stream entering the shell is referred to as the feed, which means the raw syngas that is supplied before any processing by the hydrogen permeation membranes.
The other stream located inside the eight tubes is the tube-side stream or the permeate. It is the destination of the hydrogen that is transported from the retentate through the palladium membranes. At the inlet of the permeate, a non-hydrogen gas of molecular nitrogen (N2) is supplied in the current model. This intentionally added nitrogen (or other gas) is referred to as the sweep gas or the sweep [200,201,202,203,204,205], and this is common in membrane-based separation or distillation processes [206,207,208,209,210,211,212]. While the use of such sweep gas is, strictly speaking, optional, it helps improve the permeation process by ensuring that the permeate side can never be saturated with hydrogen, and it increases the difference in the hydrogen partial pressures across the membrane, leading to an increased hydrogen permeation flux [213,214,215,216]. The added sweep gas can play another function of controlling the temperature [217], but this is not considered here as we assume the membrane reactor to be isothermal for simplicity. Other gases can also be used as a sweep gas, such as argon, Ar [218,219,220,221]; steam, H2O [222,223,224,225]; helium, He [226,227,228,229]; or a mixture of gases [230,231,232,233,234,235]. The use of nitrogen here seems adequate [236,237].
The shell-and-tube reactor configuration was used in another study [238] while computationally modeling the propane steam reforming process under high temperatures (between 750 K and 900 K). In the propane steam reforming process, propane (C3H8) reacts with steam (H2O) to produce molecular hydrogen (H2) and carbon dioxide (CO2). This reformation reaction takes place within a porous medium acting as a catalyst. Unlike the present study, the model in that external study did not include hydrogen separation. The tubes in the reactor were used in that external study to heat the surrounding porous shell through passing hot gases.
Another computational study for propane steam reforming (as a method for producing hydrogen from propane) used the shell-and-tube configuration [239,240,241,242]. There is a difference between the modeled reactor in that study and the modeled reactor in the present work. The reactor in the external study was a reformer with a catalytic chemical reaction taking place inside it, while here it is a separator with selective permeance (selective mass transport) but without a chemical reaction between gaseous species. However, there are some similarities between the two studies. One similarity is that both modeling methods assumed a steady flow (no change with time). Also, the flow entering the tubes (heating gas) and the flow entering the shell (reactant propane and steam) in the other study flow in opposite directions longitudinally, which is the same situation here for the retentate stream and the permeating stream.
Outside the field of hydrogen generation or separation (purification), shell-and-tube reactors can be used for heat transfer processes, when the fluid in the tubes is used to heat or change the phase of the fluid in the shell (or vice versa) without direct contact. The heat transfer performance can be improved by adding baffles of different geometries that alter the path of the fluid in the shell and control its level of mixing through intensified turbulence [243,244,245]. The shell-and-tube configuration is also viable in multi-phase flows [246,247,248].
In other research work, a palladium-based membrane reactor of steam methane reforming (MR-SMR) was optimized with two objectives, which are (1) minimum heat exchange and (2) minimum power consumption—through reducing the pressure gradient that should be counteracted by an external compressor unit [249,250]. In that reported work, by controlling the exterior wall temperature, it was found possible to achieve a saving of 5.1% for the power consumption and a drop of 1.4% in the rate of heat exchange. The membrane reactor was tubular (a single tube inside a single tube). It was treated as a plug-flow reactor, given the high aspect ratio (large length compared to the diameter). The outer annular tube is the reaction zone (where methane reacts with steam to produce hydrogen), while the inner circular tube is the permeation zone (hydrogen permeates from the outer tube to the inner tube through a separating palladium membrane). The outer annular tube has a catalyst in the form of particles uniformly filling the outer annular tube, making it a porous ‘packed bed’. The palladium membrane was dense.
Figure 2 shows a proposed cross-section of the membrane reactor, indicating the areas of the eight tubes (yellow color) and the area of the shell (white color). To obtain more uniform gaps between walls, one of the tubes was placed exactly at the center of the shell. The remaining seven tubes were arranged with equal circumferential spacing. It should be mentioned that the plug-flow reactor modeling conducted here utilized the details about the layout of tubes when calculating the total perimeter of the palladium membranes (but the shell outer diameter was not utilized in the permeation calculation). In addition, such a layout provides justification for the adopted membrane perimeter, and also helps the reader envision how the membrane reactor looks in a real situation.
Table 1 shows some details about the shell-and-tube geometry. The tubes and shell are assumed to have zero thickness, and the palladium membranes cover the entire surface of each tube. Being a reduced-order one-dimensional model, there are no baffles in the membrane reactor [251,252,253].

3.3. Fixed Conditions

Table 2 shows some details about the retentate stream, whose inlet conditions correspond to the raw syngas (the feed) supplied to the membrane reactor. The temperature and pressure are considered uniform in that stream. The temperature and pressure values are not specified because they are not common in all simulation cases, but are varied and their influences are studied later in multiple simulations.
The mixture molecular weight of the feed (MWmix) is calculated as a mole-weighted average of the molecular weights of the constituent gases, according to [261,262,263,264,265,266,267].
M W m i x = M W H 2   X H 2 + M W C O   X C O + M W C O 2   X C O 2
where (MW) refers to the molecular weight, (X) refers to the mole fraction, and the subscripts refer to the individual gases.
The mass fraction (Y) for each constituent gas in the feed is dependent upon its mole fraction (X) according to [268].
Y i = M W i   X i M W m i x
where the subscript (i) refers to any gas of the constituent gases in the feed.
The standard volume flow rate [269,270,271,272,273] of a gas (as a compressible medium whose volume is considerably sensitive to pressure and temperature) is the virtual volume flow if the temperature and pressure are at specified reference standard values [274,275,276,277,278]. These standard conditions are taken here as Tstnd = 0 °C (32 °F, 273.15 K) and Pstnd (absolute pressure) = 105 Pa (0.9869 atm, 14.504 psi). These two standard values of temperature and pressure are based on the International Union of Pure and Applied Chemistry (IUPAC) [279,280,281,282,283,284,285].
An ideal gas is a gas that obeys the ideal-gas equation of state, which can be expressed as
P   V = N   R ¯   T
where (P) is the pressure, (V) is the volume, (N) is the number of moles, ( R ¯ ) is the universal gas constant, and (T) is the absolute temperature, expressed in kelvins. For an ideal gas, each mole occupies a volume of 22,711 cm3 (22.711 L) at the standard pressure and temperature mentioned earlier. Thus, the standard volume is directly proportional to the number of moles, which in turn (and assuming no change occurs in the chemical composition of the gas) is directly proportional to the mass. This value can be obtained from the ideal-gas equation of state by solving for the standard volume per mole (Vstnd/N) as follows:
V s t n d N = R ¯ T s n d P s t n d = 1 0 6 8.3145 J m o l . K 273.15   K 1 0 5   P a = 22,711 c m 3 m o l
where the multiplier (106) is inserted to convert the volume unit from cubic meters (m3) to cubic centimeters (cm3), and the value of the universal gas constant ( R ¯ ) is a physical constant taken from the U.S. National Institute of Standards and Technology [286]. The value shown in the above equation is a truncated version of the published one used in the computation, which is 8.314462618 J/mol.K.
Table 3 shows some details about the permeate stream, whose inlet conditions correspond to the sweep gas supplied to the membrane reactor. The temperature and pressure are considered uniform in that stream.
The flow rate of the sweep gas is decided based on the target mole fraction of hydrogen in the permeate outlet (XH2,per−out), the mass flow rate of hydrogen in the syngas (ṁH2,feed), and the target hydrogen recovery (β) according to
m ˙ s w e e p = M W N 2   ( 1 X H 2 , p e r o u t ) M W H 2   X H 2 , p e r o u t     β   m ˙ H 2 , f e e d
The above relation can be used under the condition that there is only nitrogen in the sweep gas (which is true in all the simulations in the present work). It is a special relation and should not be used in other configurations, such as when the sweep gas is a mixture.
In the present work, β = 0.95, XH2,per−out = 0.4, and ṁH2,feed = 1.5499 kg/h (this is the product of the feed mass flow of 60 kg/h, and the mass fraction of hydrogen in the feed of 0.025832). With the previously mentioned molecular weights of hydrogen (Table 2) and nitrogen (Table 3), we obtain ṁsweep = 30.692 kg/h as mentioned in Table 3.
The mass flow rates (and the standard volume flow rates) and the chemical compositions of the sweep gas and the feed gas are fixed in all the simulations of this work.

4. Hydrogen Permeation Metrics

After performing a complete steady-state simulation for the palladium membrane permeation of hydrogen in a segmented plug-flow reactor, the assessment of the overall permeation process is facilitated using a number of quantitative scalar quantities (referred to here as performance metrics or permeation metrics) that are convenient to utilize for comparing the performance of different simulations. The use of quantitative metrics is an effective way of concisely assessing and benchmarking the performance of two processes or systems, and they are adopted in many fields [288,289]. In this section, five performance metrics for hydrogen separation are discussed.

4.1. Membrane Length

The first permeation performance metric adopted here was the membrane length needed to achieve the target hydrogen recovery (β = 95%). This is designated by the symbol (L).
A smaller membrane length was desired, because it means a shorter membrane reactor, which is a better design due to reduced cost and space.

4.2. Average Hydrogen Permeation Mass Flux

The second permeation performance metric adopted here was the average mass flux (mass flow rate per unit area) of the permeated hydrogen through the membrane. It is designated by the symbol ( M ¯ H2). Since the local mass flux of the permeated hydrogen is subject to variation from one segment to another in the membrane reactor model, the average of all segmental values is used. This metric is related to the corrected (second-iteration) segment-level molar flux of permeating hydrogen through the palladium membrane (JH2,i) as
M ¯ H 2 = 3.6   M W H 2 1 n i = 1 n J H 2 , i  
where (MWH2 = 2.01588 kg/kmol) is the molecular weight of the molecular hydrogen, and the factor (3.6) appears to enable unit conversion such that the average mass flux is expressed in (kg/m2.h) when the segmental mole flux (JH2,i) is expressed in (mol/m2.s).
A higher average mass flux is desired, as it shows more intense utilization of the membrane surface (better use of each unit area).
In the present study, because the total permeated hydrogen is fixed (by fixing the inlet mass flow of hydrogen in the syngas feed and fixing the target hydrogen recovery), and the membrane perimeter is also fixed, the average hydrogen mass flux ( M ¯ H2) and the membrane length (L) are not independent. Instead, they are inversely proportional to each other, and their product should be invariant.

4.3. Log Mean Pressure-Square-Root Difference

The third permeation performance metric adopted here is a global (membrane-level, not a segment-level) pressure-square-root difference, which is a membrane-level difference in the hydrogen partial pressure raised to the power of 0.5, between the retentate stream (higher value) and the permeate stream (lower value). It is designated here by the symbol (LMPsrD), or simply (LMPD), and is called log (or logarithmic) mean pressure-square-root difference. This difference in the square root of the hydrogen partial pressure (PsrD) stimulates the permeation through the palladium membrane. Because this stimulus driving force can vary along the membrane segments, an average value was sought. Instead of a simple arithmetic average over all segments, a logarithm-based average was used, which takes into account the differences at the left end and at the right end of the entire membrane reactor (where the flow inlets and outlets are located). Unlike computing a simple average of the driving force (the difference in the square root of the hydrogen partial pressure, or PsrD) along the membrane, the LMPD is related to the change in this driving force between the two ends of the hydrogen separation modules only, and this makes this metric simpler to compute without the need to know the details of the driving force between the two side ends of the hydrogen separation module. This resembles an approach of calculating an overall temperature difference in the field of heat transfer within heat exchangers, which is called log mean temperature difference (or LMTD); and this is a representative temperature difference between the colder stream and the hotter stream [290,291,292,293,294,295,296]. The global log mean pressure-square-root difference is calculated as follows:
L M P D = Δ P H 2 0.5 L H S , 1 Δ P H 2 0.5 R H S , n l n Δ P H 2 0.5 L H S , 1 l n Δ P H 2 0.5 R H S , n  
where (ΔPH20.5)LHS,1 is the driving force at the left end of the membrane reactor (at x = 0), (ΔPH20.5)RHS,n is the driving force at the right end of the membrane reactor (at x = L), and the (ln) function is the natural logarithm. The LMPD should lie between (ΔPH20.5)LHS,1 and (ΔPH20.5)RHS,n, regardless of which of them is larger than the other.
A higher log mean pressure-square-root difference (LMPD) is desired, as it is an indicator of the average driving force for hydrogen permeation between the left end and the right end of the membrane reactor, and is interpreted in a similar way to the LMTD in heat exchanges, being the driving force for heat transfer [297].

4.4. Global Apparent Permeance

The fourth permeation performance metric adopted here is the global apparent permeance (k’app), which is a solution-dependent variable that represents the effective permeance based on the obtained profile of hydrogen permeation in the membrane reactor. It has the same unit of the actual (ideal) permeance (k’), which is an input parameter that depends on the temperature and the membrane length, and thus can be computed before the numerical simulation of the plug-flow reactor. To explain the usefulness and significance of the global apparent permeance (k’app), it should be noted that the actual permeance (k’) is a local value, corresponding to a specific segment (or even a specific point), and it depends on the temperature and the membrane thickness, as in Equation (A14). However, the apparent permeance is computed based on the overall performance of the membrane reactor as a whole, as follows:
k a p p = 1 n i = 1 n J H 2 , i L M P D  
Therefore, two dehydrogenation palladium membrane reactors may have the same actual permeance because they have the same uniform temperature and the same uniform thickness, but due to variations in the dehydrogenation performance, they can have different global apparent permeance values. This makes the global apparent permeance a good metric for assessing the dehydrogenation process.
The numerator in the above equation is the arithmetic average of the molar flux of permeating hydrogen through the palladium membrane (average of the n segmental values), while the denominator is the log mean pressure-square-root difference (LMPD).
A higher apparent permeance is desired, as it shows that the membrane reactor yields a higher useful output (hydrogen molar flux) for a given average input (hydrogen pressure driving force).

4.5. Efficiency Factor

The fifth and last permeation performance metric adopted here is the efficiency factor (ηk), which is simply the ratio of the global apparent permeance to the local actual permeance, expressed as a percentage. Therefore,
η k = k a p p k  
The efficiency factor is the only permeation metric presented here that is non-dimensional, which makes it unambiguous and identical in any system of units.
The “efficiency factor” is not strictly an efficiency as used in energy conversion by a heat engine or an energy conversion process [298,299,300,301,302,303], being a useful output energy that is a fraction of an input heat energy, and this case the efficiency is limited to an upper ceiling value of 100% [304]. Instead, the “efficiency factor” is a ratio between two quantities that have the same dimensional units but have different meanings, and either of them is allowed to exceed the other. Thus, the “efficiency factor” used here can exceed 100%. This is somehow similar to the concept of the Octane Number or the Research Octane Number (RON) as a quality scale for gasoline (petrol) automotive fuels, where this rating is not strictly bound by an upper limit of 100 [305,306,307,308,309,310,311,312,313,314,315].
A higher efficiency factor is desired, as it is directly proportional to the global apparent permeance; thus, it is related to the goodness of the membrane reactor operation with regard to hydrogen permeation.

5. Results

5.1. Base Case and Spatial Resolution Test

The first implementation of the plug-flow membrane reactor model is referred to as the base case or reference simulation. To run a simulation, three thermodynamic properties that describe the two flowing streams in the reactor are needed. These three properties are as follows: (1) temperature (assumed uniform in the entire reactor); (2) retentate pressure (assumed uniform in the retentate stream); (3) permeate pressure (assumed uniform in the permeate stream).
The uniform-temperature assumption avoids ambiguity when computing the actual (ideal) permeance of the palladium membrane, which is temperature-dependent. Thus, the actual permeance is also uniform in the entire membrane, and this eliminates an undesirable distracting influence from varying permeance.
The three aforementioned thermodynamic properties are to be varied individually later, deviating from their base values. The influence of each of them on the hydrogen permeation is examined.
The base case is considered to have representative (realistic) values of the membrane streams [316,317,318], which are summarized in Table 4.
The local actual hydrogen permeance (k’) for the base temperature of (300 °C, 573.15 K) is 10.263 × 10−4 mol/m2.s.Pa0.5.
Table 5 gives numerical results for the base case after a complete simulation of the segmented plug-flow reactor. The normal resolution of 200 segments is compared with another resolution with twice the number of segments (400 segments) having the same membrane length found necessary for the 200-segment case.
Also, the absolute percentage deviations of these results between the two resolutions are shown in the table. Such comparison reveals small deviations not exceeding 2% in all four permeation performance metrics other than the membrane length, which was forced to be identical in the two simulations. Namely, these metrics are (1) average hydrogen permeation mass flux, (2) log mean pressure-square-root difference, (3) global apparent hydrogen permeance, (4) efficiency factor.
For the hydrogen recovery, it was a specified input in the case of using 200 segments, where the membrane length was a computed output. In the case of using 400 segments, the membrane length was the specified input (to ensure geometric consistency between the two simulations), while the hydrogen recovery was computed from the segmented plug-flow reactor model, in an inverse simulation mode. There is an insignificant difference (0.01% absolute change) in the two values of the hydrogen recovery.
In addition to the demonstrated quantitative agreement between the normal-resolution simulation with 200 segments and the high-resolution validation simulation with 400 segments, the qualitative agreement can also be observed in Figure 3. This figure compares the cumulative hydrogen recovery along the base-case plug-flow reactor model, for the 200-segment resolution and the 400-segment resolution. The profiles under both resolutions are visually indistinguishable.
Figure 4 shows the segmental hydrogen recovery profile (contribution of each reactor segment to the total hydrogen recovery). The normal resolution (200 segments) corresponds to nearly twice the values of the high resolution (400 segments). Although this may indicate a mismatch between the two simulations, it is actually appropriate because each segment under the normal resolution has twice the membrane area of a segment under the high resolution. For a meaningful visual comparison, a third curve was added to the figure by doubling the segmental hydrogen recovery obtained under high resolution (because each data point, in this case, represents only half of what a data point represents in the case of normal resolution). When the added curve (the adjusted high-resolution curve) is compared with the normal resolution curve, it is seen that both curves are visually identical.
The nonlinear profile of the segmental recovery is attributed to the underlying variation in the driving force for the segmental hydrogen permeation (the segmental pressure-square-root difference), which is a nonlinear function of the hydrogen mole fraction in the permeate stream and the hydrogen mole fraction in the retentate stream.
The variations in the hydrogen mole fraction in the retentate stream as the retentate gas traverses the membrane reactor from the left to the right are shown in Figure 5. It decreases from 30% at the left end corresponding to x/L = 0 to 2.10% at the right end corresponding to x/L = 1. The variation is weakly nonlinear with the distance. The exit mole fraction is not exactly zero, but this is consistent with the target hydrogen recovery of 95%. The value of the exit mole fraction of hydrogen (2.10%) and the longitudinal pattern of that mole fraction for the normal resolution (200 segments) are consistent with those for the validation resolution (400 segments).
The variations in the hydrogen mole fraction in the permeate stream as the permeate gas traverses the membrane reactor from the right to the left are shown in Figure 6. It increases from 0% at the right end corresponding to x = L (a normalized variable x/L = 1 is used in the figure) to the target 40% at the left end corresponding to x = 0. The variation is weakly nonlinear with the distance, as in the case of the retentate stream. Comparing the mole fraction patterns with 200 segments and with 400 segments suggests independence of the results on the spatial resolution. Thus, using 200 segments is adequate and there is no need to use a higher resolution.
With such good observed agreement between the normal spatial resolutions and the high testing resolution, the choice of 200 segments (normal resolution) was considered satisfactory and was adopted in all remaining simulations.

5.2. Influence of Temperature

After completing the segmented flow reactor under a set of representative operational conditions, the temperature of the flows in the membrane reactor was varied in a discrete manner, leading to seven temperature values that are higher than the base value of 300 °C. The base value and the six additional temperatures are as follows:
  • 300 °C (base);
  • 350 °C;
  • 400 °C;
  • 500 °C;
  • 600 °C;
  • 700 °C;
  • 800 °C.
It is understood that high temperatures may pose practical challenges for the materials and the process feasibility. Despite this, the model allows for exploring the change in the permeation performance due to elevated temperatures. This is achieved by monitoring the changes in the five performance metrics due to the changes in the temperature alone. The other two design variables (retentate pressure and permeate pressure) are kept at their base values (40 atm absolute and 20 atm absolute, respectively).
Figure 7 shows the influence of the temperature on the first hydrogen permeation metric, which is the membrane length. The figure shows a monotonic nonlinear decline in the needed membrane length as the temperature increases. This dependence can be described by a nonlinear power-type regression model, which is also shown in the figure. In the equation shown within the figure, (x) refers to the temperature in degrees Celsius, T (°C), while (y) refers to the membrane length in meters, L (cm). Therefore,
L c m = 3.0364 × 1 0 6   T ( ° C ) 1.5676  
Thus, the temperature is an instrumental variable in limiting the membrane length (for a given target hydrogen recovery) to less than 25% of its base value. Thus, it is equally instrumental to increase the hydrogen recovery for a given geometric length.
Figure 8 shows the influence of the temperature on the second hydrogen permeation metric, which is the average hydrogen permeation mass flux. There is an almost linear increase with temperature, which can be described by the following linear regression model:
M ¯ H 2 ( k g / m 2 . h ) = 1.4491 + 0.0084087   T ( ° C )
with an (R2) value of 0.9990. This value that accompanies the linear regression equation is called the coefficient of determination. It helps in evaluating the goodness of the regression fit, with a perfect fit having a value of 1.0, which means that the data points lie exactly on a straight line [319,320,321,322,323,324,325,326,327,328]. Therefore, the provided regression model is almost perfect within the examined range of temperatures.
Thus, the temperature is an instrumental variable to improve the permeation flux. This is justified by the increase in the actual local permeance when the temperature increases, following an Arrhenius-type dependence, which actually appears in different fields other than the current application of hydrogen permeation (such as chemical reaction rates [329]).
The third hydrogen permeation metric, which is the logarithmic mean pressure-square-root difference (LMPD), maintains its base values due to fixed inlet and outlet pressures. Therefore, its implicit dependence on the temperature is not manifested. Similarly, the left-end pressure-square-root difference, (ΔPH20.5)LHS,1, and the right-end pressure-square-root difference, (ΔPH20.5)RHS,n remain the same when the temperature of the membrane reactor is varied while fixing the inlet and outlet pressures. Thus, they maintain their base values. Therefore, there is no need to visualize these three variables in a dedicated figure with the temperature being the independent variable (because they would simply appear as three straight horizontal lines in such a figure).
The lack of influence of temperature on this permeation metric is valid, due to fixing the pressures in all the cases while changing only the temperatures.
Figure 9 shows how the fourth and fifth hydrogen permeation metrics change as the temperature increases. These metrics are the global apparent permeance and the efficiency factor, respectively. In addition, the figure also includes the actual local permeance. While both the actual (local) permeance and the global apparent permeance increase nonlinearly with the temperature, their ratio remains constant at the base value of 67.09%. Because their ratio is, by definition, the efficiency factor, the efficiency factor remains constant at its base value when the operational temperature of the palladium membrane reactor module increases. The increase in the permeance values exactly follows an exponential function of the following form:
k   o r   k a p p m o l / m 2 . s . P a 0.5 e μ / ( T ° C + 273.15 )
where (μ) is a constant, being the activation energy divided by the universal gas constant (E/ R ¯ ).
Therefore, while increasing the temperature leads to better permeance, it is useless for improving the efficiency factor.

5.3. Influence of Retentate Pressure

After exploring the role of the temperature as a design variable to improve permeation in a membrane reactor, this subsection explores the similar role of the retentate pressure. The retentate pressure was varied in a discrete manner, leading to a number of pressures higher than the base value of 40 atm (absolute). The base value and seven additional absolute retentate pressures are as follows:
  • 40 atm (base);
  • 45 atm;
  • 50 atm;
  • 60 atm;
  • 70 atm;
  • 80 atm;
  • 100 atm;
  • 120 atm.
The goal here is to investigate the relation between this retentate pressure and the five performance metrics. The other two design variables (temperature and permeate pressure) are kept at their base values (300 °C and 20 atm absolute, respectively).
Figure 10 shows the influence of the retentate pressure on the first hydrogen permeation metric, which is the membrane length. Similarly to the observed relation with the temperature, there is a favorable nonlinear decline in the membrane length as the retentate pressure increases. This relation can be described by the following fifth-order polynomial:
L c m = 4.43331 × 1 0 5   r 4 0.015423   r 3 + 2.0356   r 2 119.74   r + 2812.2  
where (r) refers to the retentate absolute pressure, expressed in atm, PRet (atm, absolute). The retentate pressure was found to be an instrumental variable in limiting the membrane length (for a given target hydrogen recovery) to less than 25% of its base value at about 110 atm (absolute). In other words, it can be utilized to increase the hydrogen recovery for a given geometric length.
Figure 11 shows the influence of the retentate pressure on the second hydrogen permeation metric, which is the average hydrogen permeation mass flux. There is an approximate linear increase with pressure, which can be described by the following linear regression model:
M ¯ H 2 k g / m 2 . h = 0.75288 + 0.051720   P R e t ( a t m , a b s o l u t e )  
with an (R2) value of 0.9931, which is close to the perfect value of 1.0. Thus, the retentate pressure is useful for improving the permeation flux. This is justified by the increase in the driving force for permeation when the retentate pressure increases, due to the resulting proportional increase in the hydrogen partial pressure in the retentate stream.
Figure 12 shows that the third hydrogen permeation metric (log mean pressure-square-root difference—LMPD) increases logarithmically with the retentate pressure. The relation between the log mean pressure-square-root difference and the absolute pressure of the retentate stream over the range of considered pressures can be approximated as the following logarithmic function:
L M P D P a 0.5 = 442.40   l n P R e t ( a t m , a b s o l u t e ) 1412.6  
At an absolute retentate pressure of about 47 atm, the left-end pressure-square-root difference and the right-end pressure-square-root difference are equal, and become matching with the log mean pressure-square-root difference. The gap between the left-end and right-end pressure-square-root difference becomes larger as the retentate pressure increases.
Figure 13 shows how the global apparent permeance and the efficiency factor change as the retentate pressure increases. The actual local permeance is fixed at its base value of 10.263 × 10−4 mol/m2.s.Pa0.5, but the apparent global permeance increases slightly as the retentate pressure increases, and it approaches the actual local permeance at high retentate pressure. The efficiency factor increases from its base value of 67.09% at 40 atm (absolute) to 98.71% at 120 atm (absolute).

5.4. Influence of Permeate Pressure

In this subsection, the impact of the permeate pressure (the last design variable considered in this study) on the permeation metrics is discussed. Seven additional simulations with different values of the absolute permeate pressure were conducted with the aim of capturing the trend of the permeation performance when the permeate pressure decreased from the base value of 20 atm (absolute) to a small value near the absolute vacuum (zero atm absolute) [330,331,332]. The base value and seven additional absolute permeate pressures are as follows:
  • 20 atm (base);
  • 17.5 atm;
  • 15 atm;
  • 10 atm;
  • 5 atm;
  • 1 atm;
  • 0.5 atm;
  • 0.2 atm.
The other two design variables (temperature and retentate pressure) are kept at their base values (300 °C and 40 atm absolute, respectively). Unlike the temperature and the retentate pressure, where larger values than the base values were investigated, smaller values of permeate pressures are investigated here to seek better permeation, since this permeation pressure has a resistive effect for permeation (not a supportive effect as the temperature and the retentate pressure).
Figure 14 shows the influence of the permeate pressure on the first hydrogen permeation metric, which is the membrane length. There is a nonlinear (nearly exponential) increase in the membrane length as the retentate pressure increases. This relation can be described by the following regression model:
L c m = 93.105   e 0.07192 P P e r a t m , a b s o l u t e
Although manipulating the permeate pressure can reduce the membrane length (for a given target hydrogen recovery) to less than 25% of its base value (which was also achievable by manipulating the temperature or the retentate pressure), a partial vacuum is needed to achieve this, with pressures below the normal atmospheric pressure of 1 atm. This imposes complications and expense, and may not be acceptable. However, the reduction in the permeate pressure within more realistic values (like 10 atm, absolute) has still a favorable effect, and the membrane length can decrease to less than half of its base value by halving the absolute permeate pressure.
Figure 15 shows the influence of the permeate pressure on the second hydrogen permeation metric, which is the average hydrogen permeation mass flux. There is approximately an exponential decline in the average hydrogen permeation mass flux as the permeate pressure increases. This can be described by the following nonlinear regression model:
M ¯ H 2 k g / m 2 . h = 5.0339   e 0.07192 P P e r a t m , a b s o l u t e
If the absolute permeate pressure can be reduced to half of its base value, the average hydrogen permeation mass flux can be intensified to more than twice its base value. In the base case (with the base absolute permeate pressure of 20 bar), the average hydrogen permeation mass flux is 1.15054 kg/m2.h. This increases to 2.48568 kg/m2.h at an absolute permeate pressure of 10 bar (which is an increase by a multiplicative factor of 2.160 compared to the base case), and increases further to 5.10756 kg/m2.h at an absolute permeate pressure of 0.2 bar (which is an increase by a multiplicative factor of 4.439 compared to the base case).
Figure 16 shows that the third hydrogen permeation metric (log mean pressure-square-root difference) increases logarithmically with retentate pressure. The relation between the pressure-square-root (LMPD) and the absolute pressure of the permeate stream over the range of considered pressures can be approximated as an exponential function as
L M P D P a 0.5 = 571.05   e 0.04471   P P e r a t m , a b s o l u t e  
Below the absolute permeate pressure of about 17.5 atm, the left-end pressure-square-root difference is larger than the right-end pressure-square-root difference, which is reversed at higher permeate pressures.
Figure 17 shows how the global apparent permeance and the efficiency factor decrease as the permeate pressure increases. The actual local permeance is fixed at its base value of 10.263 × 10−4 mol/m2.s.Pa0.5, but the apparent global permeance decreases from a value higher than the actual local permeance (thus, an efficiency factor above 100%) at very low permeate pressures to a value smaller than the actual local permeance as the permeate pressure increases (thus, an efficiency factor below 100%) as the absolute permeate pressure is increases beyond about 5 atm. Specifically, the efficiency factor reaches 102.53% and 118.43% at absolute permeate pressures of 5 atm and 0.2 atm, respectively (compared to the base value of 67.09% at 20 atm, absolute).

5.5. Profiles with Extreme Design Variables

The final method to highlight the general impacts of the three design variables (temperature, retentate pressure, and permeate pressure) on the hydrogen permeation within a generic palladium membrane reactor is contrasting some longitudinal profiles when the membrane reactor operates at the extreme values (of the ranges considered in the current study) assigned to each design variable while keeping the other two design variables at their base values.
This means contrasting results from three different simulations with the following conditions in Table 6.
The cumulative hydrogen recovery profiles for the three extreme cases are contrasted in Figure 18. The left-end value (0%) and the right-end value (95%) are implicitly imposed by the enforced inlet or outlet conditions. Thus, all profiles should coincide at both the left and right edges of the membrane reactor, but they can deviate in the intermediate zone. The cumulative hydrogen recovery for the extreme temperature shows more linearity than the other two profiles, with the extreme permeate pressure corresponding to a curved profile, resembling a second-order function of the normalized membrane coordinate (x/L), with a gradually declining slope as x/L increases. The extreme retentate pressure simulation shows an intermediate behavior between the other two extreme cases, in terms of the cumulative hydrogen recovery profiles.
This behavior is better explained by visualizing the segmental contribution to the total hydrogen recovery for these three extreme cases in Figure 19. While the profile for the extreme temperature is not horizontal, it first decreases and then increases, making it behave as if changing around a uniform mean value. Thus, the cumulative value (integration of the profile) should be close to a straight upward line as shown earlier. On the other hand, the extreme permeate pressure causes a nearly linear declining profile for the segmental hydrogen recovery, which when integrated, gives a nearly quadratic profile as shown before. The extreme retentate pressure simulation shows an intermediate behavior between the other two extreme cases, as noted earlier.
The hydrogen mole fractions in the retentate stream and the permeate stream for the three extreme cases are contrasted in Figure 20 and Figure 21, respectively. The simulation with the extreme permeate pressure shows a regular variation for both streams. The simulation with the extreme temperature shows an accelerated decline in the hydrogen mole fractions in both streams as the coordinate x approaches the right end (permeate inlet, retentate exit). This can be attributed to the rapidly elevated segmental hydrogen recovery in this region, after steadily declining in the left part of the membrane reactor. Again, the extreme retentate pressure simulation shows an intermediate behavior between the other two extreme cases.

6. Discussion

This section addresses four topics. First, it briefly discusses how some of the presented components in this study can add new insight to the literature and fill a potential gap in it. Second, it addresses the question of whether or not the results are limited to ideal gases. Third, it comments on the type of validation used for the presented reduced-order model. Fourth, we provide a visual summary that compares the influence levels of the three design variables examined in this study.

6.1. The Contributions of the Study

The contribution of this work to the fields of energy systems, computational modeling, hydrogen production, and carbon capture includes presenting a simple plug-flow reactor computational model for the membrane-based hydrogen separation, which takes a short time to give rough predictions as a precursor of time-consuming three-dimensional computational fluid dynamics (CFD) models performed using specialized computer programs [333,334,335,336,337]. The simple plug-flow reactor model developed here can be automated using spreadsheet software without resorting to complicated computer programming or expensive simulation packages. The model was checked by comparing cases with different spatial resolutions. The model uses a predictor-corrector approach in the computation, where an initial value (in a predictor step) is refined (in a corrector step) to yield the final value at each segment of the segmented reactor. An apparent lack of such a model in the literature is one of the motivations for disseminating this study. The present study also demonstrates examples of consolidated metrics for comparing and judging the permeation performance of hydrogen. This can guide researchers when analyzing or interpreting similar problems. In addition, the study describes the impact of three different design variables on the hydrogen permeation performance, accompanied by good-fit regression models. This step helps in having a broad estimation of the advantage of manipulating each of these variables, which can be weighed against the expenses or practical difficulty in a realistic setting, thus helping in selecting optimum operational conditions. Aside from the technical aspect of the current study, the study may also be viewed as an educational asset for diversifying and enriching teaching content in various courses (modules), such as chemical engineering, material science, energy systems, industrial processes, or numerical methods [338,339,340,341,342]. The simplicity of our proposed reduced-order model (ROM) becomes particularly advantageous here, where students can be exposed to the covered modeling concepts without having a deep background in many subjects. The students can replicate the presented simulations to ensure their understanding of the problem and can even expand this work through innovative projects.

6.2. Ideal-Gas Law and Compressibility Factor

While inspecting the steps of the proposed numerical membrane-based permeation model for hydrogen, confusion may arise regarding any restriction of applicability to gases that obey the ideal-gas law (ideal-gas equation of state). This can occur because the proposed model has steps (steps e and j) where the molar flux of permeating hydrogen through the palladium membrane (with the unit mol/m2.s, which is moles per square meter per second) was expressed as a standard-temperature–pressure volume flow rate (with the unit scc/min or sccm, which is standard cubic centimeters per minute). In particular, in step e, the conversion constant (22,711 scc/mol) was explicitly utilized, and it was derived earlier in Equation (6) using the ideal-gas equation of state. However, such restriction does not exist. The standard cubic centimeter (scc) for a gas is simply another unit of the amount of substance, like the mole (mol) unit and the kilomole (kmol) unit. All the numerical values used to derive the conversion constant (22,711 scc/mol) represent universal constants themselves. These values are the universal gas constant ( R ¯ = 8.3145 J/mol.K), the units multiplier (106 cm3/m3), the standard absolute temperature (273.15 K), and the standard pressure (105 Pa). Therefore, the unit (scc) is always linearly related to the unit (mol) through a fixed multiplication factor that is independent of the type of gas, the operating temperature, and the operating pressure. The use of scc in the model (rather than mol for example) can be viewed as an arbitrary choice for compliance with others who use this unit for gaseous flow rates in academic, commercial, or governmental sectors [343,344,345,346], and for convenience due to dealing with larger numbers having many significant digits before the decimal point.
Some studies may adopt slightly different values for the standard temperature and standard pressure than those adopted here. For example, the pressure value of 1 atm (101,325 Pa) may be considered the standard pressure instead of 105 Pa [347,348,349,350,351]. In such a case, the conversion constant (22,711 scc/mol) derived here consequently should change slightly to 22,414 scc/mol (after multiplying 22,711 by 105 and dividing it by 101,325), but this does not invalidate the fact that the scc and the mol are two related units.
In relation to the aforementioned remark, Table 2 in the present study provides the inlet standard volume flow rate (in sccm) for the retentate stream. This value is what is needed as a boundary condition to start using the proposed numerical hydrogen permeation model. The same table also provides the equivalent mass flow rate (in kg/h and lbm/h). This mass flow rate is auxiliary information. It is not directly utilized in the proposed permeation model.
The deviation from the ideal gas condition (which is utilized in the current study) is expressed in terms of a non-dimensional number called compressibility factor or deviation factor, Z [352,353,354,355,356,357,358], defined as
Z = P ρ   R   T = P   v R   T  
where (P) is the absolute pressure, (ρ) is the density, (R) is the specific gas constant, (T) is the absolute temperature, and (v) is the specific volume (the reciprocal of the density).
For ideal gases, the compressibility factor has the value of Z = 1 [359,360,361,362,363,364,365,366]. The more the deviation from the unity compressibility factor, the more the error resulting from using the ideal-gas law. The compressibility factor depends on the type of the gas, its temperature, and its pressure. Each gas has a characteristic temperature called the critical temperature or Tc [367,368,369,370,371,372], above which a vapor-liquid phase change becomes no longer observable. When the temperature of the gas is well above its critical point, this gas generally does not deviate significantly from being ideal, over a wide range of pressure [373,374]. All the simulations presented in the current study are at relatively high temperatures (from 300 °C to 800 °C, or 573.15 K to 1073.15 K), which are much more than the critical temperature of all the four gases (H2, CO, CO2, N2) included in all the simulations in the current study. The critical temperature of these gases ranges from 33.2 K for H2 to 304.2 K for CO2 [375,376,377,378,379]. Therefore, these gases are not expected to deviate largely from the ideal gas behavior in the performed simulations. For example, experiments showed that the compressibility factor for hydrogen at 75 °C (348.15 K) varies from 1.0024 at an absolute pressure of 4.5400 atm to 1.0455 at an absolute pressure of 85.269 atm [380]. As another example to support the ideal-gas assumption adopted in the current study, reported calculations based on the Reference Fluid Thermodynamic and Transport Properties Database or (REFPROP) [381,382,383,384] of the United States National Institute of Standards and Technology (NIST) [385,386,387,388,389] showed that the compressibility factor for hydrogen at 125 °C (398.15 K) varies from 1.0005 at an absolute pressure of 0.9869 atm to 1.0481 at an absolute pressure of 98.69 atm [390].

6.3. Self-Validation

In the current study, we presented self-validation by comparing results from our proposed reduced-order model (ROM) at a normal spatial resolution with results from the same ROM but at a higher spatial resolution. The outcome of this validation was very promising and satisfactory. Although further validation through comparisons with experimental results or three-dimensional computational fluid dynamics (CFD) tools may appear to add value to our study, we do not incorporate such high-fidelity validations. This should not harm the objectives of the study, which were fully accomplished. The study still presents a novel ROM for a membrane-type plug-flow reactor (PFR) as a preliminary stage before proceeding with a more-detailed computational modeling or an experimental prototype. Our ROM is not intended to replace such a more detailed analysis. Rather, it is aimed to assist it by providing a quick and convenient pre-analysis step that does not require specialized CFD software or prohibitive laboratory work (either of them can be extremely costly, time-consuming, and available only to institutionally backed researchers). In addition, our ROM is largely built upon first principles in chemistry and physics rather than unproven or complex concepts. Thus, the performed self-validation is considered sufficient for supporting the success of our ROM algorithm. We point out that our ROM is favorably made simple and easy to implement, but this also means it lacks the ability to capture some features that appear in a high-fidelity analysis. Such features include heat transfer effects, eddy structures, and temporal unsteadiness [391,392,393]. On the other hand, a three-dimensional CFD simulation is expected to have a lot of uncertainty given the large number of options and user-defined parameters that need to be selected during the simulation. Added to the different geometric and physical phenomena involved in a CFD model but not involved in our ROM, it is felt that a comparison with such CFD results is not a critical element of this study and can be skipped without major harm.

6.4. Priority Design Chart

Based on the findings of our study, the temperature of the membrane reactor for the palladium-based hydrogen separation is the most influential design variable affecting the hydrogen permeation performance, while the permeate pressure is the least influential, and the retentate pressure has an intermediate level of importance. We depict this variation among the three control variables in Figure 22, taking the shape of an inverted triangle. The relative importance of each variable in this figure is designated by the font size as well as by the area it occupies within the triangle. We refer to this sketch as a “priority design chart”.

7. Conclusions

A discrete plug-flow reactor model was established for simulating the one-dimensional isothermal permeation of hydrogen through palladium membranes, and was applied for separating hydrogen from a pressurized hot syngas stream. The model assumed a shell-and-tube cross-section for estimating the exact membrane area. A set of inlet conditions for the feed syngas and an inlet sweep nitrogen were established. Also, a target hydrogen recovery of 95% and a target hydrogen mole fraction of 40% at the permeate exit were enforced. The required membrane length was computed to achieve these targets. The model consisted of 200 segments, with each segment having a length of about 2 cm. After a validity test, it was found that this spatial segmentation was adequate.
Then, the impact of three operational variables was investigated, when varied individually starting from the reference values such that better hydrogen permeation can be obtained. These design variables are the temperature, retentate-side pressure, and permeate-side pressure. For quantitative evaluation of the hydrogen permeation improvement, five assessment metrics were discussed, including the apparent global permeance and the efficiency factor. Linear or nonlinear regression models were provided for the membrane length, the average permeation mass flux, and the log mean pressure-square-root difference. In addition, the one-dimensional profiles of some permeation-related quantities (such as the segmental hydrogen recovery) at the extreme value of each design variable were contrasted. The present study can serve as a rough guide for palladium-based hydrogen separation, showing designers some potential gains in hydrogen permeation by manipulating some operational conditions. The designers may then seek an optimum compromise between the expected gain and the incurred complications or implementation costs.
Considering the relative changes in the three investigated design variables and the dependent hydrogen permeation metrics in 21 simulations (one reference simulation and 20 additional ones), temperature is the most effective variable for improving the permeation performance. The retentate pressure comes in second place. The permeate pressure is the least powerful way to improve hydrogen permeation.

Funding

This research received no funding.

Data Availability Statement

Data elements and parameters used in this study are described and contained within the manuscript itself.

Conflicts of Interest

The author declares no conflicts of interest.

Appendix A. Continuous Plug-Flow Reactor Modeling

This appendix gives a brief description of modeling a continuous plug-flow reactor (PFR).
The Center for Chemical Process Safety (CCPS) within the American Institute of Chemical Engineers (AIChE) defines a plug-flow reactor (PFR) as “A tubular reactor where the feed is continuously introduced at one end and the products continuously removed from the other end. The concentration/temperature in the reactor is not uniform” [394]. This definition is commended for being generic enough to cover different varieties of the PFR model, based on specific sets of assumptions implied. It can be added that in a plug-flow reactor, the plug-like flow of substance does not have any mixing along the axial direction (no axial mixing between different ‘plugs’ of material) [395].
If the axial (longitudinal) axis is denoted by ( x ), the concentration of the ith species is denoted by ( C i ), the volumetric reaction rate is denoted by ( r ), the linear velocity is denoted by ( u ), and the stoichiometric coefficient of the ith species is denoted by ( ν i ), then the following partial differential equation governs the temporal and spatial evolution of the concentration along the PFR [396]:
C i t + C i x = 1 u   ν i   r  
If the flow is assumed to be steady, the above partial differential equation (PDE) is reduced to an ordinary differential equation (ODE):
d C i d x = 1 u   ν i   r  
Furthermore, if the reaction is first-order, which means that the reaction rate depends linearly on the concentration of only one reactant [397], then the axial gradient of the concentration ( C ) of that reacting species in the PFR becomes
d C d x = 1 u   k   C  
where ( k ) is the rate coefficient of the reaction, which can depend on the absolute temperature ( T ) following an exponential profile [398], such that
k = k 0   e ε   T  
where ( k 0 ) and ( ε ) are constants.

Appendix B. Modeling Hydrogen Permeation

Appendix B.1. Segmental Plug-Flow Reactor

Hydrogen permeation through a palladium membrane is a nonlinear phenomenon [399], which can be understood and investigated through numerical simulations rather than simple symbolic expressions [400]. For simulating the hydrogen permeation in the simplified hydrogen membrane reactor (HMR), piece-wise constant profiles are utilized to replace the continuous distributions along the membrane reactor, and no partial differential equations or sources of unsteadiness are involved. The membrane reactor is divided into (n) segments, having a small thickness and a constant cross-section. Each segment is assumed to have a uniform gas composition in either the permeate stream or in the retentate stream, and a uniform permeation flux. The flow is steady (time-independent), with a feed syngas entering the reactor from the left end (and leaving the reactor from the right end after losing 95% of its hydrogen moles), and a sweep gas entering the reactor from the right end (and leaving the reactor from the left end after gaining the same amount of hydrogen lost from the syngas, reaching a composition of 40% hydrogen and 60% nitrogen, by mole or volume). Such a segmental approach for handling the one-dimensional evolution of the permeation process along the membrane reactor resembles a discretized version of a plug-flow reactor, PFR [401]. Instead of having infinitely small segments (or plugs) in a true PFR, these segments here are finite. Also, instead of solving ordinary differential equations to find the composition distribution as a function of the longitudinal coordinate as in true PFR problems [402], a numerical approach based on a nonlinear mathematical framework involving scalar algebraic equations only, without a need to solve vector equations or to process matrices [403] is used to describe that composition in adjacent segments, by ensuring continuity of the permeate flow and the retentate flow at the interfaces of each pair of adjacent segments. The hydrogen permeation occurs through the palladium membranes from the retentate to the permeate, with the permeation rate approximated as being uniform (having the same value) within each segment, but it can change from one segment to another. In other words, the smoothly changing permeation rate is discretized and approximated as a piecewise-constant function. Similarly, the driving force for hydrogen permeation (which stems from the partial pressures of hydrogen in the retentate and the permeate sides) is also approximated as a piecewise-constant function.

Appendix B.2. Modeling Algorithm

Figure A1 explains the segmentation of the hypothetical hydrogen membrane reactor (HMR). It shows a portion of its left end, with the two most-left segments. The left end of the membrane reactor is where the coordinate (x) is assigned the value zero. Because the segmental plug-flow reactor (PFR) model is aimed to capture the hydrogen permeation with no consideration of the flow pattern, only the membrane surface is of concern here, because it is where permeation occurs. No gradients are allowed in the direction perpendicular to the longitudinal axis (thus, no gradients perpendicular to the membrane surface). The membrane is simplified in the figure as a cylindrical surface, separating the permeate stream (inside) from the retentate stream (outside).
The left-hand side (LHS) of the first segment is the left end of the entire membrane (where x = 0).
The LHS of the other segments (all segments except the first one) is the right-hand side (RHS) of the previous segment located to its left.
Figure A1. An illustration of the segments of the membrane reactor near its left end.
Figure A1. An illustration of the segments of the membrane reactor near its left end.
Processes 13 01455 g0a1
The index number (i) for a segment can take the value from (1) for the first segment at the left edge of the membrane reactor to (n) for the last segment at the right edge of the membrane reactor. The algorithm for calculating the segment-wise composition in the permeate and the retentate streams for a generic segment is provided below.
(a)
Start with a known hydrogen mole fraction in the retentate at the LHS (XH2,Ret-LHS,i), standard volume flow rate of retentate at the LHS (QRet-LHS,i), hydrogen mole fraction of permeate at the LHS (XH2,Per-LHS,i), and standard volume flow rate of permeate at the LHS (QPer-LHS,i) of the segment (say segment number i).
For this step, there are two routes of calculations.
For the first segment (i = 1), implement the inlet conditions and the target mole fraction as follows:
X H 2 , R e t L H S , 1 = 30 %   ( s p e c i f i e d   i n l e t   c o n d i t i o n )
Q R e t L H S , 1 = 970,068   s c c m   ( s p e c i f i e d   i n l e t   c o n d i t i o n )
Q H 2 , R e t L H S , 1 = X H 2 , R e t L H S , 1   Q R e t L H S , 1
In the last equation, the product (XH2,Ret-LHS,1 QRet-LHS,1) gives the standard volume flow rate of the hydrogen content in the feed syngas.
X H 2 , P e r L H S , 1 = 40 %   ( s p e c i f i e d   t a r g e t )
Q P e r L H S , 1 = β     Q H 2 , R e t L H S , 1 X H 2 , P e r L H S , 1
In the last equation, multiplying the standard volume flow rate of the hydrogen content in the feed syngas (QH2,Ret-LHS,1) by the target hydrogen recovery (β) gives the target standard volume flow rate of the hydrogen content in the permeate stream as it exits the membrane reactor from its left end. Dividing this value further by the target mole fraction of the exiting permeate stream (XH2,Per-LHS,1) gives the target standard volume flow of the exiting permeate stream (QPer-LHS,1), which is composed of molecular hydrogen and molecular nitrogen. Performing these calculations gives QPer-LHS,1 = 691,173 sccm.
For other segments except the first one (i = 2, 3, … n), the connectivity condition of the segments (interfacing condition) can be used as follows:
X H 2 , R e t L H S , i = X H 2 , R e t R H S , i 1
Q R e t L H S , i = Q R e t R H S , i 1
Q H 2 , R e t L H S , i = Q H 2 , R e t R H S , i 1
X H 2 , P e r L H S , i = X H 2 , P e r R H S , i 1
Q P e r L H S , i = Q P e r R H S , i 1
where the values of (XH2,Ret-RHS,i−1), (QRet-RHS,i−1), (QH2,Ret-RHS,i−1), (XH2,Per-RHS,i−1), and (QRet-RHS,i−1) should be available from the analysis of the previous segment (numbered i − 1).
(b)
Compute (QH2,Per-LHS,i), which is the standard volume flow rate of the hydrogen content in the permeate stream at the LHS of the current segment being analyzed (say segment i), as follows:
Q H 2 , P e r L H S , i = X H 2 , P e r L H S , i   Q P e r L H S , i
(c)
Compute (ΔPH20.5)LHS,i, which is the difference in the partial pressures of hydrogen raised to the power of 0.5 (this difference is the driving force for hydrogen permeation through the palladium membrane) at the LHS of the current segment being analyzed (say segment i), as follows:
Δ P H 2 0.5 L H S , i = X H 2 , R e t L H S , i   P R e t 0.5 X H 2 , P e r L H S , i   P P e r 0.5
where (PRet) is the absolute pressure of the retentate stream, and (PPer) is the absolute pressure of the permeate stream. Since the retentate stream is assumed to have a constant pressure, there is no need to add details (a subscript index) about the location for its value to be used in the above equation. The same reason justifies not specifying a particular location for the permeate pressure.
(d)
Compute (JH2,i), which is a predicted (first-iteration) segment-level molar flux of permeating hydrogen through the palladium membrane based on the conditions at LHS of the current segment being analyzed (say segment i), as follows:
J H 2 , i = k Δ P H 2 0.5 L H S , i
This predicted flux value is an initial estimation, based on LHS conditions only. In a subsequent step of the present algorithm, it is refined by including the effects of RHS conditions of the segment.
The above equation is referred to as the Richardson equation [404,405]. It is based on applying both Fick’s law for diffusion of hydrogen atoms across the metal membrane, and Sieverts’ law relating the concentration of the hydrogen atoms in the metal membrane (the equilibrium between gaseous hydrogen molecules next to the metal membrane and dissociated hydrogen atoms in the metal membrane) to the square root of the partial pressure of the adjacent molecular hydrogen gas [406,407,408]. The factor (k’) is an ideal (or local, or actual) permeance for hydrogen permeation. It is the product of the Fick’s diffusion coefficient (for the hydrogen atoms’ diffusion across the metal membrane) and the Sieverts’ solubility constant (the Sieverts’ law constant for the dissociation of a H2 molecule into two H atoms), and divided by the thickness of the metal membrane. In the current model, the factor (k’) is calculated as
k = A δ e E / ( R ¯ T )
where (A) is a pre-exponential factor for hydrogen permeation, (δ) is the thickness of the palladium membrane, (E) is an activation energy for hydrogen permeation, and (T) is the absolute temperature (in kelvins). The values of (A) and (E) used here are [409,410].
A = 2.2 × 10−7 mol/m/s/Pa0.5
E = 15,670 J/mol
with R ¯ = 8.3145 J/mol.K, the value of (E/ R ¯ ) in the above equation becomes 1884.7 K.
The thickness of the palladium membrane is set to 80 μm, which is considered a reasonable value [411,412].
(e)
Convert the LHS-based first-iteration molar flux (JH2,i) to a predicted (first-iteration) segment-level standard volume flow rate of permeating hydrogen (QH2,i) for the current segment being analyzed (say segment i).
This temporary standard volume flow rate value is an initial estimation, based on LHS conditions only. In a subsequent step of the present algorithm, it is refined by including the effect of RHS conditions. It is computed as follows:
Q H 2 , i = λ   J H 2 , i   A s e g
where (Aseg) is the membrane surface area in one segment, and (λ) is a constant that arises from a necessary unit conversion, as follows:
λ = 22,711 s c c m o l   60 s m i n
where (scc) stands for (standard cubic centimeters). This gives λ = 1.36266 × 106 scc.s/mol.min.
When computing the segmental membrane area, the envisioned tube layout in the shell-and-tube design and its specified dimensions become necessary.
If the tube diameter is designated by the symbol (d), and the length of the entire membrane reactor is designated by the symbol (L), then for eight tubes and (n) segments, the length of a single segment is
Δ x = L n
and the membrane area within a single segment is
A s e g = 8 Δ x   π   d
where (π = 3.14159) is the traditional mathematical constant.
(f)
Compute (XH2,Ret-RHS,i) and (XH2,Per-RHS,i), which are predicted (first-iteration) mole fractions of hydrogen in the retentate stream and the permeate stream, respectively, at the RHS of the current segment being analyzed (say segment i), as follows:
X H 2 , R e t R H S , i = Q H 2 , R e t L H S , i Q H 2 , i Q R e t L H S , i Q H 2 , i
X H 2 , P e r R H S , i = Q H 2 , P e r L H S , i Q H 2 , i Q P e r L H S , i Q H 2 , i
(g)
Compute (ΔPH20.5)RHS,i, which is the difference in the partial pressures of hydrogen raised to the power of 0.5 (as the driving force for hydrogen permeation) at the RHS of the current segment being analyzed (say segment i), as follows:
Δ P H 2 0.5 R H S , i = X H 2 , R e t R H S , i   P R e t 0.5 X H 2 , P e r R H S , i   P P e r 0.5
(h)
Compute (ΔPH20.5)i, which is the difference in the partial pressures of hydrogen raised to the power of 0.5 assigned to the current segment being analyzed (say segment i). It is taken as the arithmetic average of the LHS value and the RHS value, as follows:
Δ P H 2 0.5 i = 0.5 Δ P H 2 0.5 L H S , i + Δ P H 2 0.5 R H S , i
(i)
Compute (JH2,i), which is a corrected (second-iteration) segment-level molar flux of permeating hydrogen through the palladium membrane, which includes driving forces for permeation at both sides of the current segment being analyzed (say segment i), as follows:
J H 2 , i = k Δ P H 2 0.5 i
(j)
Convert the corrected, segment-level molar flux (JH2,i) to a corresponding updated (refined) segment-level standard volume flow rate of permeating hydrogen (QH2,i) for the current segment being analyzed (say segment i), as follows:
Q H 2 , i = λ J H 2 , i   A s e g
This is considered the final representation of the segment-level permeation of hydrogen.
(k)
Compute (RH2,i), which is the hydrogen recovery due to the current segment being analyzed (say segment i), as follows:
R H 2 , i = Q H 2 , i / Q H 2 , R e t L H S , 1
The denominator (QH2,Ret-LHS,1) in the previous equation is basically the standard volume flow rate of hydrogen in the inlet feed syngas. Thus, it is the standard volume flow of hydrogen available for permeation.
The segmental contribution to the overall recovery (RH2,i) can help in studying the distribution of the permeation, and thus identifying portions of the membrane reactor that are more effective than others.
(l)
Optional: Compute ( R ^ H 2 , i ), which is the cumulative hydrogen recovery, due to all previous segments of the membrane reactor in addition to the current segment being analyzed (say segment i), as follows:
R ^ H 2 , i = m = 1 i R H 2 , i
While this value is not necessary for continuing the calculation process, it is part of the data visualization of results here. It is thus beneficial to explain how it is obtained.
(m)
Compute (QRet-RHS,i) and (QH2,Ret-RHS,i), which are the standard volume flow rate of the retentate stream and the hydrogen content in that retentate stream, respectively, at the RHS of the current segment being analyzed (say segment i), as follows:
Q R e t R H S , i = Q R e t L H S , i Q H 2 , i
Q H 2 , R e t R H S , i = Q H 2 , R e t L H S , i Q H 2 , i
(n)
Compute (XH2,Ret-RHS,i), which is the corrected (second-iteration) mole fraction of hydrogen in the retentate stream at the RHS of the current segment being analyzed (say segment i), as follows:
X H 2 , R e t R H S , i = Q H 2 , R e t R H S , i / Q R e t R H S , i
(o)
Compute (QPer-RHS,i) and (QH2,Per-RHS,i), which are the standard volume flow rate of the permeate stream and the hydrogen content in that permeate stream, respectively, at the RHS of the current segment being analyzed (say segment i), as follows:
Q P e r R H S , i = Q P e r L H S , i Q H 2 , i
Q H 2 , P e r R H S , i = Q H 2 , P e r L H S , i Q H 2 , i
(p)
Compute (XH2,Per-RHS,i), which is the corrected (second-iteration) mole fraction of hydrogen in the permeate stream at the RHS of the current segment being analyzed (say segment i), as follows:
X H 2 , P e r R H S , i = Q H 2 , P e r R H S , i / Q P e r R H S , i
(q)
Set the obtained RHS conditions of the current segment being analyzed (say segment i) as LHS conditions at the next adjacent segment to be analyzed (segment i + 1), and repeat the computation procedure sequentially for all remaining segments until the last membrane segment (segment n).
The following values should be obtained for each segment:
  • (ΔPH20.5)LHS,i;
  • JH2,i;
  • QH2,i;
  • XH2,Ret-RHS,i and XH2,Per-RHS,i;
  • (ΔPH20.5)RHS,i;
  • (ΔPH20.5)i;
  • JH2,i;
  • QH2,i;
  • RH2,i;
  • Optional: R ^ H 2 , i
  • QRet-RHS,i and QH2,Ret-RHS,i;
  • XH2,Ret-RHS,i;
  • QPer-RHS,i and QH2,Per-RHS,i;
  • XH2,Per-RHS,i.
(r)
Compute ( R ^ H 2 , n ), which is the cumulative hydrogen recovery at the last segment. It is the overall hydrogen recovery by the entire membrane reactor, and it is obtained by simply adding the segment-level hydrogen recovery (RH2,i) for all the (n) segments of the membrane reactor. The total cumulative value is the target hydrogen recovery (β). Therefore,
β   o r   R ^ H 2 , n = m = 1 n R H 2 , i
The hydrogen recovery is an important success criterion not only from the chemical perspective, but also from an economic perspective. A higher hydrogen recovery leads to a lower cost per unit mass of hydrogen produced [413,414].
In the presented algorithm, the length of the entire membrane (L) is a nonlinear function of the overall hydrogen recovery (β). Since (β) is fixed at a desired value of 95% (in all the simulation cases of this study), a matching membrane length (L) should be obtained. Solving such a nonlinear system of equations is achieved here using the Goal Seek tool in the Microsoft Excel software program Version (2019). This tool is useful only when the equation or system of equations relates a single independent variable to a single dependent variable [415]. This condition is satisfied in the segmented plug-flow reactor problems here, with the independent variable being the membrane length, and the dependent variable being the hydrogen recovery.

References

  1. Bell, D.A.; Towler, B.F.; Fan, M. Coal Gasification and Its Applications; William Andrew/Elsevier: Oxford, UK, 2011; ISBN 978-1-4377-7851-9. [Google Scholar]
  2. Marzouk, O.A.; Huckaby, E.D. Assessment of Syngas Kinetic Models for the Prediction of a Turbulent Nonpremixed Flame. In Proceedings of the Fall Meeting of the Eastern States Section of the Combustion Institute 2009, College Park, ML, USA, 18–21 October 2009; pp. 726–751. [Google Scholar] [CrossRef]
  3. Couto, N.; Rouboa, A.; Silva, V.; Monteiro, E.; Bouziane, K. Influence of the Biomass Gasification Processes on the Final Composition of Syngas. Energy Procedia 2013, 36, 596–606. [Google Scholar] [CrossRef]
  4. Żogała, A. Equilibrium Simulations of Coal Gasification—Factors Affecting Syngas Composition. J. Sustain. Min. 2014, 13, 30–38. [Google Scholar] [CrossRef]
  5. Wagner, N.J.; Coertzen, M.; Matjie, R.H.; van Dyk, J.C. Chapter 5—Coal Gasification. In Applied Coal Petrology; Suárez-Ruiz, I., Crelling, J.C., Eds.; Elsevier: Burlington, NJ, USA, 2008; pp. 119–144. ISBN 978-0-08-045051-3. [Google Scholar]
  6. Guo, X.; Tang, Y.; Wang, Y.; Eble, C.F.; Finkelman, R.B.; Li, P. Evaluation of Carbon Forms and Elements Composition in Coal Gasification Solid Residues and Their Potential Utilization from a View of Coal Geology. Waste Manag. 2020, 114, 287–298. [Google Scholar] [CrossRef]
  7. Wang, Z.; Yang, J.; Li, Z.; Xiang, Y. Syngas Composition Study. Front. Energy Power Eng. China 2009, 3, 369–372. [Google Scholar] [CrossRef]
  8. Al-Zareer, M.; Dincer, I.; Rosen, M.A. Influence of Selected Gasification Parameters on Syngas Composition from Biomass Gasification. J. Energy Resour. Technol. 2018, 140, 041803. [Google Scholar] [CrossRef]
  9. Jayanarasimhan, A.; Pathak, R.M.; Shivapuji, A.M.; Rao, L. Tar Formation in Gasification Systems: A Holistic Review of Remediation Approaches and Removal Methods. ACS Omega 2024, 9, 2060–2079. [Google Scholar] [CrossRef]
  10. NETL, [National Energy Technology Laboratory of the United States Department of Energy]. Reactions & Transformations. Available online: https://netl.doe.gov/research/coal/energy-systems/gasification/gasifipedia/reaction-transformations (accessed on 5 May 2022).
  11. Frolov, S.M.; Panin, K.S.; Smetanyuk, V.A. Gasification of Liquid Hydrocarbon Waste by the Ultra-Superheated Mixture of Steam and Carbon Dioxide: A Thermodynamic Study. Energies 2024, 17, 2126. [Google Scholar] [CrossRef]
  12. Talmadge, M.; Biddy, M.; Dutta, A.; Jones, S.; Meyer, A. Syngas Upgrading to Hydrocarbon Fuels Technology Pathway; National Renewable Energy Laboratory (NREL), and Pacific Northwest National Laboratory (PNNL) of the United States Department of Energy (DoE): Richland, WA, USA, 2013.
  13. Lin, S.; Harada, M.; Suzuki, Y.; Hatano, H. Hydrogen Production from Coal by Separating Carbon Dioxide during Gasification. Fuel 2002, 81, 2079–2085. [Google Scholar] [CrossRef]
  14. Vow (ETIA SAS)|Syngas-High Energy Pyrolysis Gas for Power Applications. Available online: https://www.biogreen-energy.com/syngas (accessed on 6 May 2025).
  15. Suxing, S.; Yu, X.; Li, J.; Liu, X.; Sui, L.; Zhang, J.; Fu, Z.; Shao, Y. Effects of Hydrogen Addition on the Thermal Performance and Emissions of Biomass Syngas Combustion in a Horizontal Boiler. Energies 2024, 17, 2632. [Google Scholar] [CrossRef]
  16. Marzouk, O.A. Portrait of the Decarbonization and Renewables Penetration in Oman’s Energy Mix, Motivated by Oman’s National Green Hydrogen Plan. Energies 2024, 17, 4769. [Google Scholar] [CrossRef]
  17. Timilsina, M.S.; Chaudhary, Y.; Shah, A.K.; Lohani, S.P.; Bhandari, R.; Uprety, B. Syngas Composition Analysis for Waste to Methanol Production: Techno-Economic Assessment Using Machine Learning and Aspen Plus. Renew. Energy 2024, 228, 120574. [Google Scholar] [CrossRef]
  18. Costa, E. Industry 5.0 and SDG 9: A Symbiotic Dance towards Sustainable Transformation. Sustain. Earth Rev. 2024, 7, 4. [Google Scholar] [CrossRef]
  19. Marzouk, O.A.; Nayfeh, A.H. Mitigation of Ship Motion Using Passive and Active Anti-Roll Tanks. In Proceedings of the ASME 2007 International Design Engineering Technical Conferences and Computers and Information in Engineering Conference (IDETC-CIE 2007), ASME [American Society of Mechanical Engineers]. Las Vegas, NV, USA, 20 May 2009; pp. 215–229. [Google Scholar] [CrossRef]
  20. TGG, [The Global Goals] SDG 9|Goal 9: Industry, Innovation and Infrastructure. Available online: https://globalgoals.org/goals/9-industry-innovation-and-infrastructure (accessed on 3 May 2025).
  21. Marzouk, O.A.; Nayfeh, A.H. Control of Ship Roll Using Passive and Active Anti-Roll Tanks. Ocean Eng. 2009, 36, 661–671. [Google Scholar] [CrossRef]
  22. Kynčlová, P.; Upadhyaya, S.; Nice, T. Composite Index as a Measure on Achieving Sustainable Development Goal 9 (SDG-9) Industry-Related Targets: The SDG-9 Index. Appl. Energy 2020, 265, 114755. [Google Scholar] [CrossRef]
  23. Marzouk, O.A. Summary of the 2023 (1st Edition) Report of TCEP (Tracking Clean Energy Progress) by the International Energy Agency (IEA), and Proposed Process for Computing a Single Aggregate Rating. E3S Web Conf. 2025, 601, 00048. [Google Scholar] [CrossRef]
  24. Küfeoğlu, S. (Ed.) SDG-9: Industry, Innovation and Infrastructure. In Emerging Technologies: Value Creation for Sustainable Development; Springer International Publishing: Cham, Switzerland, 2022; pp. 349–369. ISBN 978-3-031-07127-0. [Google Scholar]
  25. Cavo, M.; Rivarolo, M.; Gini, L.; Magistri, L. An Advanced Control Method for Fuel Cells—Metal Hydrides Thermal Management on the First Italian Hydrogen Propulsion Ship. Int. J. Hydrogen Energy 2023, 48, 20923–20934. [Google Scholar] [CrossRef]
  26. Marzouk, O.A. Expectations for the Role of Hydrogen and Its Derivatives in Different Sectors through Analysis of the Four Energy Scenarios: IEA-STEPS, IEA-NZE, IRENA-PES, and IRENA-1.5 °C. Energies 2024, 17, 646. [Google Scholar] [CrossRef]
  27. Villavicencio Calzadilla, P.; Mauger, R. The UN’s New Sustainable Development Agenda and Renewable Energy: The Challenge to Reach SDG7 While Achieving Energy Justice. J. Energy Nat. Resour. Law 2018, 36, 233–254. [Google Scholar] [CrossRef]
  28. Marzouk, O.A. Power Density and Thermochemical Properties of Hydrogen Magnetohydrodynamic (H2MHD) Generators at Different Pressures, Seed Types, Seed Levels, and Oxidizers. Hydrogen 2025, 6, 31. [Google Scholar] [CrossRef]
  29. Estevão, J.; Lopes, J.D. SDG7 and Renewable Energy Consumption: The Influence of Energy Sources. Technol. Forecast. Soc. Change 2024, 198, 123004. [Google Scholar] [CrossRef]
  30. Marzouk, O.A. Jatropha Curcas as Marginal Land Development Crop in the Sultanate of Oman for Producing Biodiesel, Biogas, Biobriquettes, Animal Feed, and Organic Fertilizer. Rev. Agric. Sci. 2020, 8, 109–123. [Google Scholar] [CrossRef] [PubMed]
  31. Marcillo-Delgado, J.C.; Ortego, M.I.; Pérez-Foguet, A. A Compositional Approach for Modelling SDG7 Indicators: Case Study Applied to Electricity Access. Renew. Sustain. Energy Rev. 2019, 107, 388–398. [Google Scholar] [CrossRef]
  32. Marzouk, O.A. Facilitating Digital Analysis and Exploration in Solar Energy Science and Technology through Free Computer Applications. Eng. Proc. 2022, 31, 75. [Google Scholar] [CrossRef]
  33. Halder, P.; Babaie, M.; Salek, F.; Haque, N.; Savage, R.; Stevanovic, S.; Bodisco, T.A.; Zare, A. Advancements in Hydrogen Production, Storage, Distribution and Refuelling for a Sustainable Transport Sector: Hydrogen Fuel Cell Vehicles. Int. J. Hydrogen Energy 2024, 52, 973–1004. [Google Scholar] [CrossRef]
  34. Marzouk, O.A. Growth in the Worldwide Stock of E-Mobility Vehicles (by Technology and by Transport Mode) and the Worldwide Stock of Hydrogen Refueling Stations and Electric Charging Points between 2020 and 2022. In Construction Materials and Their Processing; Key Engineering Materials; Trans Tech Publications Ltd.: Bäch, Switzerland, 2023; Volume 972, pp. 89–96. ISBN 978-3-0364-0498-1. [Google Scholar] [CrossRef]
  35. Jayakumar, A.; Madheswaran, D.K.; Kannan, A.M.; Sureshvaran, U.; Sathish, J. Can Hydrogen Be the Sustainable Fuel for Mobility in India in the Global Context? Int. J. Hydrogen Energy 2022, 47, 33571–33596. [Google Scholar] [CrossRef]
  36. Mahadevia, D.; Mukhopadhyay, C.; Lathia, S.; Gounder, K. The Role of Urban Transport in Delivering Sustainable Development Goal 11: Learning from Two Indian Cities. Heliyon 2023, 9, e19453. [Google Scholar] [CrossRef] [PubMed]
  37. Marzouk, O.A. Toward More Sustainable Transportation: Green Vehicle Metrics for 2023 and 2024 Model Years. In Intelligent Sustainable Systems; Nagar, A.K., Jat, D.S., Mishra, D.K., Joshi, A., Eds.; Lecture Notes in Networks and Systems; Springer Nature: Singapore, 2024; Volume 817, pp. 261–272. ISBN 978-981-9978-85-4. [Google Scholar] [CrossRef]
  38. DESA, [United Nations Department of Economic and Social Affairs]. Sustainable Transport (SDG 11). Available online: https://sdgs.un.org/topics/sustainable-transport (accessed on 3 May 2025).
  39. Marzouk, O.A. Recommended LEED-Compliant Cars, SUVs, Vans, Pickup Trucks, Station Wagons, and Two Seaters for Smart Cities Based on the Environmental Damage Index (EDX) and Green Score. In Innovations in Smart Cities Applications Volume 7; Ben Ahmed, M., Boudhir, A.A., El Meouche, R., Karaș, İ.R., Eds.; Lecture Notes in Networks and Systems; Springer Nature: Cham, Switzerland, 2024; Volume 906, pp. 123–135. ISBN 978-3-031-53823-0. [Google Scholar] [CrossRef]
  40. Franco, S. Sustainable Cities and Communities: The Road towards SDG 11. TeMA—J. Land Use Mobil. Environ. 2022, 15, 341–344. [Google Scholar] [CrossRef]
  41. Mulibana, L.; du Toit, J. Transport Planning Research Toward Implementing SDG 11 in South Africa. In SDGs in Africa and the Middle East Region; Leal Filho, W., Abubakar, I.R., da Silva, I., Pretorius, R., Tarabieh, K., Eds.; Springer International Publishing: Cham, Switzerland, 2022; pp. 1–24. ISBN 978-3-030-91260-4. [Google Scholar]
  42. Marzouk, O.A. Zero Carbon Ready Metrics for a Single-Family Home in the Sultanate of Oman Based on EDGE Certification System for Green Buildings. Sustainability 2023, 15, 13856. [Google Scholar] [CrossRef]
  43. Liu, H.; Lu, H.; Hu, H. CO2 Capture and Mineral Storage: State of the Art and Future Challenges. Renew. Sustain. Energy Rev. 2024, 189, 113908. [Google Scholar] [CrossRef]
  44. Marzouk, O.A. Assessment of Global Warming in Al Buraimi, Sultanate of Oman Based on Statistical Analysis of NASA POWER Data over 39 Years, and Testing the Reliability of NASA POWER against Meteorological Measurements. Heliyon 2021, 7, e06625. [Google Scholar] [CrossRef]
  45. Doni, F.; Gasperini, A.; Soares, J.T. What Is the SDG 13? In SDG13—Climate Action: Combating Climate Change and Its Impacts; Emerald Publishing Limited: Leeds, UK, 2020; pp. 21–30. ISBN 978-1-78756-915-7. [Google Scholar]
  46. Marzouk, O.A. Evolution of the (Energy and Atmosphere) Credit Category in the LEED Green Buildings Rating System for (Building Design and Construction: New Construction), from Version 4.0 to Version 4.1. J. Infrastruct. Policy Dev. 2024, 8, 5306. [Google Scholar] [CrossRef]
  47. TGG, [The Global Goals] SDG 13|Goal 13: Climate Action. Available online: https://globalgoals.org/goals/13-climate-action (accessed on 3 May 2025).
  48. Bakirci, M. Smart City Air Quality Management through Leveraging Drones for Precision Monitoring. Sustain. Cities Soc. 2024, 106, 105390. [Google Scholar] [CrossRef]
  49. Marzouk, O.A. Compilation of Smart Cities Attributes and Quantitative Identification of Mismatch in Rankings. J. Eng. 2022, 2022, 5981551. [Google Scholar] [CrossRef]
  50. Akuraju, V.; Pradhan, P.; Haase, D.; Kropp, J.P.; Rybski, D. Relating SDG11 Indicators and Urban Scaling—An Exploratory Study. Sustain. Cities Soc. 2020, 52, 101853. [Google Scholar] [CrossRef]
  51. Marzouk, O.A. Urban Air Mobility and Flying Cars: Overview, Examples, Prospects, Drawbacks, and Solutions. Open Eng. 2022, 12, 662–679. [Google Scholar] [CrossRef]
  52. Abastante, F.; Lami, I.M.; Gaballo, M. Pursuing the SDG11 Targets: The Role of the Sustainability Protocols. Sustainability 2021, 13, 3858. [Google Scholar] [CrossRef]
  53. Marzouk, O.A. Aerial E-Mobility Perspective: Anticipated Designs and Operational Capabilities of eVTOL Urban Air Mobility (UAM) Aircraft. Edelweiss Appl. Sci. Technol. 2025, 9, 413–442. [Google Scholar] [CrossRef]
  54. Marzouk, O.A. Adiabatic Flame Temperatures for Oxy-Methane, Oxy-Hydrogen, Air-Methane, and Air-Hydrogen Stoichiometric Combustion Using the NASA CEARUN Tool, GRI-Mech 3.0 Reaction Mechanism, and Cantera Python Package. Eng. Technol. Appl. Sci. Res. 2023, 13, 11437–11444. [Google Scholar] [CrossRef]
  55. Jamrozik, A.; Grab-Rogaliński, K.; Tutak, W. Hydrogen Effects on Combustion Stability, Performance and Emission of Diesel Engine. Int. J. Hydrogen Energy 2020, 45, 19936–19947. [Google Scholar] [CrossRef]
  56. Xie, T.; Wang, P. Analysis of NO Formation in Counter-Flow Premixed Hydrogen-Air Flame. Trans. Can. Soc. Mech. Eng. 2013, 37, 851–859. [Google Scholar] [CrossRef]
  57. Marzouk, O.A. Hydrogen Utilization as a Plasma Source for Magnetohydrodynamic Direct Power Extraction (MHD-DPE). IEEE Access 2024, 12, 167088–167107. [Google Scholar] [CrossRef]
  58. Goldmeer, J.; Catillaz, J.; Donohue, J. Hydrogen as a Fuel for Gas Turbines—A Pathway to Lower CO2; General Electric (GE): Boston, MA, USA, 2022. [Google Scholar]
  59. Javoy, S.; Mevel, R.; Paillard, C.E. A Study of N2O Decomposition Rate Constant at High Temperature: Application to the Reduction of Nitrous Oxide by Hydrogen. Int. J. Chem. Kinet. 2009, 41, 357–375. [Google Scholar] [CrossRef]
  60. Weiland, N.T.; Strakey, P.A. Global NOx Measurements in Turbulent Nitrogen-Diluted Hydrogen Jet Flames; National Energy Technology Laboratory (NETL): San Diego, CA, USA, 2007.
  61. Marzouk, O.A. Dataset of Total Emissivity for CO2, H2O, and H2O-CO2 Mixtures; over a Temperature Range of 300–2900 K and a Pressure-Pathlength Range of 0.01–50 Atm.m. Data Brief 2025, 59, 111428. [Google Scholar] [CrossRef] [PubMed]
  62. Ghiat, I.; Al-Ansari, T. A Review of Carbon Capture and Utilisation as a CO2 Abatement Opportunity within the EWF Nexus. J. CO2 Util. 2021, 45, 101432. [Google Scholar] [CrossRef]
  63. Madejski, P.; Chmiel, K.; Subramanian, N.; Kuś, T. Methods and Techniques for CO2 Capture: Review of Potential Solutions and Applications in Modern Energy Technologies. Energies 2022, 15, 887. [Google Scholar] [CrossRef]
  64. Ruan, R.; Wang, G.; Li, S.; Wang, M.; Lin, H.; Tan, H.; Wang, X.; Liu, F. The Effect of Alkali and Alkaline Earth Metals (AAEMs) on Combustion and PM Formation during Oxy-Fuel Combustion of Coal Rich in AAEMs. Energy 2024, 293, 130695. [Google Scholar] [CrossRef]
  65. Marzouk, O.A. Radiant Heat Transfer in Nitrogen-Free Combustion Environments. Int. J. Nonlinear Sci. Numer. Simul. 2018, 19, 175–188. [Google Scholar] [CrossRef]
  66. Rajabloo, T.; Valee, J.; Marenne, Y.; Coppens, L.; De Ceuninck, W. Carbon Capture and Utilization for Industrial Applications. Energy Rep. 2023, 9, 111–116. [Google Scholar] [CrossRef]
  67. Marzouk, O.A. Temperature-Dependent Functions of the Electron–Neutral Momentum Transfer Collision Cross Sections of Selected Combustion Plasma Species. Appl. Sci. 2023, 13, 11282. [Google Scholar] [CrossRef]
  68. Donskoy, I. Techno-Economic Efficiency Estimation of Promising Integrated Oxyfuel Gasification Combined-Cycle Power Plants with Carbon Capture. Clean Technol. 2023, 5, 215–232. [Google Scholar] [CrossRef]
  69. Marzouk, O.A. Detailed and Simplified Plasma Models in Combined-Cycle Magnetohydrodynamic Power Systems. Int. J. Adv. Appl. Sci. 2023, 10, 96–108. [Google Scholar] [CrossRef]
  70. Dinca, C.; Slavu, N.; Cormoş, C.-C.; Badea, A. CO2 Capture from Syngas Generated by a Biomass Gasification Power Plant with Chemical Absorption Process. Energy 2018, 149, 925–936. [Google Scholar] [CrossRef]
  71. Marzouk, O.A. Technical Review of Radiative-Property Modeling Approaches for Gray and Nongray Radiation, and a Recommended Optimized WSGGM for CO2/H2O-Enriched Gases. Results Eng. 2025, 25, 103923. [Google Scholar] [CrossRef]
  72. Bhavsar, A.; Hingar, D.; Ostwal, S.; Thakkar, I.; Jadeja, S.; Shah, M. The Current Scope and Stand of Carbon Capture Storage and Utilization ∼ A Comprehensive Review. Case Stud. Chem. Environ. Eng. 2023, 8, 100368. [Google Scholar] [CrossRef]
  73. Dziejarski, B.; Krzyżyńska, R.; Andersson, K. Current Status of Carbon Capture, Utilization, and Storage Technologies in the Global Economy: A Survey of Technical Assessment. Fuel 2023, 342, 127776. [Google Scholar] [CrossRef]
  74. Tomasko, D.L.; Li, H.; Liu, D.; Han, X.; Wingert, M.J.; Lee, L.J.; Koelling, K.W. A Review of CO2 Applications in the Processing of Polymers. Ind. Eng. Chem. Res. 2003, 42, 6431–6456. [Google Scholar] [CrossRef]
  75. Singh, A.; Saini, V.; Jain, S.; Gour, A. Techno-Economic, Environmental, and Policy Perspectives of Carbon Capture to Fuel Technologies. Int. Res. J. Adv. Eng. Hub. 2024, 2, 1387–1403. [Google Scholar] [CrossRef]
  76. Bukar, A.M.; Asif, M. Technology Readiness Level Assessment of Carbon Capture and Storage Technologies. Renew. Sustain. Energy Rev. 2024, 200, 114578. [Google Scholar] [CrossRef]
  77. Shabbani, H.J.K.; Othman, M.R.; Al- Janabi, S.K.; Barron, A.R.; Helwani, Z. H2 Purification Employing Pressure Swing Adsorption Process: Parametric and Bibliometric Review. Int. J. Hydrogen Energy 2024, 50, 674–699. [Google Scholar] [CrossRef]
  78. Singh, A.P.; Singh, S.; Ganguly, S.; Patwardhan, A.V. Steam Reforming of Methane and Methanol in Simulated Macro & Micro-Scale Membrane Reactors: Selective Separation of Hydrogen for Optimum Conversion. J. Nat. Gas Sci. Eng. 2014, 18, 286–295. [Google Scholar] [CrossRef]
  79. Kassi, A.H.; Al-Hattab, T.A. A Review: Membrane Reactor for Hydrogen Production: Modeling and Simulation. Eng. Chem. 2023, 4, 17–31. [Google Scholar] [CrossRef]
  80. Omidifar, M.; Akbar Babaluo, A.; Jamshidi, S. H2 Permeance and Surface Characterization of a Thin (2 Μm) Pd-Ni Composite Membrane Prepared by Electroless Plating. Chem. Eng. Sci. 2024, 283, 119370. [Google Scholar] [CrossRef]
  81. Zito, P.F.; Brunetti, A.; Barbieri, G. Hydrogen Concentration and Purification by Membrane Process: A Multistage Analysis. Renew. Energy 2023, 218, 119243. [Google Scholar] [CrossRef]
  82. Piemonte, V.; Di Paola, L.; De Falco, M.; Iulianelli, A.; Basile, A. Hydrogen Production Using Inorganic Membrane Reactors. In Advances in Hydrogen Production, Storage and Distribution; Elsevier: Amsterdam, The Netherlands, 2014; pp. 283–316. ISBN 978-0-85709-768-2. [Google Scholar]
  83. Li, Y.; Yuan, H.; Wang, X.; Wei, X.; Zhu, J.; Ma, R.; Gao, L.; Dai, H. Insight into Resistance Relaxation Characteristics of Proton Exchange Membrane Fuel Cell in Purge Operation: Experiment, Application, and Mechanisms. J. Clean. Prod. 2025, 496, 145118. [Google Scholar] [CrossRef]
  84. Marzouk, O.A. 2030 Ambitions for Hydrogen, Clean Hydrogen, and Green Hydrogen. Eng. Proc. 2023, 56, 14. [Google Scholar] [CrossRef]
  85. Huang, T.; Zou, P.; Xiao, Y.; Ding, W.; Ke, R.; Zhang, Y.; Wan, Z.; Tu, Z.; Tang, S.; Zeng, W. Comprehensive Analysis of Half-Wave Continuous Cooling Channels in Performance of Proton Exchange Membrane Fuel Cells. Int. J. Heat Mass Transf. 2025, 244, 126971. [Google Scholar] [CrossRef]
  86. Irshad, H.M.; Shahgaldi, S. Comprehensive Review of Bipolar Plates for Proton Exchange Membrane Fuel Cells with a Focus on Materials, Processing Methods and Characteristics. Int. J. Hydrogen Energy 2025, 111, 462–487. [Google Scholar] [CrossRef]
  87. Peters, T.; Caravella, A. Pd-Based Membranes: Overview and Perspectives. Membranes 2019, 9, 25. [Google Scholar] [CrossRef]
  88. Marcantonio, V.; De Falco, M.; Capocelli, M.; Bocci, E.; Colantoni, A.; Villarini, M. Process Analysis of Hydrogen Production from Biomass Gasification in Fluidized Bed Reactor with Different Separation Systems. Int. J. Hydrogen Energy 2019, 44, 10350–10360. [Google Scholar] [CrossRef]
  89. Park, S.-J.; Lee, D.-W.; Yu, C.-Y.; Lee, K.-Y.; Lee, K.-H. Hydrogen Production from a DME Reforming-Membrane Reactor Using Stainless Steel-Supported Knudsen Membranes with High Permeability. J. Membr. Sci. 2008, 318, 123–128. [Google Scholar] [CrossRef]
  90. Lee, D.-W.; Park, S.-J.; Yu, C.-Y.; Ihm, S.-K.; Lee, K.-H. Remarkable Improvement in Hydrogen Recovery and Reaction Efficiency of a Methanol Reforming−Membrane Reactor by Using a Novel Knudsen Membrane. Ind. Eng. Chem. Res. 2008, 47, 1392–1397. [Google Scholar] [CrossRef]
  91. Lee, D.-W.; Nam, S.-E.; Sea, B.; Ihm, S.-K.; Lee, K.-H. Preparation of Pt-Loaded Hydrogen Selective Membranes for Methanol Reforming. Catal. Today 2006, 118, 198–204. [Google Scholar] [CrossRef]
  92. Barg, C.; Ferreira, J.M.P.; Trierweiler, J.O.; Secchi, A.R. Simulation and Optimization of an Industrial PSA Unit. Braz. J. Chem. Eng. 2000, 17, 695–704. [Google Scholar] [CrossRef]
  93. Damle, A.S. Hydrogen Production by Reforming of Liquid Hydrocarbons in a Membrane Reactor for Portable Power Generation—Model Simulations. J. Power Sources 2008, 180, 516–529. [Google Scholar] [CrossRef]
  94. Lee, D.-W.; Park, S.-J.; Yu, C.-Y.; Ihm, S.-K.; Lee, K.-H. Study on Methanol Reforming–Inorganic Membrane Reactors Combined with Water–Gas Shift Reaction and Relationship between Membrane Performance and Methanol Conversion. J. Membr. Sci. 2008, 316, 63–72. [Google Scholar] [CrossRef]
  95. Sircar, S. Pressure Swing Adsorption. Ind. Eng. Chem. Res. 2002, 41, 1389–1392. [Google Scholar] [CrossRef]
  96. Ruthven, D.M.; Farooq, S.; Knaebel, K.S. Pressure Swing Adsorption; John Wiley & Sons: Hoboken, NJ, USA, 1996; ISBN 978-0-471-18818-6. [Google Scholar]
  97. Zhang, R.; Shen, Y.; Tang, Z.; Li, W.; Zhang, D. A Review of Numerical Research on the Pressure Swing Adsorption Process. Processes 2022, 10, 812. [Google Scholar] [CrossRef]
  98. Wiheeb, A.D.; Helwani, Z.; Kim, J.; Othman, M.R. Pressure Swing Adsorption Technologies for Carbon Dioxide Capture. Sep. Purif. Rev. 2016, 45, 108–121. [Google Scholar] [CrossRef]
  99. Ko, D.; Siriwardane, R.; Biegler, L.T. Optimization of Pressure Swing Adsorption and Fractionated Vacuum Pressure Swing Adsorption Processes for CO2 Capture. Ind. Eng. Chem. Res. 2005, 44, 8084–8094. [Google Scholar] [CrossRef]
  100. Grande, C.A. Advances in Pressure Swing Adsorption for Gas Separation. Int. Sch. Res. Not. 2012, 2012, 982934. [Google Scholar] [CrossRef]
  101. Liemberger, W.; Groß, M.; Miltner, M.; Harasek, M. Experimental Analysis of Membrane and Pressure Swing Adsorption (PSA) for the Hydrogen Separation from Natural Gas. J. Clean. Prod. 2017, 167, 896–907. [Google Scholar] [CrossRef]
  102. Vermaak, L.; Neomagus, H.W.J.P.; Bessarabov, D.G. Hydrogen Separation and Purification from Various Gas Mixtures by Means of Electrochemical Membrane Technology in the Temperature Range 100–160 °C. Membranes 2021, 11, 282. [Google Scholar] [CrossRef]
  103. Chang, K.; Li, Q.; Li, Q. Refrigeration Cycle for Cryogenic Separation of Hydrogen from Coke Oven Gas. Front. Energy Power Eng. China 2008, 2, 484–488. [Google Scholar] [CrossRef]
  104. Naeiji, E.; Noorpoor, A.; Ghanavati, H. Energy, Exergy, and Economic Analysis of Cryogenic Distillation and Chemical Scrubbing for Biogas Upgrading and Hydrogen Production. Sustainability 2022, 14, 3686. [Google Scholar] [CrossRef]
  105. Kowalczyk, P.; Gauden, P.A.; Terzyk, A.P. Cryogenic Separation of Hydrogen Isotopes in Single-Walled Carbon and Boron-Nitride Nanotubes: Insight into the Mechanism of Equilibrium Quantum Sieving in Quasi-One-Dimensional Pores. J. Phys. Chem. B 2008, 112, 8275–8284. [Google Scholar] [CrossRef]
  106. Lazar, A.; Brad, S.; Vijulie, M.; Oubraham, A. Cryogenic Distillation Experimental Stands for Hydrogen Isotopes Separation. Fusion Eng. Des. 2019, 146, 1998–2001. [Google Scholar] [CrossRef]
  107. Bogdan, C.; Brad, S.; Necula, H.; Sirosh, O.; Brill, C.; Vijulie, M.; Lazar, A.; Grafov, A. Design, Fabrication, and Test of a Prototype of Matrix Heat Exchanger for Cryogenic Distillation of Hydrogen Isotopes. Fusion Sci. Technol. 2024, 80, 443–454. [Google Scholar] [CrossRef]
  108. Naquash, A.; Riaz, A.; Yehia, F.; Chaniago, Y.D.; Lim, H.; Lee, M. Hydrogen Purification through a Membrane–Cryogenic Integrated Process: A 3 E’s (Energy, Exergy, and Economic) Assessment. Gases 2023, 3, 92–105. [Google Scholar] [CrossRef]
  109. Cavaliere, P. Hydrogen Separation and Purification. In Water Electrolysis for Hydrogen Production; Cavaliere, P., Ed.; Springer International Publishing: Cham, Switzerland, 2023; pp. 509–541. ISBN 978-3-031-37780-8. [Google Scholar]
  110. Yang, X.; Xu, Q. Product Loss Minimization of an Integrated Cryogenic Separation System. Chem. Eng. Technol. 2012, 35, 635–645. [Google Scholar] [CrossRef]
  111. Kangwanpongpan, T.; Makarov, D.; Cirrone, D.; Molkov, V. LES Model of Flash-Boiling and Pressure Recovery Phenomena during Release from Large-Scale Pressurised Liquid Hydrogen Storage Tank. Int. J. Hydrogen Energy 2024, 50, 390–405. [Google Scholar] [CrossRef]
  112. Ohira, K. Study of Nucleate Boiling Heat Transfer to Slush Hydrogen and Slush Nitrogen. Heat Trans. Asian Res. 2003, 32, 13–28. [Google Scholar] [CrossRef]
  113. Pereira, R.; Monteiro, V.; Afonso, J.L.; Teixeira, J. Hydrogen Refueling Stations: A Review of the Technology Involved from Key Energy Consumption Processes to Related Energy Management Strategies. Energies 2024, 17, 4906. [Google Scholar] [CrossRef]
  114. Ozawa, A.; Inoue, M.; Kitagawa, N.; Muramatsu, R.; Anzai, Y.; Genchi, Y.; Kudoh, Y. Assessing Uncertainties of Well-To-Tank Greenhouse Gas Emissions from Hydrogen Supply Chains. Sustainability 2017, 9, 1101. [Google Scholar] [CrossRef]
  115. Harris, P.D.; Barnes, R. The Uses of Helium and Xenon in Current Clinical Practice. Anaesthesia 2008, 63, 284–293. [Google Scholar] [CrossRef]
  116. Xie, Y.; Zhong, K. Investigation on the Performances of Vuilleumier Cycle Heat Pump Adopting Mixture Refrigerants. Entropy 2017, 19, 446. [Google Scholar] [CrossRef]
  117. Moldrik, P.; Sebesta, R. Hydrogen Fuel Cells as a Part of the System for Accumulation of Electric Energy. In Proceedings of the 2009 10th International Conference on Electrical Power Quality and Utilisation, Lodz, Poland, 15–17 September2009; pp. 1–6. [Google Scholar]
  118. Ubic, R. Covalent, Metallic, and Secondary Bonding. In Crystallography and Crystal Chemistry; Ubic, R., Ed.; Springer International Publishing: Cham, Switzerland, 2024; pp. 257–303. ISBN 978-3-031-49752-0. [Google Scholar]
  119. Soejima, M.; Tanaka, N.; Oshima, T.; Kinoshita, H.; Koda, Y. Detection of Helium in a Fire Victim: A Case Report. Forensic Sci. Int. 2021, 318, 110613. [Google Scholar] [CrossRef]
  120. Li, Y.; Zhang, W.; Wang, L.; Zhao, F.; Han, W.; Chen, G. Henry’s Law and Accumulation of Weak Source for Crust-Derived Helium: A Case Study of Weihe Basin, China. J. Nat. Gas Geosci. 2017, 2, 333–339. [Google Scholar] [CrossRef]
  121. Qin, S.; Dou, L.; Tao, G.; Li, J.; Qi, W.; Li, X.; Guo, B.; Zhao, Z.; Wang, J. Helium Enrichment Theory and Exploration Ideas for Helium-Rich Gas Reservoirs. Pet. Explor. Dev. 2024, 51, 1340–1356. [Google Scholar] [CrossRef]
  122. Lu, C.; Wang, B.; Zhu, D.; Liu, Q.; Zhang, X.; He, H. High-Spatial-Resolution Helium Detection and Its Implications for Helium Accumulation Mechanisms. Appl. Sci. 2024, 14, 3453. [Google Scholar] [CrossRef]
  123. NCBI, [National Center for Biotechnology Information] PubChem | Melting Point (Periodic Table of Elements). Available online: https://pubchem.ncbi.nlm.nih.gov/periodic-table/melting-point (accessed on 6 May 2025).
  124. Usas, S.A.; Ricardez-Sandoval, L. An Optimal Sustainable Planning Strategy for National Carbon Capture Deployment: A Review on the State of CO 2 Capture in Canada. Can. J. Chem. Eng. 2024, 102, 2332–2351. [Google Scholar] [CrossRef]
  125. Marzouk, O.A. Subcritical and Supercritical Rankine Steam Cycles, under Elevated Temperatures up to 900 °C and Absolute Pressures up to 400 Bara. Adv. Mech. Eng. 2024, 16, 16878132231221065. [Google Scholar] [CrossRef]
  126. Nijsse, F.J.M.M.; Mercure, J.-F.; Ameli, N.; Larosa, F.; Kothari, S.; Rickman, J.; Vercoulen, P.; Pollitt, H. The Momentum of the Solar Energy Transition. Nat. Commun. 2023, 14, 6542. [Google Scholar] [CrossRef]
  127. Marzouk, O.A. Land-Use Competitiveness of Photovoltaic and Concentrated Solar Power Technologies near the Tropic of Cancer. Sol. Energy 2022, 243, 103–119. [Google Scholar] [CrossRef]
  128. Magar, V.; Peña, A.; Hahmann, A.N.; Pacheco-Rojas, D.A.; García-Hernández, L.S.; Gross, M.S. Wind Energy and the Energy Transition: Challenges and Opportunities for Mexico. Sustainability 2023, 15, 5496. [Google Scholar] [CrossRef]
  129. Marzouk, O.A. Wind Speed Weibull Model Identification in Oman, and Computed Normalized Annual Energy Production (NAEP) From Wind Turbines Based on Data From Weather Stations. Eng. Rep. 2025, 7, e70089. [Google Scholar] [CrossRef]
  130. Priya, S.; Inman, D.J. (Eds.) Energy Harvesting Technologies; Springer: Boston, MA, USA, 2009; ISBN 978-0-387-76463-4. [Google Scholar]
  131. Marzouk, O.A. Thermoelectric Generators versus Photovoltaic Solar Panels: Power and Cost Analysis. Edelweiss Appl. Sci. Technol. 2024, 8, 406–428. [Google Scholar] [CrossRef]
  132. Marzouk, O.A. Tilt Sensitivity for a Scalable One-Hectare Photovoltaic Power Plant Composed of Parallel Racks in Muscat. Cogent Eng. 2022, 9, 2029243. [Google Scholar] [CrossRef]
  133. Shang, Y.; Sang, S.; Tiwari, A.K.; Khan, S.; Zhao, X. Impacts of Renewable Energy on Climate Risk: A Global Perspective for Energy Transition in a Climate Adaptation Framework. Appl. Energy 2024, 362, 122994. [Google Scholar] [CrossRef]
  134. Marzouk, O.A. Energy Generation Intensity (EGI) of Solar Updraft Tower (SUT) Power Plants Relative to CSP Plants and PV Power Plants Using the New Energy Simulator “Aladdin”. Energies 2024, 17, 405. [Google Scholar] [CrossRef]
  135. Chipangamate, N.S.; Nwaila, G.T. Assessment of Challenges and Strategies for Driving Energy Transitions in Emerging Markets: A Socio-Technological Systems Perspective. Energy Geosci. 2024, 5, 100257. [Google Scholar] [CrossRef]
  136. Murmura, M.A.; Cerbelli, S.; Annesini, M.C. Transport-Reaction-Permeation Regimes in Catalytic Membrane Reactors for Hydrogen Production. The Steam Reforming of Methane as a Case Study. Chem. Eng. Sci. 2017, 162, 88–103. [Google Scholar] [CrossRef]
  137. Oh, D.-K.; Lee, K.-Y.; Park, J.-S. Hydrogen Purification from Compact Palladium Membrane Module Using a Low Temperature Diffusion Bonding Technology. Membranes 2020, 10, 338. [Google Scholar] [CrossRef] [PubMed]
  138. Lucia, D.J.; Beran, P.S.; Silva, W.A. Reduced-Order Modeling: New Approaches for Computational Physics. Prog. Aerosp. Sci. 2004, 40, 51–117. [Google Scholar] [CrossRef]
  139. Marzouk, O.A. A Nonlinear ODE System for the Unsteady Hydrodynamic Force—A New Approach. World Acad. Sci. Eng. Technol. 2009, 39, 948–962. [Google Scholar] [CrossRef]
  140. Masoumi-Verki, S.; Haghighat, F.; Eicker, U. A Review of Advances towards Efficient Reduced-Order Models (ROM) for Predicting Urban Airflow and Pollutant Dispersion. Build. Environ. 2022, 216, 108966. [Google Scholar] [CrossRef]
  141. Marzouk, O.; Nayfeh, A. Differential/Algebraic Wake Model Based on the Total Fluid Force and Its Direction, and the Effect of Oblique Immersed-Body Motion on ‘Type-1’ and ‘Type-2’ Lock-In. In Proceedings of the 47th AIAA Aerospace Sciences Meeting including The New Horizons Forum and Aerospace Exposition, AIAA [American Institute of Aeronautics and Astronautics]. Orlando, FL, USA, 5–8 January 2009; p. AIAA 2009-1112. [Google Scholar] [CrossRef]
  142. Keiper, W.; Milde, A.; Volkwein, S. (Eds.) Reduced-Order Modeling (ROM) for Simulation and Optimization; Springer International Publishing: Cham, Switzerland, 2018; ISBN 978-3-319-75318-8. [Google Scholar]
  143. Han, G.; Li, G.; Sun, Y. Electrocatalytic Hydrogenation Using Palladium Membrane Reactors. JACS Au 2024, 4, 328–343. [Google Scholar] [CrossRef]
  144. Marzouk, O.A.; Nayfeh, A.H. Loads on a Harmonically Oscillating Cylinder. In Proceedings of the ASME 2007 International Design Engineering Technical Conferences and Computers and Information in Engineering Conference (IDETC-CIE 2007), ASME [American Society of Mechanical Engineers]. Las Vegas, NV, USA, 20 May 2009; pp. 1755–1774. [Google Scholar] [CrossRef]
  145. Khosravi, M.; Jafari, M.; Ghasemzadeh, K.; Iulianelli, A. Enhancing Styrene Production: CFD Analysis of a Pd-Based Membrane Reactor to Carry out Ethylbenzene Dehydrogenation. Fuel Process. Technol. 2025, 270, 108191. [Google Scholar] [CrossRef]
  146. Marzouk, O.A.; Nayfeh, A.H. Reduction of the Loads on a Cylinder Undergoing Harmonic In-Line Motion. Phys. Fluids 2009, 21, 083103. [Google Scholar] [CrossRef]
  147. Ghasemzadeh, K.; Ghahremani, M.; Yousefi Amiri, T.; Basile, A.; Iulianelli, A. Hydrogen Production by Silica Membrane Reactor during Dehydrogenation of Methylcyclohexane: CFD Analysis. Int. J. Hydrogen Energy 2021, 46, 19768–19777. [Google Scholar] [CrossRef]
  148. Yang, W.-W.; Tang, X.-Y.; Ma, X.; Cao, X.E.; He, Y.-L. Synergistic Intensification of Palladium-Based Membrane Reactors for Hydrogen Production: A Review. Energy Convers. Manag. 2025, 325, 119424. [Google Scholar] [CrossRef]
  149. Chompupun, T.; Limtrakul, S.; Vatanatham, T.; Kanhari, C.; Ramachandran, P.A. Experiments, Modeling and Scaling-up of Membrane Reactors for Hydrogen Production via Steam Methane Reforming. Chem. Eng. Process.—Process Intensif. 2018, 134, 124–140. [Google Scholar] [CrossRef]
  150. Berstad, D.; Nekså, P.; Gjøvåg, G.A. Low-Temperature Syngas Separation and CO2 Capture for Enhanced Efficiency of IGCC Power Plants. Energy Procedia 2011, 4, 1260–1267. [Google Scholar] [CrossRef]
  151. Brdar, R.D.; Jones, R.M. GE IGCC Technology and Experience with Advanced Gas Turbines; General Electric (GE) Power Systems: Schenectady, NY, USA, 2021. [Google Scholar]
  152. Krishnan, G.; Steele, D.; O’Brien, K.; Callahan, R.; Berchtold, K.; Figueroa, J. Simulation of a Process to Capture CO2 From IGCC Syngas Using a High Temperature PBI Membrane. Energy Procedia 2009, 1, 4079–4088. [Google Scholar] [CrossRef]
  153. Gazzani, M.; Macchi, E.; Manzolini, G. CO2 Capture in Integrated Gasification Combined Cycle with SEWGS—Part A: Thermodynamic Performances. Fuel 2013, 105, 206–219. [Google Scholar] [CrossRef]
  154. Jin, H.; Xu, Y.; Lin, R.; Han, W. A Proposal for a Novel Multi-Functional Energy System for the Production of Hydrogen and Power. Int. J. Hydrogen Energy 2008, 33, 9–19. [Google Scholar] [CrossRef]
  155. Park, S.H.; Chung, S.W.; Lee, S.K.; Choi, H.K.; Lee, S.H. Thermo-Economic Evaluation of 300 MW Class Integrated Gasification Combined Cycle with Ash Free Coal (AFC) Process. Appl. Therm. Eng. 2015, 89, 843–852. [Google Scholar] [CrossRef]
  156. Zhang, Z.; Zhu, L.; Yu, J.; Liu, H.; Hao, Q.; Huang, Y. Feasibility and Performance Analysis of a Dual-Regulation System in the Process of Methanol and Electricity Coproduction by Coal Gasification. Energy Technol. 2024, 12, 2300718. [Google Scholar] [CrossRef]
  157. Boruah, A.; Phukan, A.; Singh, S. Utilization of Coal for Hydrogen Generation. Int. J. Coal Prep. Util. 2024, 1–22. [Google Scholar] [CrossRef]
  158. Zhao, L.; Etezadi, R.; Tsotsis, T. Chapter 15—Syngas from Coal. In Advances in Synthesis Gas : Methods, Technologies and Applications; Rahimpour, M.R., Makarem, M.A., Meshksar, M., Eds.; Elsevier: Amsterdam, The Netherlands, 2023; Volume 1, pp. 363–377. ISBN 978-0-323-91871-8. [Google Scholar]
  159. Vairakannu, P. Clean Coal Technologies. In Sustainable Utilization of Natural Resources; CRC Press: Boca Raton, FL, USA, 2017; ISBN 978-1-315-15329-2. [Google Scholar]
  160. Gräbner, M.; von Morstein, O.; Rappold, D.; Günster, W.; Beysel, G.; Meyer, B. Constructability Study on a German Reference IGCC Power Plant with and without CO2-Capture for Hard Coal and Lignite. Energy Convers. Manag. 2010, 51, 2179–2187. [Google Scholar] [CrossRef]
  161. Aineto, M.; Acosta, A.; Rincón, J.M.; Romero, M. Thermal Expansion of Slag and Fly Ash from Coal Gasification in IGCC Power Plant. Fuel 2006, 85, 2352–2358. [Google Scholar] [CrossRef]
  162. Melchior, T.; Madlener, R. Economic Evaluation of IGCC Plants with Hot Gas Cleaning. Appl. Energy 2012, 97, 170–184. [Google Scholar] [CrossRef]
  163. Siefert, N.S.; Narburgh, S.; Chen, Y. Comprehensive Exergy Analysis of Three IGCC Power Plant Configurations with CO2 Capture. Energies 2016, 9, 669. [Google Scholar] [CrossRef]
  164. Pérez-Fortes, M.; Bojarski, A.D.; Velo, E.; Nougués, J.M.; Puigjaner, L. Conceptual Model and Evaluation of Generated Power and Emissions in an IGCC Plant. Energy 2009, 34, 1721–1732. [Google Scholar] [CrossRef]
  165. Valero, A.; Usón, S. Oxy-Co-Gasification of Coal and Biomass in an Integrated Gasification Combined Cycle (IGCC) Power Plant. Energy 2006, 31, 1643–1655. [Google Scholar] [CrossRef]
  166. Casella, F.; Colonna, P. Dynamic Modeling of IGCC Power Plants. Appl. Therm. Eng. 2012, 35, 91–111. [Google Scholar] [CrossRef]
  167. Merkel, T.C.; Zhou, M.; Baker, R.W. Carbon Dioxide Capture with Membranes at an IGCC Power Plant. J. Membr. Sci. 2012, 389, 441–450. [Google Scholar] [CrossRef]
  168. Oh, H.-T.; Lee, W.-S.; Ju, Y.; Lee, C.-H. Performance Evaluation and Carbon Assessment of IGCC Power Plant with Coal Quality. Energy 2019, 188, 116063. [Google Scholar] [CrossRef]
  169. Wang, T. An Overview of IGCC Systems. In Integrated Gasification Combined Cycle (IGCC) Technologies; Elsevier: Amsterdam, The Netherlands, 2017; pp. 1–80. ISBN 978-0-08-100167-7. [Google Scholar]
  170. Wolfersdorf, C.; Meyer, B. 25—The Current Status and Future Prospects for IGCC Systems. In Integrated Gasification Combined Cycle (IGCC) Technologies; Wang, T., Stiegel, G., Eds.; Woodhead Publishing: Sawston, UK, 2017; pp. 847–889. ISBN 978-0-08-100167-7. [Google Scholar]
  171. Siefert, N.S.; Litster, S. Exergy and Economic Analyses of Advanced IGCC–CCS and IGFC–CCS Power Plants. Appl. Energy 2013, 107, 315–328. [Google Scholar] [CrossRef]
  172. Descamps, C.; Bouallou, C.; Kanniche, M. Efficiency of an Integrated Gasification Combined Cycle (IGCC) Power Plant Including CO2 Removal. Energy 2008, 33, 874–881. [Google Scholar] [CrossRef]
  173. Ståhl, K.; Neergaard, M. IGCC Power Plant for Biomass Utilisation, Värnamo, Sweden. Biomass Bioenergy 1998, 15, 205–211. [Google Scholar] [CrossRef]
  174. Hossein Sahraei, M.; McCalden, D.; Hughes, R.; Ricardez-Sandoval, L.A. A Survey on Current Advanced IGCC Power Plant Technologies, Sensors and Control Systems. Fuel 2014, 137, 245–259. [Google Scholar] [CrossRef]
  175. Sofia, D.; Coca Llano, P.; Giuliano, A.; Iborra Hernández, M.; García Peña, F.; Barletta, D. Co-Gasification of Coal–Petcoke and Biomass in the Puertollano IGCC Power Plant. Chem. Eng. Res. Des. 2014, 92, 1428–1440. [Google Scholar] [CrossRef]
  176. Buttler, A.; Kunze, C.; Spliethoff, H. IGCC–EPI: Decentralized Concept of a Highly Load-Flexible IGCC Power Plant for Excess Power Integration. Appl. Energy 2013, 104, 869–879. [Google Scholar] [CrossRef]
  177. Omell, B.P.; Chmielewski, D.J. IGCC Power Plant Dispatch Using Infinite-Horizon Economic Model Predictive Control. Ind. Eng. Chem. Res. 2013, 52, 3151–3164. [Google Scholar] [CrossRef]
  178. Pruschek, R.; Oeljeklaus, G.; Kloster, R. Enhancement of the Efficiency of Coal-Fired Power Generation Systems: Final Report; Institute for Process Engineering and Power Plant Technology, University of Stuttgart: Stuttgart, Germany, 1998. [Google Scholar]
  179. Marzouk, O.A. Condenser Pressure Influence on Ideal Steam Rankine Power Vapor Cycle Using the Python Extension Package Cantera for Thermodynamics. Eng. Technol. Appl. Sci. Res. 2024, 14, 14069–14078. [Google Scholar] [CrossRef]
  180. Mitsubishi Power. IGCC Integrated Coal Gasification Combined Cycle Power Plants; Mitsubishi Power, a brand of Mitsubishi Heavy Industries, Ltd.: Kanagawa, Japan, 2021. [Google Scholar]
  181. Marzouk, O.A. Chronologically-Ordered Quantitative Global Targets for the Energy-Emissions-Climate Nexus, from 2021 to 2050. In Proceedings of the 2022 International Conference on Environmental Science and Green Energy (ICESGE), Shenyang, China, 25–27 December 2022; pp. 1–6. [Google Scholar] [CrossRef]
  182. Zhao, Y.; Ying, D.; Liu, Y.; Wentao, R. Relative Stability Analysis Method of Systems Based on Goal Seek. In Proceedings of the 2021 3rd International Conference on Industrial Artificial Intelligence (IAI), Shenyang, China, 8–11 November 2021; pp. 1–5. [Google Scholar]
  183. Guerrero, H. Solver, Scenarios, and Goal Seek Tools. In Excel Data Analysis: Modeling and Simulation; Guerrero, H., Ed.; Springer International Publishing: Cham, Switzerland, 2019; pp. 311–346. ISBN 978-3-030-01279-3. [Google Scholar]
  184. Tian, Y.; Qin, C.; Yang, Z.; Hao, D. Numerical Simulation Study on the Leakage and Diffusion Characteristics of High-Pressure Hydrogen Gas in Different Spatial Scenes. Int. J. Hydrogen Energy 2024, 50, 1335–1349. [Google Scholar] [CrossRef]
  185. Marzouk, O.A. Contrasting the Cartesian and Polar Forms of the Shedding-Induced Force Vector in Response to 12 Subharmonic and Superharmonic Mechanical Excitations. Fluid Dyn. Res. 2010, 42, 035507. [Google Scholar] [CrossRef]
  186. Ghasemzadeh, K.; Torabi, T.; Yousefi Amiri, T.; Fortunelli, A.; Iulianelli, A. Parametric and Sensitive Analysis of Pd-Ag Membrane Reactor Performance in Biogas Reforming to Generate Decarbonized Hydrogen by Computational Fluid Dynamic-Response Surface Methodology. Fuel 2024, 365, 131205. [Google Scholar] [CrossRef]
  187. Marzouk, O.A. Characteristics of the Flow-Induced Vibration and Forces With 1- and 2-DOF Vibrations and Limiting Solid-to-Fluid Density Ratios. J. Vib. Acoust. 2010, 132, 041013. [Google Scholar] [CrossRef]
  188. Kassi, A.H.; Al-Hattab, T.A. A CFD Model of Natural Gas Steam Reforming in a Catalytic Membrane Reactor: Effect of Various Operating Parameters on the Performance of CMR. Int. J. Hydrogen Energy 2024, 56, 780–796. [Google Scholar] [CrossRef]
  189. Marzouk, O.A.; Nayfeh, A.H. Characterization of the Flow over a Cylinder Moving Harmonically in the Cross-Flow Direction. Int. J. Non-Linear Mech. 2010, 45, 821–833. [Google Scholar] [CrossRef]
  190. Zakeri, R.; Fazeli, A. CFD Modeling of Hydrogen Production from Glycerol Steam Reforming in Tesla Microchannel Reactor. Fuel 2024, 357, 129646. [Google Scholar] [CrossRef]
  191. Li, X.; Zhao, Y.; Liu, Z.; Chen, H. A New Methodology for Preliminary Design of Centrifugal Impellers with Prewhirl. Proc. Inst. Mech. Eng. Part A J. Power Energy 2020, 234, 251–262. [Google Scholar] [CrossRef]
  192. Marzouk, O.A. Evolutionary Computing Applied to Design Optimization. In Proceedings of the Volume 2: 27th Computers and Information in Engineering Conference, Parts A and B, Las Vegas, NV, USA, 4–7 September 2007; ASMEDC: Las Vegas, NV, USA, 2007; pp. 995–1003. [Google Scholar] [CrossRef]
  193. Zhang, Z. Understanding and Exploiting the Nonlinear Behavior of Tuned Liquid Dampers (TLDs) for Structural Vibration Control by Means of a Nonlinear Reduced-Order Model (ROM). Eng. Struct. 2022, 251, 113524. [Google Scholar] [CrossRef]
  194. Marzouk, O.A. Flow Control Using Bifrequency Motion. Theor. Comput. Fluid Dyn. 2011, 25, 381–405. [Google Scholar] [CrossRef]
  195. Marzouk, O.A. A Two-Step Computational Aeroacoustics Method Applied to High-Speed Flows. Noise Control Eng. J. 2008, 56, 396. [Google Scholar] [CrossRef]
  196. San, J.Y.; Worek, W.M.; Lavan, Z. Entropy Generation in Convective Heat Transfer and Isothermal Convective Mass Transfer. J. Heat Transf. 1987, 109, 647–652. [Google Scholar] [CrossRef]
  197. DOE, [United States Department of Energy]. Department of Energy Fundamentals Training Handbook, Thermodynamics, Heat Transfer, and Fluid Flow, Volume 2 of 3; United States Department of Energy (DOE): Washington, DC, USA, 1992.
  198. Enerquip What’s the Difference between Parallel Flow, Counter Flow and Crossflow Heat Exchangers? Available online: https://www.enerquip.com/whats-the-difference-between-parallel-flow-counter-flow-and-crossflow-heat-exchangers (accessed on 7 May 2022).
  199. Thermex Why Counter Flow Heat Exchangers Are More Efficient. Available online: http://www.thermex.co.uk/news/blog/605-why-counter-flow-heat-exchangers-are-more-efficient (accessed on 7 May 2022).
  200. Pinacci, P.; Drago, F. Influence of the Support on Permeation of Palladium Composite Membranes in Presence of Sweep Gas. Catal. Today 2012, 193, 186–193. [Google Scholar] [CrossRef]
  201. Skoulidas, A.I.; Sholl, D.S. Multiscale Models of Sweep Gas and Porous Support Effects on Zeolite Membranes. AIChE J. 2005, 51, 867–877. [Google Scholar] [CrossRef]
  202. Akamatsu, K.; Ohta, Y.; Sugawara, T.; Kanno, N.; Tonokura, K.; Hattori, T.; Nakao, S. Stable High-Purity Hydrogen Production by Dehydrogenation of Cyclohexane Using a Membrane Reactor with Neither Carrier Gas nor Sweep Gas. J. Membr. Sci. 2009, 330, 1–4. [Google Scholar] [CrossRef]
  203. Moussaddy, S.; Maisonneuve, J. Energy from Carbon Dioxide: Experimental and Theoretical Analysis of Power Generation from Membrane-Based Sweep Gas Permeation. J. Membr. Sci. 2022, 644, 120053. [Google Scholar] [CrossRef]
  204. Li, Z.; Zhang, P.; Guan, K.; Gonzales, R.R.; Ishigami, T.; Xue, M.; Yoshioka, T.; Matsuyama, H. An Experimental Study on Recovering and Concentrating Ammonia by Sweep Gas Membrane Distillation. Process Saf. Environ. Prot. 2023, 171, 555–560. [Google Scholar] [CrossRef]
  205. Gobina, E.; Hou, K.; Hughes, R. Ethane Dehydrogenation in a Catalytic Membrane Reactor Coupled with a Reactive Sweep Gas. Chem. Eng. Sci. 1995, 50, 2311–2319. [Google Scholar] [CrossRef]
  206. Xie, Z.; Duong, T.; Hoang, M.; Nguyen, C.; Bolto, B. Ammonia Removal by Sweep Gas Membrane Distillation. Water Res. 2009, 43, 1693–1699. [Google Scholar] [CrossRef] [PubMed]
  207. Yoon, H.; Lee, J.; Lee, G.; Shin, W.; Min, T. Membrane Degassing with the Combination of Sweep Gas and Vacuum Pressure for Ammonia Removal. Environ. Sci. Water Res. Technol. 2023, 9, 467–473. [Google Scholar] [CrossRef]
  208. Shukla, S.; Benes, N.E.; Vankelecom, I.; Méricq, J.P.; Belleville, M.P.; Hengl, N.; Marcano, J.S. Sweep Gas Membrane Distillation in a Membrane Contactor with Metallic Hollow-Fibers. J. Membr. Sci. 2015, 493, 167–178. [Google Scholar] [CrossRef]
  209. Nordio, M.; Soresi, S.; Manzolini, G.; Melendez, J.; Van Sint Annaland, M.; Pacheco Tanaka, D.A.; Gallucci, F. Effect of Sweep Gas on Hydrogen Permeation of Supported Pd Membranes: Experimental and Modeling. Int. J. Hydrogen Energy 2019, 44, 4228–4239. [Google Scholar] [CrossRef]
  210. Lee, C.H.; Hong, W.H. Effect of Operating Variables on the Flux and Selectivity in Sweep Gas Membrane Distillation for Dilute Aqueous Isopropanol. J. Membr. Sci. 2001, 188, 79–86. [Google Scholar] [CrossRef]
  211. Franz, J.; Schiebahn, S.; Zhao, L.; Riensche, E.; Scherer, V.; Stolten, D. Investigating the Influence of Sweep Gas on CO2/N2 Membranes for Post-Combustion Capture. Int. J. Greenh. Gas Control 2013, 13, 180–190. [Google Scholar] [CrossRef]
  212. Jiménez-Robles, R.; Gabaldón, C.; Martínez-Soria, V.; Izquierdo, M. Simultaneous Application of Vacuum and Sweep Gas in a Polypropylene Membrane Contactor for the Recovery of Dissolved Methane from Water. J. Membr. Sci. 2021, 617, 118560. [Google Scholar] [CrossRef]
  213. Ben-Mansour, R.; Haque, M.A.; Habib, M.A.; Paglieri, S.; Harale, A.; Mokheimer, E.M.A. Effect of Temperature and Heat Flux Boundary Conditions on Hydrogen Production in Membrane-Integrated Steam-Methane Reformer. Appl. Energy 2023, 346, 121407. [Google Scholar] [CrossRef]
  214. Barbieri, G. Sweep Gas in a Membrane Reactor. In Encyclopedia of Membranes; Drioli, E., Giorno, L., Eds.; Springer: Berlin/Heidelberg, Germany, 2015; pp. 1–2. ISBN 978-3-642-40872-4. [Google Scholar]
  215. Li, Z.; Polfus, J.M.; Xing, W.; Denonville, C.; Fontaine, M.-L.; Bredesen, R. Factors Limiting the Apparent Hydrogen Flux in Asymmetric Tubular Cercer Membranes Based on LaCO27WCO3.5MoCO1.5O55.5−δ and La0.87Sr0.13CrO3−δ. Membranes 2019, 9, 126. [Google Scholar] [CrossRef]
  216. Xie, D.; Yu, J.; Wang, F.; Zhang, N.; Wang, W.; Yu, H.; Peng, F.; Park, A.-H.A. Hydrogen Permeability of Pd–Ag Membrane Modules with Porous Stainless Steel Substrates. Int. J. Hydrogen Energy 2011, 36, 1014–1026. [Google Scholar] [CrossRef]
  217. Chein, R.Y.; Chen, Y.C.; Chung, J.N. Sweep Gas Flow Effect on Membrane Reactor Performance for Hydrogen Production from High-Temperature Water-Gas Shift Reaction. J. Membr. Sci. 2015, 475, 193–203. [Google Scholar] [CrossRef]
  218. Balachandran, U.; Dorris, S.E.; Emerson, J.E.; Lee, T.H.; Lu, Y.; Park, C.Y.; Picciolo, J.J. Hydrogen Separation Membranes; Argonne National Laboratory (ANL) of the United States Department of Energy: Lemont, IL, USA, 2011.
  219. Bidica, N.; Ghimis, N.; Monea, B. Experimental Results of Deuterium/Hydrogen Co-Current Permeation through Nickel Membrane, in Non-Steady-State Surface Limited Regime. Fusion Eng. Des. 2023, 194, 113718. [Google Scholar] [CrossRef]
  220. Escolástico, S.; Ivanova, M.; Solís, C.; Roitsch, S.; Meulenberg, W.A.; Serra, J.M. Improvement of Transport Properties and Hydrogen Permeation of Chemically-Stable Proton-Conducting Oxides Based on the System BaZr1-x-yYxMyO3-δ. RSC Adv. 2012, 2, 4932. [Google Scholar] [CrossRef]
  221. Longhurst, G.R.; Kratville, A. A Simple System for Measuring Permeation of Hydrogen Through Stainless Steel. Fusion Sci. Technol. 2014, 66, 385–393. [Google Scholar] [CrossRef]
  222. Chen, W.-H.; Lin, C.-H.; Lin, Y.-L.; Tsai, C.-W.; Chein, R.-Y.; Yu, C.-T. Interfacial Permeation Phenomena of Hydrogen Purification and Carbon Dioxide Separation in a Non-Isothermal Palladium Membrane Tube. Chem. Eng. J. 2016, 305, 156–168. [Google Scholar] [CrossRef]
  223. Chiesa, P.; Romano, M.C.; Kreutz, T.G. Use of Membranes in Systems for Electric Energy and Hydrogen Production from Fossil Fuels. In Handbook of Membrane Reactors; Elsevier: Amsterdam, The Netherlands, 2013; pp. 416–455. ISBN 978-0-85709-415-5. [Google Scholar]
  224. Marín, P.; Patiño, Y.; Díez, F.V.; Ordóñez, S. Modelling of Hydrogen Perm-Selective Membrane Reactors for Catalytic Methane Steam Reforming. Int. J. Hydrogen Energy 2012, 37, 18433–18445. [Google Scholar] [CrossRef]
  225. Park, Y.; Cha, J.; Oh, H.-T.; Lee, T.; Lee, S.H.; Park, M.G.; Jeong, H.; Kim, Y.; Sohn, H.; Nam, S.W.; et al. A Catalytic Composite Membrane Reactor System for Hydrogen Production from Ammonia Using Steam as a Sweep Gas. J. Membr. Sci. 2020, 614, 118483. [Google Scholar] [CrossRef]
  226. Brunetti, A.; Caravella, A.; Drioli, E.; Barbieri, G. CHAPTER 1. Membrane Reactors for Hydrogen Production. In Membrane Engineering for the Treatment of Gases: Volume 2: Gas-Separation Issues Combined with Membrane Reactors; The Royal Society of Chemistry: London, UK, 2017; pp. 1–29. [Google Scholar]
  227. Li, X.; Li, A.; Lim, C.J.; Grace, J.R. Hydrogen Permeation through Pd-Based Composite Membranes: Effects of Porous Substrate, Diffusion Barrier and Sweep Gas. J. Membr. Sci. 2016, 499, 143–155. [Google Scholar] [CrossRef]
  228. Meng, X.; Shang, Y.; Meng, B.; Yang, N.; Tan, X.; Sunarso, J.; Liu, S. Bi-Functional Performances of BaCe0.95Tb0.05O3−δ-Based Hollow Fiber Membranes for Power Generation and Hydrogen Permeation. J. Eur. Ceram. Soc. 2016, 36, 4123–4129. [Google Scholar] [CrossRef]
  229. Roberts, R.M.; Elleman, T.S.; Iii, H.P.; Verghese, K. Hydrogen Permeability of Sintered Aluminum Oxide. J. Am. Ceram. Soc. 1979, 62, 495–499. [Google Scholar] [CrossRef]
  230. Escolástico, S.; Solís, C.; Kjølseth, C.; Serra, J.M. Outstanding Hydrogen Permeation through CO 2 -Stable Dual-Phase Ceramic Membranes. Energy Environ. Sci. 2014, 7, 3736–3746. [Google Scholar] [CrossRef]
  231. Qi, X. Electrical Conduction and Hydrogen Permeation through Mixed Proton–Electron Conducting Strontium Cerate Membranes. Solid State Ion. 2000, 130, 149–156. [Google Scholar] [CrossRef]
  232. Sakbodin, M.; Schulman, E.; Oh, S.C.; Pan, Y.; Wachsman, E.D.; Liu, D. Dual Utilization of Greenhouse Gases to Produce C2+ Hydrocarbons and Syngas in a Hydrogen-Permeable Membrane Reactor. J. Membr. Sci. 2020, 595, 117557. [Google Scholar] [CrossRef]
  233. Smith, J.B.; Aasen, K.I.; Wilhelmsen, K.; Käka, D.; Risdal, T.; Berglund, A.; Stenersen Østby, A.; Budd, M.; Bruun, T.; Werswick, B. Recent Development in the HMR Pre-Combustion Gas Power Cycle. Energy Procedia 2009, 1, 343–351. [Google Scholar] [CrossRef]
  234. Tan, X.; Tan, X.; Yang, N.; Meng, B.; Zhang, K.; Liu, S. High Performance BaCe0.8Y0.2O3−a (BCY) hollow fibre membranes for hydrogen permeation. Ceram. Int. 2014, 40, 3131–3138. [Google Scholar] [CrossRef]
  235. Unruh, D.; Pabst, K.; Schaub, G. Fischer−Tropsch Synfuels from Biomass: Maximizing Carbon Efficiency and Hydrocarbon Yield. Energy Fuels 2010, 24, 2634–2641. [Google Scholar] [CrossRef]
  236. GTI, [Gas Technology Institute]. Direct Hydrogen Production from Biomass Gasifier Using Hydrogen-Selective Membrane; Gas Technology Institute (GTI), In Collaboration with Natural Resources Research Institute University of Minnesota at Duluth: Des Plaines, IL, USA, 2007. [Google Scholar]
  237. Kinouchi, K.; Katoh, M.; Horikawa, T.; Yoshikawa, T.; Wada, M. Hydrogen permeability of palladium membrane for steam-reforming of bio-ethanol using the membrane reactor. Int. J. Mod. Phys. Conf. Ser. 2012, 6, 7–12. [Google Scholar] [CrossRef]
  238. Barnoon, P. Modeling of a High Temperature Heat Exchanger to Supply Hydrogen Required by Fuel Cells through Reforming Process. Energy Rep. 2021, 7, 5685–5699. [Google Scholar] [CrossRef]
  239. Barnoon, P.; Toghraie, D.; Mehmandoust, B.; Fazilati, M.A.; Eftekhari, S.A. Comprehensive Study on Hydrogen Production via Propane Steam Reforming inside a Reactor. Energy Rep. 2021, 7, 929–941. [Google Scholar] [CrossRef]
  240. Fornarelli, F.; Dambrosio, L.; Terlizzi, L.; Camporeale, S.M. Performance and Cost Multi Objective Optimisation of a Shell-and-Tube LHTES Device for Mid-Temperature Applications. J. Energy Storage 2024, 99, 113134. [Google Scholar] [CrossRef]
  241. Kirincic, M.; Trp, A.; Lenic, K.; Torbarina, F. Numerical Analysis of the Influence of Geometry Parameters on Charging and Discharging Performance of Shell-and-Tube Latent Thermal Energy Storage with Longitudinal Fins. Appl. Therm. Eng. 2024, 236, 121385. [Google Scholar] [CrossRef]
  242. Li, C.; Li, Q.; Ge, R. Comparison of Performance Enhancement in a Shell and Tube Based Latent Heat Thermal Energy Storage Device Containing Different Structured Fins. Renew. Energy 2023, 206, 994–1006. [Google Scholar] [CrossRef]
  243. Marzouk, O.A. Direct Numerical Simulations of the Flow Past a Cylinder Moving With Sinusoidal and Nonsinusoidal Profiles. J. Fluids Eng. 2009, 131, 121201. [Google Scholar] [CrossRef]
  244. Daneshparvar, M.R.; Beigzadeh, R. Multi-Objective Optimization of Helical Baffles in the Shell-and-Tube Heat Exchanger by Computational Fluid Dynamics and Genetic Algorithm. Energy Rep. 2022, 8, 11064–11077. [Google Scholar] [CrossRef]
  245. Marzouk, O.A. One-Way and Two-Way Couplings of CFD and Structural Models and Application to the Wake-Body Interaction. Appl. Math. Model. 2011, 35, 1036–1053. [Google Scholar] [CrossRef]
  246. Marzouk, O.A.; Huckaby, E.D. Simulation of a Swirling Gas-Particle Flow Using Different k-Epsilon Models and Particle-Parcel Relationships. Eng. Lett. 2010, 18, 7. [Google Scholar] [CrossRef]
  247. Saldanha, W.H.; Arrieta, F.R.P.; Soares, G.L. State-of-the-Art of Research on Optimization of Shell and Tube Heat Exchangers by Methods of Evolutionary Computation. Arch Comput. Methods Eng 2021, 28, 2761–2783. [Google Scholar] [CrossRef]
  248. Marzouk, O.A.; Huckaby, E.D. Modeling Confined Jets with Particles and Swril. In Machine Learning and Systems Engineering; Ao, S.-I., Rieger, B., Amouzegar, M.A., Eds.; Lecture Notes in Electrical Engineering; Springer: Dordrecht, The Netherlands, 2010; Volume 68, pp. 243–256. ISBN 978-90-481-9418-6. [Google Scholar] [CrossRef]
  249. Li, P.; Chen, L.; Xia, S.; Zhang, L. Entropy Generation Rate Minimization for Methanol Synthesis via a CO2 Hydrogenation Reactor. Entropy 2019, 21, 174. [Google Scholar] [CrossRef] [PubMed]
  250. Li, P.; Chen, L.; Xia, S.; Kong, R.; Ge, Y. Multi-Objective Optimal Configurations of a Membrane Reactor for Steam Methane Reforming. Energy Rep. 2022, 8, 527–538. [Google Scholar] [CrossRef]
  251. Bichkar, P.; Dandgaval, O.; Dalvi, P.; Godase, R.; Dey, T. Study of Shell and Tube Heat Exchanger with the Effect of Types of Baffles. Procedia Manuf. 2018, 20, 195–200. [Google Scholar] [CrossRef]
  252. Pal, E.; Kumar, I.; Joshi, J.B.; Maheshwari, N.K. CFD Simulations of Shell-Side Flow in a Shell-and-Tube Type Heat Exchanger with and without Baffles. Chem. Eng. Sci. 2016, 143, 314–340. [Google Scholar] [CrossRef]
  253. Salahuddin, U.; Bilal, M.; Ejaz, H. A Review of the Advancements Made in Helical Baffles Used in Shell and Tube Heat Exchangers. Int. Commun. Heat Mass Transf. 2015, 67, 104–108. [Google Scholar] [CrossRef]
  254. Jamsran, N.; Park, H.; Lee, J.; Oh, S.; Kim, C.; Lee, Y.; Kang, K. Influence of Syngas Composition on Combustion and Emissions in a Homogeneous Charge Compression Ignition Engine. Fuel 2021, 306, 121774. [Google Scholar] [CrossRef]
  255. Kousheshi, N.; Yari, M.; Paykani, A.; Saberi Mehr, A.; de la Fuente, G.F. Effect of Syngas Composition on the Combustion and Emissions Characteristics of a Syngas/Diesel RCCI Engine. Energies 2020, 13, 212. [Google Scholar] [CrossRef]
  256. Ribeiro, A.M.; Santos, J.C.; Rodrigues, A.E. PSA Design for Stoichiometric Adjustment of Bio-Syngas for Methanol Production and Co-Capture of Carbon Dioxide. Chem. Eng. J. 2010, 163, 355–363. [Google Scholar] [CrossRef]
  257. Giuliano, A.; Freda, C.; Catizzone, E. Techno-Economic Assessment of Bio-Syngas Production for Methanol Synthesis: A Focus on the Water–Gas Shift and Carbon Capture Sections. Bioengineering 2020, 7, 70. [Google Scholar] [CrossRef]
  258. NIST, [United States National Institute of Standards and Technology]. NIST Chemistry WebBook—Hydrogen. Available online: https://webbook.nist.gov/cgi/cbook.cgi?Name=h2&Units=SI (accessed on 7 May 2022).
  259. NIST, [United States National Institute of Standards and Technology]. NIST Chemistry WebBook—Carbon Monoxide. Available online: https://webbook.nist.gov/cgi/cbook.cgi?Name=CO&Units=SI (accessed on 7 May 2022).
  260. NIST, [United States National Institute of Standards and Technology]. NIST Chemistry WebBook—Carbon Dioxide. Available online: https://webbook.nist.gov/cgi/cbook.cgi?ID=C124389&Units=SI (accessed on 7 May 2022).
  261. Kuo, K.K. Principles of Combustion, 2nd ed.; John Wiley: Hoboken, NJ, USA, 2005; ISBN 978-0-471-04689-9. [Google Scholar]
  262. Shi, L.; Liu, X.-J.; Wang, X.; Zhu, M.-S. Prediction Method for Liquid Thermal Conductivity of Refrigerant Mixtures. Fluid Phase Equilibria 2000, 172, 293–306. [Google Scholar] [CrossRef]
  263. Agathou, M.S.; Kyritsis, D.C. An Experimental Comparison of Non-Premixed Bio-Butanol Flames with the Corresponding Flames of Ethanol and Methane. Fuel 2011, 90, 255–262. [Google Scholar] [CrossRef]
  264. Marzouk, O.A. Estimated Electric Conductivities of Thermal Plasma for Air-Fuel Combustion and Oxy-Fuel Combustion with Potassium or Cesium Seeding. Heliyon 2024, 10, e31697. [Google Scholar] [CrossRef] [PubMed]
  265. Mills, P.L. Analysis of Multiphase Polycarbonate Polymerization in a Semibatch Reactor. Chem. Eng. Sci. 1986, 41, 2939–2952. [Google Scholar] [CrossRef]
  266. Hudebine, D.; Verstraete, J.J. Molecular Reconstruction of LCO Gasoils from Overall Petroleum Analyses. Chem. Eng. Sci. 2004, 59, 4755–4763. [Google Scholar] [CrossRef]
  267. Bell, D.; Heyne, J.S.; Won, S.H.; Dryer, F.; Haas, F.M.; Dooley, S. On the Development of General Surrogate Composition Calculations for Chemical and Physical Properties. In Proceedings of the 55th AIAA Aerospace Sciences Meeting, Grapevine, TX, USA, 9–13 January 2017; p. AIAA 2017-0609. [Google Scholar]
  268. Poinsot, T.; Veynante, D. Theoretical and Numerical Combustion, 2nd ed.; Edwards: Philadelphia, PA, USA, 2005; ISBN 978-1-930217-10-2. [Google Scholar]
  269. Shikida, M.; Yoshikawa, K.; Iwai, S.; Sato, K. Flexible Flow Sensor for Large-Scale Air-Conditioning Network Systems. Sens. Actuators A Phys. 2012, 188, 2–8. [Google Scholar] [CrossRef]
  270. Heselmann, M.; Dämgen, U.; Brümmer, A. Experimental Investigation of the Distribution of Two-Phase Flow in Oil-Injected Twin-Screw Compressors. In Proceedings of the 13th International Conference on Compressors and Their Systems, London, UK, 11–13 September 2023; Read, M., Rane, S., Ivkovic-Kihic, I., Kovacevic, A., Eds.; Springer Nature: Cham, Switzerland, 2024; pp. 61–76. [Google Scholar]
  271. Johansson, P.S.; Valle, A.; Holm, H.T. Model-Uncertainty-Estimation Methodology Used on High-Flow-Rate Field Cases. SPE J. 2017, 22, 1254–1260. [Google Scholar] [CrossRef]
  272. Carrington, C.G.; Liu, Q. Calorimeter Measurements of a Heat Pump Dehumidifier: Influence of Evaporator Air Flow. Int. J. Energy Res. 1995, 19, 649–658. [Google Scholar] [CrossRef]
  273. Stoppel, L.; Fehling, T.; Geißler, T.; Baake, E.; Wetzel, T. Carbon Dioxide Free Production of Hydrogen. IOP Conf. Ser. Mater. Sci. Eng. 2017, 228, 012016. [Google Scholar] [CrossRef]
  274. Wang, X.; Zhao, Y.; Yu, H.; Liu, Y. Research on Dynamic Characteristics of Laminar Flow Meter. Flow Meas. Instrum. 2024, 97, 102619. [Google Scholar] [CrossRef]
  275. Murgia, S.; Valenti, G.; Contaldi, G.; Valenti, A. Experimental Investigation on Materials and Lubricants for Sliding-Vane Air Compressors. IOP Conf. Ser. Mater. Sci. Eng. 2015, 90, 012039. [Google Scholar] [CrossRef]
  276. Wiederkehr, R.S.; Salvadori, M.C.; Degasperi, F.T.; Cattani, M. Development of Microvalves for Gas Flow Control in Micronozzles Using PVDF Piezoelectric Polymer. J. Phys. Conf. Ser. 2008, 100, 052046. [Google Scholar] [CrossRef]
  277. Rönsch, S.; Auer, B.; Kinateder, M.; Gleichmann, K. Zeolite Heat Storage: Key Parameters from Experimental Results with Binder-Free NaY. Chem. Eng. Technol. 2020, 43, 2530–2537. [Google Scholar] [CrossRef]
  278. Dierks, S.; Kroll, A. Quantification of Methane Gas Leakages Using Remote Sensing and Sensor Data Fusion. In Proceedings of the 2017 IEEE Sensors Applications Symposium (SAS), Glassboro, NJ, USA, 13–15 March 2017; pp. 1–6. [Google Scholar]
  279. IUPAC, [International Union of Pure and Applied Chemistry]. IUPAC Compendium of Chemical Terminology, 3rd Ed. Online Version 3.0.1. Available online: https://doi.org/10.1351/goldbook.S06036 (accessed on 1 June 2024).
  280. Klinger, R.; Kolářik, C.; Fluck, J.; Hofmann-Apitius, M.; Friedrich, C.M. Detection of IUPAC and IUPAC-like Chemical Names. Bioinformatics 2008, 24, i268–i276. [Google Scholar] [CrossRef] [PubMed]
  281. Kuhn, H.J.; Braslavsky, S.E.; Schmidt, R. Chemical actinometry (IUPAC Technical Report). Pure Appl. Chem. 2004, 76, 2105–2146. [Google Scholar] [CrossRef]
  282. Pettit, L.D. The IUPAC Stability Constants Database. Chem. Int.—Newsmag. IUPAC 2006, 28, 14–15. [Google Scholar] [CrossRef]
  283. Jenkins, A.D.; Kratochvíl, P.; Stepto, R.F.T.; Suter, U.W. Glossary of basic terms in polymer science (IUPAC Recommendations 1996). Pure Appl. Chem. 1996, 68, 2287–2311. [Google Scholar] [CrossRef]
  284. Guthrie, R.D.; Jencks, W.P. IUPAC Recommendations for the Representation of Reaction Mechanisms. Available online: https://pubs.acs.org/doi/pdf/10.1021/ar00166a001 (accessed on 4 May 2025).
  285. Buck, R.P.; Rondinini, S.; Covington, A.K.; Baucke, F.G.K.; Brett, C.M.A.; Camoes, M.F.; Milton, M.J.T.; Mussini, T.; Naumann, R.; Pratt, K.W.; et al. Measurement of pH. Definition, Standards, and Procedures (IUPAC Recommendations 2002). Pure Appl. Chem. 2002, 74, 2169–2200. [Google Scholar] [CrossRef]
  286. NIST, [United States National Institute of Standards and Technology]. CODATA [Committee on Data for Science and Technology] Value: Molar Gas Constant. Available online: https://physics.nist.gov/cgi-bin/cuu/Value?r (accessed on 7 May 2022).
  287. NIST, [United States National Institute of Standards and Technology]. NIST Chemistry WebBook—Nitrogen. Available online: https://webbook.nist.gov/cgi/cbook.cgi?Name=n2 (accessed on 7 May 2022).
  288. Mallette, M.D. Sustainable Development Through the Use of Bio-Fuels and Quantitative Metrics—Blacklight. Master of Science, Environmental Pollution Control; Pennsylvania State University: Pennsylvania, PA, USA, 2008. [Google Scholar]
  289. Marzouk, O.A. Benchmarks for the Omani Higher Education Students-Faculty Ratio (SFR) Based on World Bank Data, QS Rankings, and THE Rankings. Cogent Educ. 2024, 11, 2317117. [Google Scholar] [CrossRef]
  290. Utamura, M.; Nikitin, K.; Kato, Y. A Generalised Mean Temperature Difference Method for Thermal Design of Heat Exchangers. Int. J. Nucl. Energy Sci. Technol. 2008, 4, 11. [Google Scholar] [CrossRef]
  291. Cao, Y. Equivalent Enthalpy Model and Log-Mean Enthalpy Difference Method for Heat-Mass Transfer in Cooling Towers. ASME J. Heat Mass Transf. 2024, 146, 104501. [Google Scholar] [CrossRef]
  292. Hurd, N.L. Mean Temperature Difference in the Field or Bayonet Tube. Ind. Eng. Chem. 1946, 38, 1266–1271. [Google Scholar] [CrossRef]
  293. Fakheri, A. Alternative Approach for Determining Log Mean Temperature Difference Correction Factor and Number of Shells of Shell and Tube Heat Exchangers. JEH(T) 2003, 10, 407–420. [Google Scholar] [CrossRef]
  294. Lu, G.; Wang, J.; Li, Z. Investigation on Calculating Methods of Mean Temperature Differences; American Society of Mechanical Engineers Digital Collection: New York, NY, USA, 2009; pp. 33–36. [Google Scholar]
  295. Bowman, R.A.; Mueller, A.C.; Nagle, W.M. Mean Temperature Difference in Design. Trans. Am. Soc. Mech. Eng. 2023, 62, 283–293. [Google Scholar] [CrossRef]
  296. Almeshaal, M.A.; Choubani, K. Using the Log Mean Temperature Difference (LMTD) and ε-NTU Methods to Analyze Heat and Mass Transfer in Direct Contact Membrane Distillation. Membranes 2023, 13, 588. [Google Scholar] [CrossRef] [PubMed]
  297. Lienhard, J.H., IV; Lienhard, J.H., V. A Heat Transfer Textbook (Fifth Edition), 5th ed.; Phlogiston Press: Cambridge, MA, USA, 2019. [Google Scholar]
  298. Quade, K.L.; Jöst, D.; Sauer, D.U.; Li, W. Understanding the Energy Potential of Lithium-Ion Batteries: Definition and Estimation of the State of Energy. Batter. Supercaps 2023, 6, e202300152. [Google Scholar] [CrossRef]
  299. Kanoğlu, M.; Çengel, Y.A.; Dinçer, İ. Energy Conversion Efficiencies. In Efficiency Evaluation of Energy Systems; Kanoğlu, M., Çengel, Y.A., Dinçer, İ., Eds.; Springer: New York, NY, USA, 2012; pp. 55–68. ISBN 978-1-4614-2242-6. [Google Scholar]
  300. Yin, L.; Ingesson, G.; Tunestal, P.; Johansson, R.; Johansson, B. An Experimental Investigation of a Multi-Cylinder Engine with Gasoline-Like Fuel towards a High Engine Efficiency; SAE Technical Paper: Detroit, MI, USA, 2016; p. 2016–01–0763. [Google Scholar]
  301. Marzouk, O.A. Energy Generation Intensity (EGI) for Parabolic Dish/Engine Concentrated Solar Power in Muscat, Sultanate of Oman. IOP Conf. Ser. Earth Environ. Sci. 2022, 1008, 012013. [Google Scholar] [CrossRef]
  302. Idris, K.A.; Ramdan, M.I.; Idroas, M.Y. Optimising the Performance of 110cc Engines for Fuel-Efficient Vehicle Design in the Shell Eco-Marathon. Politek. Kolej Komuniti J. Eng. Technol. 2024, 9, 125–135. [Google Scholar]
  303. Caton, J.A. Maximum Efficiencies for Internal Combustion Engines: Thermodynamic Limitations. Int. J. Engine Res. 2018, 19, 1005–1023. [Google Scholar] [CrossRef]
  304. Bandyopadhyay, S. All Forms of Energy Are Equal, but Some Forms of Energy Are More Equal than Others. Clean Technol. Env. Policy 2021, 23, 2775–2776. [Google Scholar] [CrossRef]
  305. Demirbas, A.; Balubaid, M.A.; Basahel, A.M.; Ahmad, W.; Sheikh, M.H. Octane Rating of Gasoline and Octane Booster Additives. Pet. Sci. Technol. 2015, 33, 1190–1197. [Google Scholar] [CrossRef]
  306. Kubesh, J.; King, S.R.; Liss, W.E. Effect of Gas Composition on Octane Number of Natural Gas Fuels; SAE Technical Paper: San Francisco, CA, USA, 1992; p. 922359. [Google Scholar]
  307. Salgado-Conrado, L.; Lopez-Montelongo, A.; Alvarez-Macias, C.; Hernadez-Jaquez, J. Review of Heliodon Developments and Computational Tools for Building Shadow Analysis. Buildings 2022, 12, 627. [Google Scholar] [CrossRef]
  308. Ramadhan, O.M.; Al-Hyali, E.A. New Experimental and Theoretical Relation to Determine the Research Octane Number (Ron) of Authentic Aromatic Hydrocarbons Could Be Present in the Gasoline Fraction. Pet. Sci. Technol. 1999, 17, 623–635. [Google Scholar] [CrossRef]
  309. Wu, Z.; Zhang, G.; Wang, C.; Jin, S.; Ji, M.; Hu, C.; Shang, Q. Numerical Investigation on the Flame Propagation Process of Ammonia/Hydrogen Blends under Engine-Related Conditions. Int. J. Hydrogen Energy 2024, 60, 1041–1053. [Google Scholar] [CrossRef]
  310. Muhammed, T.; Tokay, B.; Conradie, A. Raising the Research Octane Number Using an Optimized Simulated Moving Bed Technology towards Greater Sustainability and Economic Return. Fuel 2023, 337, 126864. [Google Scholar] [CrossRef]
  311. Liu, H.; Wang, Z.; Wang, J.; He, X. Effects of Gasoline Research Octane Number on Premixed Low-Temperature Combustion of Wide Distillation Fuel by Gasoline/Diesel Blend. Fuel 2014, 134, 381–388. [Google Scholar] [CrossRef]
  312. Nikolaou, N.; Papadopoulos, C.E.; Gaglias, I.A.; Pitarakis, K.G. A New Non-Linear Calculation Method of Isomerisation Gasoline Research Octane Number Based on Gas Chromatographic Data. Fuel 2004, 83, 517–523. [Google Scholar] [CrossRef]
  313. Daly, S.R.; Niemeyer, K.E.; Cannella, W.J.; Hagen, C.L. Predicting Fuel Research Octane Number Using Fourier-Transform Infrared Absorption Spectra of Neat Hydrocarbons. Fuel 2016, 183, 359–365. [Google Scholar] [CrossRef]
  314. Luecke, J.; Zigler, B.T. Rapid Prediction of Fuel Research Octane Number and Octane Sensitivity Using the AFIDA Constant-Volume Combustion Chamber. Fuel 2021, 301, 120969. [Google Scholar] [CrossRef]
  315. Sun, X.; Zhang, F.; Liu, J.; Duan, X. Prediction of Gasoline Research Octane Number Using Multiple Feature Machine Learning Models. Fuel 2023, 333, 126510. [Google Scholar] [CrossRef]
  316. Leonzio, G. Methanol Synthesis: Optimal Solution for a Better Efficiency of the Process. Processes 2018, 6, 20. [Google Scholar] [CrossRef]
  317. Tan, E.C.D.; Talmadge, M.; Dutta, A.; Hensley, J.; Schaidle, J.; Biddy, M.; Humbird, D.; Snowden-Swan, L.J.; Ross, J.; Sexton, D.; et al. Process Design and Economics for the Conversion of Lignocellulosic Biomass to Hydrocarbons via Indirect Liquefaction Thermochemical Research Pathway to High-Octane Gasoline Blendstock Through Methanol/Dimethyl Ether Intermediates; National Renewable Energy Laboratory (NREL) and Pacific Northwest National Laboratory (PNNL) of the United States Department of Energy: Richland, WA, USA, 2015.
  318. Vita, A.; Italiano, C. Fuel and Hydrogen Related Problems for Conventional Steam Reforming of Natural Gas. In Current Trends and Future Developments on (Bio-) Membranes; Elsevier: Amsterdam, The Netherlands, 2020; pp. 71–89. ISBN 978-0-12-816778-6. [Google Scholar]
  319. Najjar, R.P.; Teikari, P.; Cornut, P.-L.; Knoblauch, K.; Cooper, H.M.; Gronfier, C. Heterochromatic Flicker Photometry for Objective Lens Density Quantification. Investig. Ophthalmol. Vis. Sci. 2016, 57, 1063–1071. [Google Scholar] [CrossRef] [PubMed]
  320. Herbst, R.S.; Peterman, D.R.; Tillotson, R.D.; Delmau, L.H. Aspects of the Fundamental Chemistry of Cesium Extraction from Acidic Media by Hccd. Czech J. Phys. 2006, 56, D477–D482. [Google Scholar] [CrossRef]
  321. Marzouk, O.A. Lookup Tables for Power Generation Performance of Photovoltaic Systems Covering 40 Geographic Locations (Wilayats) in the Sultanate of Oman, with and without Solar Tracking, and General Perspectives about Solar Irradiation. Sustainability 2021, 13, 13209. [Google Scholar] [CrossRef]
  322. Heumann, C.; Schomaker, M. Shalabh Introduction to Statistics and Data Analysis: With Exercises, Solutions and Applications in r, 2nd ed.; Springer: Cham, Switzerland, 2023; ISBN 978-3-031-12025-1. [Google Scholar]
  323. Hamilton, D.F.; Ghert, M.; Simpson, A.H.R.W. Interpreting Regression Models in Clinical Outcome Studies. Bone Jt. Res. 2015, 4, 152–153. [Google Scholar] [CrossRef]
  324. Kim, Y.; van der Pol, E.; Arafa, A.; Thapa, I.; Britton, C.J.; Kosti, J.; Song, S.; Joshi, V.B.; Erickson, R.M.; Ali, H.; et al. Calibration and Standardization of Extracellular Vesicle Measurements by Flow Cytometry for Translational Prostate Cancer Research. Nanoscale 2022, 14, 9781–9795. [Google Scholar] [CrossRef] [PubMed]
  325. Kulik, M.M.; Yaklich, R.W. Evaluation of Vigor Tests in Soybean Seeds: Relationship of Accelerated Aging, Cold, Sand Bench, and Speed of Germination Test to Field Performance. Crop Sci. 1982, 22, 766–770. [Google Scholar] [CrossRef]
  326. Lewis-Beck, M.S.; Skalaban, A. The R-Squared: Some Straight Talk. Political Anal. 1990, 2, 153–171. [Google Scholar] [CrossRef]
  327. Jenkins, D.G.; Quintana-Ascencio, P.F. A Solution to Minimum Sample Size for Regressions. PLoS ONE 2020, 15, e0229345. [Google Scholar] [CrossRef]
  328. Menard, S. Coefficients of Determination for Multiple Logistic Regression Analysis. Am. Stat. 2000, 54, 17–24. [Google Scholar] [CrossRef]
  329. Marzouk, O.A.; Huckaby, E.D. A Comparative Study of Eight Finite-Rate Chemistry Kinetics for CO/H2 Combustion. Eng. Appl. Comput. Fluid Mech. 2010, 4, 331–356. [Google Scholar] [CrossRef]
  330. Wang, S.; Lei, X.; Xu, B.; Jiang, W.; Kong, L.; Yang, B.; Tian, Y.; Liu, Y. Vacuum Evaporation and Condensation Thermodynamics and Evaporation Kinetics of Pure Silver. Next Mater. 2024, 4, 100189. [Google Scholar] [CrossRef]
  331. Han, X.; Li, G.; Xu, M.; Guo, C.; Wang, Y.; Feng, Y.; Li, D. Miniature Capacitance Diaphragm Gauge for Absolute Vacuum Measurement. Measurement 2022, 194, 110851. [Google Scholar] [CrossRef]
  332. Bello, I. Vacuum and Ultravacuum: Physics and Technology; CRC Press: Boca Raton, FL, USA, 2017; ISBN 978-1-315-15536-4. [Google Scholar]
  333. Adam, T.; Hashim, U. COMSOL Multiphysics Simulation in Biomedical Engineering. Adv. Mater. Res. 2014, 832, 511–516. [Google Scholar] [CrossRef]
  334. Zimmerman, W.B.J. Introduction to Comsol Multiphysics. In Multiphysics Modeling with Finite Element Methods; Series on Stability, Vibration and Control of Systems, Series A.; World Scientific: Singapore, 2006; Volume 18, pp. 1–26. ISBN 978-981-256-843-4. [Google Scholar]
  335. Marzouk, O.A.; Nayfeh, A.H. Fluid Forces and Structure-Induced Damping of Obliquely-Oscillating Offshore Structures. In Proceedings of the the Eighteenth (2008) International Offshore and Polar Engineering Conference (ISOPE-2008), Vancouver, BC, Canada, 6 July 2008; Volume 3, pp. 460–468. [Google Scholar] [CrossRef]
  336. Rybdylova, O.; Qubeissi, M.A.; Braun, M.; Crua, C.; Manin, J.; Pickett, L.M.; de Sercey, G.; Sazhina, E.M.; Sazhin, S.S.; Heikal, M. A Model for Droplet Heating and Its Implementation into ANSYS Fluent. Int. Commun. Heat Mass Transf. 2016, 76, 265–270. [Google Scholar] [CrossRef]
  337. Ekambara, K.; Sanders, R.S.; Nandakumar, K.; Masliyah, J.H. Hydrodynamic Simulation of Horizontal Slurry Pipeline Flow Using ANSYS-CFX. Ind. Eng. Chem. Res. 2009, 48, 8159–8171. [Google Scholar] [CrossRef]
  338. McLure, F.I.; Tang, K.-S.; Williams, P.J. What Do Integrated STEM Projects Look like in Middle School and High School Classrooms? A Systematic Literature Review of Empirical Studies of iSTEM Projects. Int. J. STEM Educ. 2022, 9, 73. [Google Scholar] [CrossRef]
  339. Hall, A.; Miro, D. A Study of Student Engagement in Project-Based Learning Across Multiple Approaches to STEM Education Programs. Sch. Sci. Math. 2016, 116, 310–319. [Google Scholar] [CrossRef]
  340. Marzouk, O.A. Globalization and Diversity Requirement in Higher Education. In Proceedings of the 11th World Multi-Conference on Systemics, Cybernetics and Informatics (WMSCI 2007)—The 13th International Conference on Information Systems Analysis and Synthesis (ISAS 2007), Orlando, FL, USA, 8–11 July 2007; Volume III, pp. 101–106. [Google Scholar] [CrossRef]
  341. Mun, R.U.; Hertzog, N.B. Teaching and Learning in STEM Enrichment Spaces: From Doing Math to Thinking Mathematically. Roeper Rev. 2018, 40, 121–129. [Google Scholar] [CrossRef]
  342. Boelter, C.; Link, T.C.; Perry, B.L.; Leukefeld, C. Diversifying the STEM Pipeline. J. Educ. Stud. Placed Risk (JESPAR) 2015, 20, 218–237. [Google Scholar] [CrossRef]
  343. Alicat Scientific Alicat Flow Controller—MC · MCW · MCR · MCV; Alicat Scientific, Inc. (a Halma Company): Tucson, AZ, USA, 2020.
  344. Arshi, N.; Lu, J.; Joo, Y.K.; Lee, C.G.; Yoon, J.H.; Ahmed, F. Study on Structural, Morphological and Electrical Properties of Sputtered Titanium Nitride Films under Different Argon Gas Flow. Mater. Chem. Phys. 2012, 134, 839–844. [Google Scholar] [CrossRef]
  345. Ralchenko, V.; Sychov, I.; Vlasov, I.; Vlasov, A.; Konov, V.; Khomich, A.; Voronina, S. Quality of Diamond Wafers Grown by Microwave Plasma CVD: Effects of Gas Flow Rate. Diam. Relat. Mater. 1999, 8, 189–193. [Google Scholar] [CrossRef]
  346. Rombouts, C.; Bair, M.; Barbe, J.; Wright, J.D.; Kramer, R.; Krajicek, Z. A Comparison of Primary Gas Flow Standards Spanning the Range from 10 Sccm N2 to 10 Slm N2. NCSLI Meas. 2014, 9, 46–54. [Google Scholar] [CrossRef]
  347. Ahmadi, A.; Hadipour, N.L.; Kamfiroozi, M.; Bagheri, Z. Theoretical Study of Aluminum Nitride Nanotubes for Chemical Sensing of Formaldehyde. Sens. Actuators B Chem. 2012, 161, 1025–1029. [Google Scholar] [CrossRef]
  348. Berg, R.F.; Gooding, T.; Vest, R.E. Constant Pressure Primary Flow Standard for Gas Flows from 0.01cm3/Min to 100cm3/Min (0.007–74 μmol/s). Flow Meas. Instrum. 2014, 35, 84–91. [Google Scholar] [CrossRef]
  349. Fillet, R.; Nicolas, V.; Celzard, A.; Fierro, V. Solar Evaporation Performance of 3D-Printed Concave Structures Filled with Activated Carbon under Low Convective Flow. Chem. Eng. J. 2023, 457, 141168. [Google Scholar] [CrossRef]
  350. Jing, J.; Yang, L.; Tang, X.; He, P.; Tang, K. Numerical Simulation of Multiphase Flow Erosion in the Gas Well Relief Line Elbow under Supercritical Conditions. J. Pipeline Syst. Eng. Pract. 2023, 14, 04023031. [Google Scholar] [CrossRef]
  351. Shahid, M.; Chambers, B.; Sankarasubramanian, S. Methane and Oxygen from Energy-efficient, Low Temperature in Situ Resource Utilization Enables Missions to Mars. AIChE J. 2023, 69, e18010. [Google Scholar] [CrossRef]
  352. Cao, Y.; Liang, W.; Law, C.K. Real Gas Effects in High-Pressure Ignition of n -Dodecane/Air Mixtures. J. Phys. Chem. A 2024, 128, 3604–3612. [Google Scholar] [CrossRef]
  353. Ding, H.; Zhang, Y.; Dong, Y.; Wen, C.; Yang, Y. High-Pressure Supersonic Carbon Dioxide (CO2) Separation Benefiting Carbon Capture, Utilisation and Storage (CCUS) Technology. Appl. Energy 2023, 339, 120975. [Google Scholar] [CrossRef]
  354. Gaganis, V.; Homouz, D.; Maalouf, M.; Khoury, N.; Polychronopoulou, K. An Efficient Method to Predict Compressibility Factor of Natural Gas Streams. Energies 2019, 12, 2577. [Google Scholar] [CrossRef]
  355. Kartal, M.A.; Atakök, G.; Ersoy, S. Cooling and Multiphase Analysis of Heated Environmentally Friendly R152A (C2H4F2) Fluid Coming from the Production Process According to Nist Indicators. Appl. Sci. 2024, 14, 4143. [Google Scholar] [CrossRef]
  356. Manfredi, M.; Persico, G.; Spinelli, A.; Gaetani, P.; Dossena, V. Design and Commissioning of Experiments for Supersonic ORC Nozzles in Linear Cascade Configuration. Appl. Therm. Eng. 2023, 224, 119996. [Google Scholar] [CrossRef]
  357. Ruiz Maraggi, L.M.; Moscardelli, L.G. Modeling Hydrogen Storage Capacities, Injection and Withdrawal Cycles in Salt Caverns: Introducing the GeoH2 Salt Storage and Cycling App. Int. J. Hydrogen Energy 2023, 48, 26921–26936. [Google Scholar] [CrossRef]
  358. Srinivasan, N.; Zhang, H.; Yang, S. VLE-Based Phase Field Method to Simulate High-Pressure Diffuse Interface with Phase Change. In Proceedings of the AIAA SCITECH 2024 Forum, Orlando, FL, USA, 8 January 2024. [Google Scholar]
  359. Mahmoud, M. Development of a New Correlation of Gas Compressibility Factor (Z-Factor) for High Pressure Gas Reservoirs. J. Energy Resour. Technol. 2013, 136, 012903. [Google Scholar] [CrossRef]
  360. Jung, H.-Y.; Kim, M.S.; Ko, A.-R.; Lee, J.I. Investigation of CO2 Leak Accident in SFR Coupled with S-CO2 Brayton Cycle. Ann. Nucl. Energy 2017, 103, 212–226. [Google Scholar] [CrossRef]
  361. Marzouk, O.A. The Sod Gasdynamics Problem as a Tool for Benchmarking Face Flux Construction in the Finite Volume Method. Sci. Afr. 2020, 10, e00573. [Google Scholar] [CrossRef]
  362. Kim, M.S.; Oh, B.S.; Jung, H.-Y.; Bae, S.J.; Lee, J.I. Experimental and Numerical Study of Critical Flow Model Development for Supercritical CO2 Power Cycle Application; American Society of Mechanical Engineers Digital Collection: New York, NY, USA, 2018. [Google Scholar]
  363. Kim, M.S.; Bae, S.J.; Son, S.; Oh, B.S.; Lee, J.I. Study of Critical Flow for Supercritical CO2 Seal. Int. J. Heat Mass Transf. 2019, 138, 85–95. [Google Scholar] [CrossRef]
  364. Yan, P.; Gori, G.; Zocca, M.; Guardone, A. SU2-COOL: Open-Source Framework for Non-Ideal Compressible Fluid Dynamics. Comput. Phys. Commun. 2025, 307, 109394. [Google Scholar] [CrossRef]
  365. Kandula, M.; Nufer, B.M.; Felt, A.M.; Hicks, N.G.; Aranyos, T.J.; Webster, G.K.; Coll, G.T. Compressibility, Reynolds Number and Thermal Effects in Pressure-Decay Based Leak Detection Systems. In Proceedings of the AIAA Aviation 2019 Forum, Dallas, TX, USA, 17–21 June 2019; p. AIAA 2019-3437. [Google Scholar]
  366. Sarmast, S.; Rouindej, K.; Fraser, R.A.; Dusseault, M.B. Sizing-Design Method for Compressed Air Energy Storage (CAES) Systems: A Case Study Based on Power Grid in Ontario. Energy Convers. Manag. 2023, 277, 116656. [Google Scholar] [CrossRef]
  367. Bischoff, J.L.; Rosenbauer, R.J.; Pitzer, K.S. The System NaCl-H2O: Relations of Vapor-Liquid near the Critical Temperature of Water and of Vapor-Liquid-Halite from 300° to 500 °C. Geochim. Cosmochim. Acta 1986, 50, 1437–1444. [Google Scholar] [CrossRef]
  368. Dauchot, J.-P.; Dath, J.-P. Oxidation of Carbon Monoxide on Thin Film Catalysts: Characterization by a Critical Temperature Measurement. J. Catal. 1984, 86, 373–383. [Google Scholar] [CrossRef]
  369. Guo, G.; Wang, F.; Liu, G.-Q.; Luo, S.-J.; Guo, R.-B. Calculation on the Phase Equilibrium and Critical Temperature of CH4/CO2. Process Saf. Environ. Prot. 2018, 113, 369–377. [Google Scholar] [CrossRef]
  370. Ingebritsen, S.E.; Hayba, D.O. Fluid Flow and Heat Transport near the Critical Point of H 2 O. Geophys. Res. Lett. 1994, 21, 2199–2202. [Google Scholar] [CrossRef]
  371. Karnaukhov, V.A.; Oeschler, H.; Avdeyev, S.P.; Duginova, E.V.; Rodionov, V.K.; Budzanowski, A.; Karcz, W.; Bochkarev, O.V.; Kuzmin, E.A.; Chulkov, L.V.; et al. Critical Temperature for the Nuclear Liquid-Gas Phase Transition. Phys. Rev. C 2003, 67, 011601. [Google Scholar] [CrossRef]
  372. Kestin, J.; Korfali, Ö.; Sengers, J.V. Density Expansion of the Viscosity of Carbon Dioxide near the Critical Temperature. Phys. A Stat. Mech. Its Appl. 1980, 100, 335–348. [Google Scholar] [CrossRef]
  373. Dincer, I.; Zamfirescu, C. 1.5 Thermodynamic Aspects of Energy. In Comprehensive Energy Systems; Elsevier: Amsterdam, The Netherlands, 2018; pp. 153–211. ISBN 978-0-12-814925-6. [Google Scholar]
  374. Tiab, D.; Donaldson, E.C. Introduction to Petroleum Geology. In Petrophysics; Elsevier: Amsterdam, The Netherlands, 2016; pp. 23–66. ISBN 978-0-12-803188-9. [Google Scholar]
  375. Sankar Behera, U.; Kumar Prasad, S.; Byun, H.-S. Experimental Validation on the Phase Separation for the 2-(Diisopropylamino)Ethyl Methacrylate and Poly[2-(Diisopropylamino)Ethyl Methacrylate] in Supercritical CO2. J. Mol. Liq. 2024, 393, 123553. [Google Scholar] [CrossRef]
  376. Shang, Z.; Yang, Y.; Zhang, L.; Sun, H.; Zhong, J.; Zhang, K.; Yao, J. Hydrogen Adsorption and Diffusion Behavior in Kaolinite Slit for Underground Hydrogen Storage: A Hybrid GCMC-MD Simulation Study. Chem. Eng. J. 2024, 487, 150517. [Google Scholar] [CrossRef]
  377. Shen, Y.; Qiu, C.; Liu, D.; Tao, X.; Wan, A.; Zhang, Z.; Gan, Z. Experimental Study on a Closed-Cycle Joule-Thomson Cryocooler Working at Liquid Hydrogen Temperature. Appl. Therm. Eng. 2023, 234, 121291. [Google Scholar] [CrossRef]
  378. Wei-yu, C.; Sun, L.; Zhou, J.; Li, X.; Huang, L.; Xia, G.; Meng, X.; Wang, K. Toward Predicting Interfacial Tension of Impure and Pure CO2-Brine Systems Using Robust Correlative Approaches. ACS Omega 2024, 9, 7937–7957. [Google Scholar] [CrossRef]
  379. Xie, Y.; Dong, H.; Zhang, S.; Lu, X.; Ji, X. Solubilities of CO2, CH4, H2, CO and N2 in Choline Chloride/Urea. Green Energy Environ. 2016, 1, 195–200. [Google Scholar] [CrossRef]
  380. Mihara, S.; Sagara, H.; Arai, Y.; Saito, S. The Compressibility Factors of Hydrogen-Methane, Hydrogen-Ethane and Hydrogen-Propane Gaseous Mixtures. J. Chem. Eng. Jpn./JCEJ 1977, 10, 395–399. [Google Scholar] [CrossRef]
  381. White, C.W.; Weiland, N.T. Evaluation of Property Methods for Modeling Direct-Supercritical CO2 Power Cycles. J. Eng. Gas Turbines Power 2017, 140, 011701. [Google Scholar] [CrossRef]
  382. Seneviratne, K.N.; Hughes, T.J.; Johns, M.L.; Marsh, K.N.; May, E.F. Surface Tension and Critical Point Measurements of Methane + Propane Mixtures. J. Chem. Thermodyn. 2017, 111, 173–184. [Google Scholar] [CrossRef]
  383. Rane, S.; Kovačević, A.; Stošić, N.; Smith, I. Analysis of Real Gas Equation of State for CFD Modelling of Twin Screw Expanders with R245fa, R290, R1336mzz(Z) and R1233zd(E). Int. J. Refrig. 2021, 121, 313–326. [Google Scholar] [CrossRef]
  384. Clementoni, E.M.; Cox, T.L. Comparison of Carbon Dioxide Property Measurements for an Operating Supercritical Brayton Cycle to the REFPROP Physical Property Database; American Society of Mechanical Engineers Digital Collection: New York, NY, USA, 2014. [Google Scholar]
  385. Obein, G.; Bousquet, R.; Nadal, M.E. New NIST Reference Goniospectrometer. In Optical Diagnostics; SPIE: Pune, India, 2005; Volume 5880, pp. 241–250. [Google Scholar]
  386. Pearce, N.J.G.; Perkins, W.T.; Westgate, J.A.; Gorton, M.P.; Jackson, S.E.; Neal, C.R.; Chenery, S.P. A Compilation of New and Published Major and Trace Element Data for NIST SRM 610 and NIST SRM 612 Glass Reference Materials. Geostand. Newsl. 1997, 21, 115–144. [Google Scholar] [CrossRef]
  387. Lafarge, T.; Possolo, A. The NIST Uncertainty Machine. NCSLI Meas. 2015, 10, 20–27. [Google Scholar] [CrossRef]
  388. Ferraiolo, D.F.; Sandhu, R.; Gavrila, S.; Kuhn, D.R.; Chandramouli, R. Proposed NIST Standard for Role-Based Access Control. ACM Trans. Inf. Syst. Secur. 2001, 4, 224–274. [Google Scholar] [CrossRef]
  389. Linstrom, P.J.; Mallard, W.G. The NIST Chemistry WebBook:  A Chemical Data Resource on the Internet. J. Chem. Eng. Data 2001, 46, 1059–1063. [Google Scholar] [CrossRef]
  390. PNNL, [Pacific Northwest National Laboratory of the United States Department of Energy]. H2 Tools—Hydrogen Compressibility at Different Temperatures and Pressures. Available online: https://h2tools.org/hyarc/hydrogen-data/hydrogen-compressibility-different-temperatures-and-pressures (accessed on 20 September 2022).
  391. Todić, B.; Pravilović, R.; Nikačević, N. Design and Simulations of a Helical Oscillatory Baffled Reactor for Biochemical Reactions. Chem. Eng. Process. -Process Intensif. 2024, 203, 109895. [Google Scholar] [CrossRef]
  392. Marzouk, O.A.; Nayfeh, A.H. Hydrodynamic Forces on a Moving Cylinder with Time-Dependent Frequency Variations. In Proceedings of the 46th AIAA Aerospace Sciences Meeting and Exhibit, Reno, NV, USA, 7–10 January 2008; p. AIAA 2008-680. [Google Scholar] [CrossRef]
  393. Wang, J.; Shen, L.; Wang, L.; Sundén, B. Effect of Hydrogen Permeation and Operating Parameters on Thermochemical Performance of Solar-Driven Steam Methane Reforming Membrane Reactor. Chem. Eng. Sci. 2024, 295, 120157. [Google Scholar] [CrossRef]
  394. CCPS, [Center for Chemical Process Safety of the American Institute of Chemical Engineers]. Plug Flow Reactor (PFR). Available online: https://www.aiche.org/ccps/resources/glossary/process-safety-glossary/plug-flow-reactor-pfr (accessed on 20 September 2022).
  395. Ming, D.; Glasser, D.; Hildebrandt, D.; Glasser, B.; Metzer, M. Attainable Region Theory: An Introduction to Choosing an Optimal Reactor; Wiley: Hoboken, NJ, USA, 2016; ISBN 978-1-119-24470-7. [Google Scholar]
  396. Rosa, D.; Goes, P.; Manzi, J. Steady-State Plug Flow Reactor Analysis by Means of Minimum Entropy. In Computer Aided Chemical Engineering; Elsevier: Amsterdam, The Netherlands, 2017; Volume 40, pp. 1165–1170. ISBN 978-0-444-63965-3. [Google Scholar]
  397. Curtis, R.; Nguyen, C.; Lower, S. 2.3: First-Order Reactions. Available online: https://chem.libretexts.org/Bookshelves/Physical_and_Theoretical_Chemistry_Textbook_Maps/Supplemental_Modules__(Physical_and_Theoretical_Chemistry)/Kinetics/02%3A_Reaction_Rates/2.03%3A_First-Order_Reactions (accessed on 20 September 2022).
  398. Phuakpunk, K.; Chalermsinsuwan, B.; Assabumrungrat, S. Comparison of Chemical Reaction Kinetic Models for Corn Cob Pyrolysis. Energy Rep. 2020, 6, 168–178. [Google Scholar] [CrossRef]
  399. Marzouk, O.A.; Nayfeh, A.H. New Wake Models with Capability of Capturing Nonlinear Physics. In Proceedings of the ASME 2008 27th International Conference on Offshore Mechanics and Arctic Engineering (OMAE 2008), Estoril, Portugal, 27 July 2009; pp. 901–912. [Google Scholar] [CrossRef]
  400. Marzouk, O.A.; Nayfeh, A.H. A Study of the Forces on an Oscillating Cylinder. In Proceedings of the ASME 2007 26th International Conference on Offshore Mechanics and Arctic Engineering (OMAE 2007), San Diego, CA, USA, 20 May 2009; pp. 741–752. [Google Scholar] [CrossRef]
  401. AIChE, [American Institute of Chemical Engineers]. Continuous Reactors. Available online: https://www.aiche.org/ccps/resources/glossary/process-safety-glossary/continuous-reactors (accessed on 6 May 2025).
  402. Tuckerman, M.E. Lecture 25: Plug Flow Reactors and Comparison to Continuously Stirred Tank Reactors. Available online: https://chem.libretexts.org/Courses/New_York_University/CHEM-UA_652%3A_Thermodynamics_and_Kinetics/Lecture_25%3A_Plug_flow_reactors_and_comparison_to_continuously_stirred_tank_reactors (accessed on 9 May 2022).
  403. Marzouk, O.A. Coupled Differential-Algebraic Equations Framework for Modeling Six-Degree-of-Freedom Flight Dynamics of Asymmetric Fixed-Wing Aircraft. Int. J. Appl. Adv. Sci. 2025, 12, 30–51. [Google Scholar] [CrossRef]
  404. Brencio, C.; Fontein, F.W.A.; Medrano, J.A.; Di Felice, L.; Arratibel, A.; Gallucci, F. Pd-Based Membranes Performance under Hydrocarbon Exposure for Propane Dehydrogenation Processes: Experimental and Modeling. Int. J. Hydrogen Energy 2022, 47, 11369–11384. [Google Scholar] [CrossRef]
  405. Campo, M.; Tanaka, A.; Mendes, A.; Sousa, J.M. Characterization of Membranes for Energy and Environmental Applications. In Advanced Membrane Science and Technology for Sustainable Energy and Environmental Applications; Elsevier: Amsterdam, The Netherlands, 2011; pp. 56–89. ISBN 978-1-84569-969-7. [Google Scholar]
  406. Caravella, A.; Hara, S.; Drioli, E.; Barbieri, G. Sieverts Law Pressure Exponent for Hydrogen Permeation through Pd-Based Membranes: Coupled Influence of Non-Ideal Diffusion and Multicomponent External Mass Transfer. Int. J. Hydrogen Energy 2013, 38, 16229–16244. [Google Scholar] [CrossRef]
  407. Olander, D.; Konashi, K.; Yamawaki, M. Uranium–Zirconium Hydride Fuel. In Comprehensive Nuclear Materials; Elsevier: Amsterdam, The Netherlands, 2012; pp. 313–357. ISBN 978-0-08-056033-5. [Google Scholar]
  408. Suzuki, A.; Yukawa, H. A Review for Consistent Analysis of Hydrogen Permeability through Dense Metallic Membranes. Membranes 2020, 10, 120. [Google Scholar] [CrossRef] [PubMed]
  409. Koffler, S.A.; Hudson, J.B.; Ansell, G.S. Hydrogen Permeation through Alpha-Palladium. Trans. Metall. Soc. Am. Inst. Min. Metall. Pet. Eng. 1969, 245, 1735–1740. [Google Scholar]
  410. Morreale, B.D.; Ciocco, M.V.; Enick, R.M.; Morsi, B.I.; Howard, B.H.; Cugini, A.V.; Rothenberger, K.S. The Permeability of Hydrogen in Bulk Palladium at Elevated Temperatures and Pressures. J. Membr. Sci. 2003, 212, 87–97. [Google Scholar] [CrossRef]
  411. Yuan, M.; Lee, K.; Van Campen, D.G.; Liguori, S.; Toney, M.F.; Wilcox, J. Hydrogen Purification in Palladium-Based Membranes: An Operando X-Ray Diffraction Study. Ind. Eng. Chem. Res. 2019, 58, 926–934. [Google Scholar] [CrossRef]
  412. Marzouk, O.A. Performance Analysis of Shell-and-Tube Dehydrogenation Module. Int. J. Energy Res. 2017, 41, 604–610. [Google Scholar] [CrossRef]
  413. Nordio, M.; Wassie, S.A.; Van Sint Annaland, M.; Pacheco Tanaka, D.A.; Viviente Sole, J.L.; Gallucci, F. Techno-Economic Evaluation on a Hybrid Technology for Low Hydrogen Concentration Separation and Purification from Natural Gas Grid. Int. J. Hydrogen Energy 2021, 46, 23417–23435. [Google Scholar] [CrossRef]
  414. Marzouk, O.A. Levelized Cost of Green Hydrogen (LCOH) in the Sultanate of Oman Using H2A-Lite with Polymer Electrolyte Membrane (PEM) Electrolyzers Powered by Solar Photovoltaic (PV) Electricity. E3S Web Conf. 2023, 469, 00101. [Google Scholar] [CrossRef]
  415. Microsoft Corporation Use Goal Seek to Find the Result You Want by Adjusting an Input Value. Available online: https://support.microsoft.com/en-us/office/use-goal-seek-to-find-the-result-you-want-by-adjusting-an-input-value-320cb99e-f4a4-417f-b1c3-4f369d6e66c7 (accessed on 11 May 2022).
Figure 1. Inlet conditions and flow directions for retentate and permeate streams.
Figure 1. Inlet conditions and flow directions for retentate and permeate streams.
Processes 13 01455 g001
Figure 2. The proposed cross-section of the shell-and-tube membrane reactor (all dimensions are in cm).
Figure 2. The proposed cross-section of the shell-and-tube membrane reactor (all dimensions are in cm).
Processes 13 01455 g002
Figure 3. Commutative hydrogen recovery for the base case, with two spatial resolutions (the 200 segments represent the normal resolution adopted in the study; the 400 segments represent the finer resolution for testing the resolution insensitivity).
Figure 3. Commutative hydrogen recovery for the base case, with two spatial resolutions (the 200 segments represent the normal resolution adopted in the study; the 400 segments represent the finer resolution for testing the resolution insensitivity).
Processes 13 01455 g003
Figure 4. Segmental hydrogen recovery for the base case, with two spatial resolutions (the 200 segments represent the normal resolution). There is a third curve added, with the segmental recovery values in the case of 400 segments being doubled.
Figure 4. Segmental hydrogen recovery for the base case, with two spatial resolutions (the 200 segments represent the normal resolution). There is a third curve added, with the segmental recovery values in the case of 400 segments being doubled.
Processes 13 01455 g004
Figure 5. Hydrogen mole fraction in the retentate stream for the base case, with two spatial resolutions (the 200 segments represent the normal resolution).
Figure 5. Hydrogen mole fraction in the retentate stream for the base case, with two spatial resolutions (the 200 segments represent the normal resolution).
Processes 13 01455 g005
Figure 6. The hydrogen mole fraction in the permeate stream for the base case, with two spatial resolutions (the 200 segments represent the normal resolution).
Figure 6. The hydrogen mole fraction in the permeate stream for the base case, with two spatial resolutions (the 200 segments represent the normal resolution).
Processes 13 01455 g006
Figure 7. The membrane length as a function of the temperature in the membrane reactor.
Figure 7. The membrane length as a function of the temperature in the membrane reactor.
Processes 13 01455 g007
Figure 8. The average hydrogen permeation mass flux as a function of the temperature in the membrane reactor.
Figure 8. The average hydrogen permeation mass flux as a function of the temperature in the membrane reactor.
Processes 13 01455 g008
Figure 9. Local actual permeance, global apparent permeance, and efficiency factor as functions of temperature in membrane reactor.
Figure 9. Local actual permeance, global apparent permeance, and efficiency factor as functions of temperature in membrane reactor.
Processes 13 01455 g009
Figure 10. The membrane length as a function of the retentate pressure in the membrane reactor.
Figure 10. The membrane length as a function of the retentate pressure in the membrane reactor.
Processes 13 01455 g010
Figure 11. The average hydrogen permeation mass flux as a function of the retentate pressure in the membrane reactor.
Figure 11. The average hydrogen permeation mass flux as a function of the retentate pressure in the membrane reactor.
Processes 13 01455 g011
Figure 12. The pressure-square-root difference at the left end of the membrane reactor (x = 0), the pressure-square-root difference at the right end of the membrane reactor (x = L), and the log mean pressure-square-root difference (LMPD) as functions of the retentate pressure in the membrane reactor.
Figure 12. The pressure-square-root difference at the left end of the membrane reactor (x = 0), the pressure-square-root difference at the right end of the membrane reactor (x = L), and the log mean pressure-square-root difference (LMPD) as functions of the retentate pressure in the membrane reactor.
Processes 13 01455 g012
Figure 13. Local actual permeance, global apparent permeance, and efficiency factor as functions of retentate pressure in membrane reactor.
Figure 13. Local actual permeance, global apparent permeance, and efficiency factor as functions of retentate pressure in membrane reactor.
Processes 13 01455 g013
Figure 14. The membrane length as a function of the permeate pressure in the membrane reactor.
Figure 14. The membrane length as a function of the permeate pressure in the membrane reactor.
Processes 13 01455 g014
Figure 15. The average hydrogen permeation mass flux as a function of the permeate pressure in the membrane reactor.
Figure 15. The average hydrogen permeation mass flux as a function of the permeate pressure in the membrane reactor.
Processes 13 01455 g015
Figure 16. The pressure-square-root difference at the left end of the membrane reactor (x = 0), the pressure-square-root difference at the right end of the membrane reactor (x = L), and the log mean pressure-square-root difference (LMPD) as functions of the permeate pressure in the membrane reactor.
Figure 16. The pressure-square-root difference at the left end of the membrane reactor (x = 0), the pressure-square-root difference at the right end of the membrane reactor (x = L), and the log mean pressure-square-root difference (LMPD) as functions of the permeate pressure in the membrane reactor.
Processes 13 01455 g016
Figure 17. Local actual permeance, global apparent permeance, and efficiency factor as functions of permeate pressure in membrane reactor.
Figure 17. Local actual permeance, global apparent permeance, and efficiency factor as functions of permeate pressure in membrane reactor.
Processes 13 01455 g017
Figure 18. Commutative hydrogen recovery with the extreme values considered for the temperature, absolute retentate pressure, and absolute permeate pressure.
Figure 18. Commutative hydrogen recovery with the extreme values considered for the temperature, absolute retentate pressure, and absolute permeate pressure.
Processes 13 01455 g018
Figure 19. Segmental hydrogen recovery with the extreme values considered for the temperature, absolute retentate pressure, and absolute permeate pressure.
Figure 19. Segmental hydrogen recovery with the extreme values considered for the temperature, absolute retentate pressure, and absolute permeate pressure.
Processes 13 01455 g019
Figure 20. The hydrogen mole fraction in the retentate stream with the extreme values considered for the temperature, absolute retentate pressure, and absolute permeate pressure.
Figure 20. The hydrogen mole fraction in the retentate stream with the extreme values considered for the temperature, absolute retentate pressure, and absolute permeate pressure.
Processes 13 01455 g020
Figure 21. The hydrogen mole fraction in the permeate stream with the extreme values considered for the temperature, absolute retentate pressure, and absolute permeate pressure.
Figure 21. The hydrogen mole fraction in the permeate stream with the extreme values considered for the temperature, absolute retentate pressure, and absolute permeate pressure.
Processes 13 01455 g021
Figure 22. A visualization of the importance of the three design variables covered in the current study (a priority design chart) showing that the temperature has the highest priority, while the permeate pressure has the least priority.
Figure 22. A visualization of the importance of the three design variables covered in the current study (a priority design chart) showing that the temperature has the highest priority, while the permeate pressure has the least priority.
Processes 13 01455 g022
Table 1. Some fixed geometric details about the proposed shell-and-tube membrane reactor.
Table 1. Some fixed geometric details about the proposed shell-and-tube membrane reactor.
Geometric FeatureValue
Shell diameter5.000 cm (1.969 in)
Tube diameter 1.250 cm (0.4921 in)
Number of tubes 8
Shell cross-section area (excluding tubes)9.817 cm2 (1.522 in2)
Tubes cross-section area (all 8 tubes)9.817 cm2 (1.522 in2)
Shell-to-tubes area ratio1:1
Tube cross-section area (single tube)1.227 cm2 (0.1902 in2)
Table 2. Some common fixed properties of the retentate stream.
Table 2. Some common fixed properties of the retentate stream.
ConditionValue
Inlet mole fraction, H230% [254]
Inlet mole fraction, CO50% [255]
Inlet mole fraction, CO220% [256,257]
Molecular weight, H22.01588 kg/kmol [258]
Molecular weight, CO28.0101 kg/kmol [259]
Molecular weight, CO244.0095 kg/kmol [260]
Molecular weight, mixture23.412 kg/kmol
Inlet mass fraction, H20.025832
Inlet mass fraction, CO0.598207
Inlet mass fraction, CO20.375961
Inlet mass flow rate60 kg/h (132.28 lbm/h)
Inlet standard volume flow rate970,068 sccm (standard cubic centimeters per minute)
Target hydrogen recovery95% (by mass, by mole, or by standard volume—identical)
Table 3. Some common fixed properties of the permeate stream.
Table 3. Some common fixed properties of the permeate stream.
ConditionValue
Inlet gas 100% N2
Molecular weight, N2 28.0134 kg/kmol [287]
Inlet mass flow rate 30.692 kg/h (67.664 lbm/h)
Inlet standard volume flow rate414,704 sccm (standard cubic centimeters per minute)
Target outlet mole fraction of H2 40%
Table 4. Temperature and pressures of base case (reference simulation).
Table 4. Temperature and pressures of base case (reference simulation).
Fluid Property Value
Temperature300 °C (572.00 °F)
Absolute retentate pressure40.0 atm (587.84 psia)
Absolute permeate pressure20.0 atm (293.92 psia)
Table 5. Some results obtained for the base case with 200 segments (normal resolution) and 400 segments (refined resolution for testing only).
Table 5. Some results obtained for the base case with 200 segments (normal resolution) and 400 segments (refined resolution for testing only).
Result ValueAbsolute Percentage Change
n = 200 Segmentsn = 400 Segments V a l u e 200 V a l u e 400 V a l u e 200
Membrane length (cm)407.359407.3590% (identical)
Average hydrogen permeation mass flux (kg/m2.h)1.1511.1500.01%
Pressure-square-root difference at left end (Pa0.5)202.345202.3450% (identical)
Pressure-square-root difference at right end (Pa0.5)260.655268.8963.16%
Log mean pressure-square-root difference (Pa0.5)230.271234.051.64%
Global apparent hydrogen permeance (mol/m2.s.Pa0.5)6.8849 × 10−46.7732 × 10−41.62%
Efficiency factor (%)67.09%66.00%1.62%
Hydrogen recovery (%)95.000%94.991%0.01%
Table 6. Some details about the extreme cases.
Table 6. Some details about the extreme cases.
Extreme Case Number Temperature Absolute Retentate PressureAbsolute Permeate Pressure
1800 °C (1472.00 °F)40.0 atm (587.84 psia)20.0 atm (293.92 psia)
2300 °C (572.00 °F)120.0 atm (1763.5 psia)20.0 atm (293.92 psia)
3300 °C (572.00 °F)40.0 atm (587.84 psia)0.20 atm (2.9392 psia)
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Marzouk, O.A. Reduced-Order Modeling (ROM) of a Segmented Plug-Flow Reactor (PFR) for Hydrogen Separation in Integrated Gasification Combined Cycles (IGCC). Processes 2025, 13, 1455. https://doi.org/10.3390/pr13051455

AMA Style

Marzouk OA. Reduced-Order Modeling (ROM) of a Segmented Plug-Flow Reactor (PFR) for Hydrogen Separation in Integrated Gasification Combined Cycles (IGCC). Processes. 2025; 13(5):1455. https://doi.org/10.3390/pr13051455

Chicago/Turabian Style

Marzouk, Osama A. 2025. "Reduced-Order Modeling (ROM) of a Segmented Plug-Flow Reactor (PFR) for Hydrogen Separation in Integrated Gasification Combined Cycles (IGCC)" Processes 13, no. 5: 1455. https://doi.org/10.3390/pr13051455

APA Style

Marzouk, O. A. (2025). Reduced-Order Modeling (ROM) of a Segmented Plug-Flow Reactor (PFR) for Hydrogen Separation in Integrated Gasification Combined Cycles (IGCC). Processes, 13(5), 1455. https://doi.org/10.3390/pr13051455

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop