Next Article in Journal
Fundamental Study of the Operational Characteristics of Recombination Catalysts for Hydrogen Risk Mitigation at Low Temperatures
Previous Article in Journal
Design of a Proton Exchange Membrane Electrolyzer
Previous Article in Special Issue
Green Hydrogen: Pathway to Net Zero Green House Gas Emission and Global Climate Change Mitigation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Power Density and Thermochemical Properties of Hydrogen Magnetohydrodynamic (H2MHD) Generators at Different Pressures, Seed Types, Seed Levels, and Oxidizers

College of Engineering, University of Buraimi, Al Buraimi 512, Sultanate of Oman
Hydrogen 2025, 6(2), 31; https://doi.org/10.3390/hydrogen6020031
Submission received: 5 April 2025 / Revised: 28 April 2025 / Accepted: 30 April 2025 / Published: 2 May 2025

Abstract

:
Hydrogen and some of its derivatives (such as e-methanol, e-methane, and e-ammonia) are promising energy carriers that have the potential to replace conventional fuels, thereby eliminating their harmful environmental impacts. An innovative use of hydrogen as a zero-emission fuel is forming weakly ionized plasma by seeding the combustion products of hydrogen with a small amount of an alkali metal vapor (cesium or potassium). This formed plasma can be used as a working fluid in supersonic open-cycle magnetohydrodynamic (OCMHD) power generators. In these OCMHD generators, direct-current (DC) electricity is generated straightforwardly without rotary turbogenerators. In the current study, we quantitatively and qualitatively explore the levels of electric conductivity and the resultant volumetric electric output power density in a typical OCMHD supersonic channel, where thermal equilibrium plasma is accelerated at a Mach number of two (Mach 2) while being subject to a strong applied magnetic field (applied magnetic-field flux density) of five teslas (5 T), and a temperature of 2300 K (2026.85 °C). We varied the total pressure of the pre-ionization seeded gas mixture between 1/16 atm and 16 atm. We also varied the seed level between 0.0625% and 16% (pre-ionization mole fraction). We also varied the seed type between cesium and potassium. We also varied the oxidizer type between air (oxygen–nitrogen mixture, 21–79% by mole) and pure oxygen. Our results suggest that the ideal power density can reach exceptional levels beyond 1000 MW/m3 (or 1 kW/cm3) provided that the total absolute pressure can be reduced to about 0.1 atm only and cesium is used for seeding rather than potassium. Under atmospheric air–hydrogen combustion (1 atm total absolute pressure) and 1% mole fraction of seed alkali metal vapor, the theoretical volumetric power density is 410.828 MW/m3 in the case of cesium and 104.486 MW/m3 in the case of potassium. The power density can be enhanced using any of the following techniques: (1) reducing the total pressure, (2) using cesium instead of potassium for seeding, and (3) using air instead of oxygen as an oxidizer (if the temperature is unchanged). A seed level between 1% and 4% (pre-ionization mole fraction) is recommended. Much lower or much higher seed levels may harm the OCMHD performance. The seed level that maximizes the electric power is not necessarily the same seed level that maximizes the electric conductivity, and this is due to additional thermochemical changes caused by the additive seed. For example, in the case of potassium seeding and air combustion, the electric conductivity is maximized with about 6% seed mole fraction, while the output power is maximized at a lower potassium level of about 5%. We also present a comprehensive set of computed thermochemical properties of the seeded combustion gases, such as the molecular weight and the speed of sound.

1. Introduction

1.1. Background

Hydrogen is the lightest element in the periodic table, but it is the most plentiful element in the universe [1,2] (oxygen is the most abundant element in the earth’s crust by mass [3,4]). Hydrogen is an important building block in several chemical processes and power systems [5,6], as listed in Table 1.
Hydrogen facilitates the transition to a sustainable (green) low-emission economy [95,96,97,98], taking advantage of the absence of released carbon dioxide (CO2) due to the consumption of hydrogen [99,100,101,102]. The production of green hydrogen is based on water electrolysis powered by electricity generated from renewable energy sources (such as solar energy [103,104,105] and wind energy [106,107,108,109]), and thus it does not involve the release of greenhouse gases (GHG) [110,111,112], making it an attractive way to mitigate GHG emissions without resorting to carbon capture technologies as needed in the case of blue hydrogen [113,114,115,116] that is originating from fossil fuels.
In an earlier study, we proposed a novel use for hydrogen (either green hydrogen or blue hydrogen) in open-cycle magnetohydrodynamic power generators [117], in which direct power extraction (DPE) is realized with the aid of the Lorentz force. In this concept, hot gases formed as combustion products of hydrogen are seeded with a small amount of an alkali metal (either cesium “Cs” or potassium “K”) [118,119,120], which allows for an appreciable level of thermal ionization leading to a suitable electric conductivity of the weakly ionized plasma gas, which is accelerated through a linear channel [121,122,123,124]. The high-speed plasma loses part of its energy in reaction to an applied magnetic field, where an induced electric field allows collection of an induced electric current to power an external electric load [125,126]. This concept of hydrogen-powered magnetohydrodynamic power generation, which we refer to here as H2MHD, can be combined with a classical power cycle [127,128,129] for better utilization of the input heat released from the zero-emission hydrogen combustion through a dual-cycle power plant, comprising a higher-temperature open-cycle MHD generator and either a lower-temperature closed-cycle Rankine system or a lower-temperature open-cycle Brayton system [130,131,132]. A triple combined cycle that incorporates the open-cycle MHD generator as a topping layer (high temperature) with an intermediate layer having an open-cycle Brayton gas turbine (intermediate temperature) and then a layer of a closed-cycle steam turbine (low temperature) is also possible, to recover part of the rejected heat containing in the still-hot plasma gas leaving the MHD generator [133,134,135]. Introducing an even lower-temperature refrigerant-based organic Rankine cycle [136,137,138] as an additional terminal downstream level (very low temperature) was proposed, yielding a quadruple combined cycle [139].
Figure 1 depicts how the proposed hydrogen magnetohydrodynamic (H2MHD) topping cycle operates. The plasma velocity vector is designated by the symbol ( u ), the applied magnetic field vector (applied magnetic-field flux density vector) is designated by the symbol ( B ), and the electric current-density vector is designated by the symbol ( J ). Hydrogen is combusted with an oxidizer, which is either air (in the case of the conventional air-combustion mode) or oxygen (in the case of the intensified oxy-combustion mode). Oxy-combustion (the combustion of HHO or Brown’s gas) offers the advantage of elevating the combustion temperature due to the absence of diluent nitrogen [140,141,142]. A small amount of solid alkali salt, such as potassium carbonate (K2CO3) [143] is added in the form of an aqueous solution that is introduced as a fine spray by atomization, to serve as the source of the vapor alkali metal after combustion [144,145]. Either potassium element or cesium element is readily vaporized well below the combustion temperature. Potassium vaporizes at 760 °C only [146], while cesium vaporizes at a lower boiling point of 671 °C [147]. The hot-seeded hydrogen-combustion products are accelerated to supersonic speeds using a de Laval nozzle (convergent–divergent nozzle or CD nozzle) [148,149,150]. The term “channel” or “linear channel” here refers to the portion of the divergent section of the convergent–divergent nozzle where an external magnetic field is applied perpendicular to the plasma’s direction of bulk motion, and an induced electric emf (electromotive force) is produced; thus, an electric current-density can be collected through electrodes (positive cathode and negative anode) to be delivered to an external load. Like solar photovoltaic (PV) modules [151,152] and thermoelectric generators (TEG) [153,154], MHD generators operate on the principle of direct power conversion (from input energy to electricity in a single step), without an intermediate turbogenerator [155,156], and without reciprocating or rotary parts [157,158]. This is a remarkable feature in MHD generators.
The power density (or volumetric power density) is a performance metric for power generation units, describing the power output per unit volume occupied. This metric quantifies the effective utilization of space. For MHD generators, the intensive high temperature and high rate of power extraction from the plasma gives this technology a unique favorable characteristic when compared to other power technologies. For example, the Sakhalin MHD pulsed power generator experimentally demonstrated a temporary power density of about 90.7 MW/m3 (computed as an electric power output of 510 MW divided by an MHD linear channel volume of 5.625 m3) [159,160,161]. Another shock-tunnel MHD generator experimentally demonstrated a high power density of 140 MW/m3 [162]. These power densities are much higher than typical values of automotive engines (around 15 MW/m3, the volume here refers to the displacement volume [163,164]), permanent magnet (PM) synchronous motors (SM) (around 7.5 MW/m3 [165]), and triboelectric nanogenerator (TENG) energy-harvesting units (up to about 5 MW/m3 [166]), and microbial fuel cells (MFCs) (between 1 × 10–6 MW/m3 and 5 × 10–5 MW/m3 [167,168,169]). Numerical two-dimensional modeling of a hypothetical MHD channel with methane (CH4) combustion suggested a lower power density of 17.9 MW/m3 (computed through dividing an electric power of 31.4 MW by a volume of 1.75 m3) [170]. This is not remarkably high, but it is still encouraging.
There is another metric for expressing the effective utilization of a power system, which is the specific power of gravimetric power density (W/kg). However, this is more challenging to estimate compared to the volumetric power density; estimating the specific power requires more detailed and precise knowledge about the particular power system and its components, including the minor auxiliary ones if they largely affect the weight [171,172]. The specific power (or the power-to-weight ratio [173,174]) is of particular importance in aeronautical and space propulsion systems [175,176,177], which are outside the scope of this study that is concerned with terrestrial power systems. Thus, the volumetric power density is adopted here. The specific energy or the gravimetric energy density metric (Wh/kg) [178,179], and the energy density or volumetric energy density metric (Wh/L) [180,181] pertain more to energy storage units (such as batteries), and thus are not used in our study.

1.2. Goal of the Study

Because the proposed H2MHD concept is a new technology for utilizing hydrogen, there is a need for extensive investigations regarding the technological feasibility of this technology. This study aims to partly fill this huge knowledge gap through estimating the volumetric power density of an ideal H2MHD channel, at a representative value for the applied magnetic field (the applied magnetic-field flux density), a reasonable supersonic Mach number (the ratio of the plasma speed to the speed of sound), and a suitable temperature representative of hydrogen combustion.
The performance of the volumetric power density is estimated for a wide range of operational conditions, with different total pressures, different levels of seeding, two types of the seed metal, and two types of the oxidizer.
The results of this study provide several insights into the use of OCMHD in general, rather than H2MHD solely because some of the generic findings reported here can actually be applicable to other OCMHD systems with fuels other than hydrogen, including coal and fuels derived from renewable biomass [182,183,184,185]. Thus, this study is considered useful for MHD power generators in general, and for hydrogen use in these power generators in particular.

2. Research Method

2.1. Assumptions

The current study is centered on zero-dimensional computational modeling (point modeling or system-level modeling) of a hypothetical supersonic hydrogen magnetohydrodynamic channel. Therefore, no exact geometry is defined; instead, the analysis assumes uniform local properties everywhere. Despite losing boundary effects and temporal transience [186,187,188,189,190,191] that can be captured using computational fluid dynamics (CFD) methods [192,193,194,195,196], our simplifying assumption is important to allow having a closed-form expression for the power density, which is conveniently applied to estimate theoretical upper bounds. In addition, our assumption is useful in generalizing the outcomes of this study, making them broadly valid, rather than being limited to a specific H2MHD channel.
In addition, due to the large number of variables involved in the H2MHD problem, we need to fix some variables at reasonable values while changing others to explore their impact. Therefore, three quantities are fixed in the current study, and these fixed quantities are (1) the temperature, (2) the applied magnetic-field flux density, and (3) the Mach number. These fixed parameters are set to the values listed in Table 2. The significance of these fixed parameters lies also in the fact they are the same values assumed in our previous work mentioned earlier in which we proposed the H2MHD concept (but the term “M2HHD” is coined here for the first time).

2.2. Mathematical Equations

For a plasma speed ( u ), an applied magnetic-field flux density magnitude ( B ), and an electric conductivity ( σ ), the theoretical upper limit for the volumetric electric power density ( P V ) that can be extracted from the moving plasma gas is [207].
P V = 1 4 σ   u 2 B 2
This power density is restricted to ideal conditions, where the external load is matched (optimized) such that its electric resistance is equal to the equivalent internal electric resistance of the MHD generator [208,209,210,211].
The plasma speed ( u ) can be expressed in terms of the speed of sound ( a ) and the Mach number ( M ) as [212].
u = M   a
Using Equation (2) in Equation (1) allows for expressing the power density ( P V ) in a different form as
P V = 0.25   σ   M 2   a 2   B 2    
The plasma electric conductivity ( σ ) in our study is computed according to an algorithm we described and verified in detail in a previous study [213], which is based on the robust Frost’s two-conductivity approximation that takes into account the scattering of electrons by neutrals as well as the scattering of electrons by charged particles (ions and other electrons) [214], and thus interpolates the electric conductivity for the partially ionized plasma in OCMHD channels between the two extreme values for which an exact solution is known (namely, the extreme case of very weakly ionized plasma, and the extreme case of fully ionized plasma) [215,216,217]. In our algorithm, the plasma’s electric conductivity depends on the total static pressure ( p ) (the term “total” here means the sum of partial pressures of involved gaseous species, not the stagnation zero-speed pressure, and not the gauge pressure [218]), the absolute temperature ( T ), and the pre-ionization mole fractions ( X i , where i is an index referring to each involved gaseous species in the gaseous mixture formed after adding the seed alkali metal vapor to the hydrogen combustion products). Since the temperature is fixed in the current study at 2300 K, practically, the electric conductivity depends only on the total static pressure ( p ) and the mole fractions ( X i ). Therefore, we can symbolically represent the electric conductivity ( σ ) as ( σ ( p , X i ) ) to emphasize its functional dependence on these plasma properties.
The speed of sound in the pre-ionization plasma ( a ) is modeled here using the ideal gas approximation, making it a function of the absolute temperature ( T ) and the mole fractions ( X i ) only [219,220,221]. This speed of sound can be expressed as
a = C p , m i x C p , m i x R m i x   R m i x   T
where ( C p , m i x ) is the specific heat capacity at constant pressure of the gas mixture after seeding but before ionization, and ( R m i x ) is the specific gas constant for this gas mixture. The pre-ionization mixture’s specific heat capacity ( C p , m i x ) is computed from the specific heat capacities of the individual gaseous species ( C p , i for the species with the index i ) that make up the pre-ionization gas mixture, and these individual specific heat capacities are described according to the updated nine-coefficient edition of NASA (United States National Aeronautics and Space Administration) nonlinear fitting functions for computing thermodynamic properties [222,223,224,225]. Therefore, we can analytically express the speed of sound ( a ) as ( a ( T , X i ) ) to emphasize its functional dependence on the plasma properties ( T ) and ( X i ).
The nondimensional quantity ( C p , m i x / ( C p , m i x R m i x ) ) in Equation (4) is called the specific heat ratio or the adiabatic index, and here it is assigned the symbol ( γ m i x ). Therefore, the previous equation for the speed of sound ( a ) can be rewritten as [226]
a = γ m i x   R m i x   T
The gas constant ( R i ) for each gaseous species having the index i is obtained using
R i = R ¯ M i
where ( R ¯ ) is the universal (molar) gas constant, and ( M i ) is the molecular weight of the species with an index ( i ). In the current study, we have [227,228,229]
R ¯ = 8314.462618 J k m o l . K
The molecular weights of the gaseous species that appear in the current study are listed in Table 3. These were taken values from the NIST (United States National Institute of Standards and Technology) individual gaseous species Chemistry WebBook database [230,231,232,233].
The molecular weight of the gas mixture ( M m i x ) is computed as [238]
M m i x = i = 1 n s M i   X i
where ( n s ) is the number of gaseous species in the pre-ionization gas mixture. We have ( n s = 2 ) when the oxidizer is pure oxygen (oxy-hydrogen combustion), while ( n s = 3 ) when the oxidizer is air (air–hydrogen combustion). In the case of oxy-combustion, the pre-ionization mixture is treated here as a mixture of water vapor (H2O) and the vaporized alkali metal (either Cs or K). In the case of air–hydrogen combustion, the pre-ionization mixture is treated here as a mixture of water vapor (H2O), excess nitrogen (N2), and the vaporized alkali metal (either Cs or K).
After obtaining ( M m i x ), the specific gas constant of the pre-ionization gas mixture ( R m i x ) can be computed as
R m i x = R ¯ M m i x
Then, the mass fraction of each individual species ( Y i ) can be computed as [239]
Y i = X i   M i M m i x
Then, the specific heat capacity of the mixture is computed from the specific heat capacities of the individual species through mass-weighted averaging as
C p , m i x = i = 1 n s C p , i   Y i
The NASA nonlinear fitting function for estimating ( C p , i ) has the following general structure, with seven terms:
C p , i = R i b 1 , i   T 2 + b 2 , i   T 1 + b 3 , i + b 4 , i   T + b 5 , i   T 2 + b 6 , i   T 3 + b 7 , i   T 4
It should be noted that the fitting coefficients ( b 1 , i to b 7 , i ) in this equation correspond to the higher temperature range (from 1000 K to 6000 K); there is another set of fitting coefficients for a lower-temperature range (from 200 K to 1000 K), but these are irrelevant in our study (too cold to be of interest) (Table 4).
The combustion of hydrogen with air is approximated as a complete stoichiometric reaction of molecular hydrogen (H2) with a mixture having mole fractions of 79% for molecular nitrogen (N2) and 21% for molecular oxygen (O2) [240,241,242]. Therefore, the combustion products are idealized as having only water vapor (H2O) and excess non-reacting molecular nitrogen (N2) according to the following reaction equation [243,244]:
2 H 2 + O 2 + 3.762 N 2 2 H 2 O + 3.762 N 2
Vapor seed metal (either cesium or potassium) is then added to the hydrogen combustion products with a predetermined target mole fraction. This seeded mixture is then analyzed under thermal equilibrium to obtain the electrochemical properties of the post-ionization plasma (such as its electric conductivity and mole fractions of ions).
The combustion of hydrogen with oxygen is approximated as a complete stoichiometric reaction of molecular hydrogen (H2) with molecular oxygen (O2), resulting in pure water vapor (H2O) as the single combustion product. This is described by the following reaction equation:
2 H 2 + O 2 2 H 2 O
Vapor seed metal (either cesium or potassium) is then added to this combustion product water vapor with a predetermined target mole fraction. This seeded mixture is then analyzed under thermal equilibrium to obtain the electrochemical properties of the post-ionization plasma.
We conclude the current mathematical description of the modeling methodology by presenting a customized version of the expression governing the ideal volumetric power density in Equation (3), by setting the Mach number ( M = 2 ) and the magnitude of the applied magnetic-field flux density ( B = 5   T ). This gives the following special expression pertaining to our study:
P V W m 3 = 25   σ S m   a m s 2
To conveniently express the values of the power density ( P V ), the larger unit of (MW/m3) is adopted rather than the basic SI unit of (W/m3), thus the previous equation is modified to
P V M W m 3 = 25 10 6   σ S m   a m s 2
While the speed of sound ( a ) in the above equation should correspond to the post-ionization plasma condition (taking into account the small changes in the gaseous mixture because of the partial ionization of some alkali seed atoms), we use the pre-ionization value that corresponds to an all-neutral (no ions, no electrons) seeded gas mixture. This establishes consistency with the reported total pressure values as a controllable variable, where these pressures correspond to the pre-ionization condition. This matter is practically negligible because only a small portion of the seeded alkali metal can be thermally ionized at the fixed temperature of 2300 K, causing the mole fraction of either the alkali ions or the free electrons to be below 0.1% in all cases we cover, and even much less than this in many of the analyzed cases.

2.3. Varied Conditions

In Table 5, we list the quantities to be varied in the current study while estimating the ideal volumetric power density ( P V ) for a magnetohydrodynamic (MHD) generator channel, as well as the values assigned to these varied quantities. These quantities are listed below:
(1) The total absolute pressure ( p ) after the vaporous seed alkali metal is mixed with the hydrogen combustion gases. The unit is the standard atmosphere (1 atm = 101,325 Pa = 1.01325 bar). Nine values are assigned to this continuous variable, sampling it between a very low value of 1/16 atm (6332.812 Pa or 6.332812 kPa) to a high value of 16 atm (1.6212 MPa or 1621.2 kPa). These values allow for exploring the impact of this variable over a wide range. The nine selected pressures form a geometric sequence with a multiplicative progression constant of two.
(2) The mole fraction ( X s e e d ) of the vaporous seed alkali metal after it is mixed with the hydrogen combustion gases. This fraction is expressed as a percentage. Five values are assigned to this continuous variable, sampling it between a very low value of 0.0625% or (1/16)% to a high value of 16%. These values allow for exploring the impact of this variable over a wide range. The five selected seed mole fractions form a geometric sequence with a multiplicative progression constant of four.
(3) The type of the vaporous seed alkali metal to be added to the hydrogen combustion gases. This is a binary variable (this means it can take one of two values only). The seed type here can be either cesium (Cs) or potassium (K). These two particular chemical elements are chosen here as the possible seed metal that serves as the source of electrons due to their easy thermal ionization compared to the combustion products gases (H2O or N2). The ionization energy of cesium (Cs, atomic number 55; located in the alkali metal group IA or group 1 and period 6 in the periodic table of chemical elements) is 3.893 eV/particle (375.6 kJ/mol), and this is the lowest value among all chemical elements in the periodic table [245,246,247]. We point out here that 1 eV/particle (electronvolt per particle) is equivalent to 96.49 kJ/mol (kilojoule per mole) [248,249,250]. The ionization energy of potassium (K, atomic number 19; located in the alkali metal group IA or group 1 and period 4 in the periodic table of chemical elements) is 4.34 eV/particle (419 kJ/mol), and this is the fourth lowest value among all chemical elements in the periodic table [251,252,253]. Although francium (Fr, atomic number 87; located in the alkali metal group IA or group 1 and period 7 in the periodic table) has a lower ionization energy of (3.9 eV/particle or 376 kJ/mol) than potassium [254,255], francium is radioactive and very unstable (has no practical uses) due to its small half-life of only 22 min [256,257,258]. Although rubidium (Rb, atomic number 37; located in the alkali metal group IA or group 1 and period 5 in the periodic table) has a lower ionization energy of (4.177 eV/particle or 403.0 kJ/mol) than potassium [259,260], rubidium is a rare element, and its use is typically limited to scientific research [261,262,263]. The ionization energy of molecular nitrogen (N2) is very large, being 15.6 eV/particle (1505 kJ/mol) approximately [264,265]. The ionization energy of atomic nitrogen (N) is also large, being 14.5 eV/particle (1399 kJ/mol) approximately [266,267]. The dissociation energy of molecular nitrogen (N2) into two atoms is also much lower (9.76 eV/mole or 942 kJ/mol) [268,269], but it is still higher than the ionization energy of either potassium or cesium. The ionization energy of the water molecules (H2O) is very large, being 12.6 eV/particle (1216 kJ/mol) approximately [270]. The dissociation energy of the water molecules (H2O) into a hydrogen atom (H) and a hydroxyl radical (OH) radical is approximately 5.16 eV/particle (498 kJ/mol) [271], making this dissociation less probable than the ionization of either potassium or cesium.
(4) The type of oxidizer to be used in the combustion of hydrogen. This is a binary variable, and it can be either air or pure oxygen (O2).
Therefore, there is a total of 180 conditions (data points) to be computationally represented in the current study (because there are nine pressures, five seed mole fractions, two seed types, and two oxidizer types).

2.4. Power Density Criterion

In our study, it is sufficient to estimate the volumetric electric power density output under the various operational conditions for the H2MHD (hydrogen magnetohydrodynamic) concept, as this is its primary goal and major contribution. However, the value of our study is boosted when a criterion for this power density is also recommended, to help in judging the feasibility of the H2MHD concept and the preferred set of conditions suitable for it.
To justify such a criterion (a power density threshold), we consider that a mid-size power plant has a power capacity on the order of 600 MW [272,273,274,275]. If an H2MHD power density of 50 MW/m3 is possible, then this 600 MW power capacity can be accomplished with a compact space of only 12 m3 in the case of a properly operated H2MHD generator. This small volume can be formed, for example, as a divergent channel with a width of 1 m, a length of 6 m, and a mean height of 2 m [276,277,278,279]. To account for the assumptions made in the current study, and the losses ignored in our system-level approach, we set our threshold to twice this value, making it 100 MW/m3. This means that 100 MW/m3 in our analysis may actually be reduced to half of this value in reality. To show the gigantic amount of power delineated by this criterion, it is helpful to compare this threshold (100 MW/m3) with the typical area power density of photovoltaic (PV) solar panels under good sunshine, which is only about 200 W/m2 (0.0002 MW/m2) [280,281,282].
We emphasize that this suggested criterion of 100 MW/m3 here is not a precise value, but involves some arbitrariness. However, it is considered a reasonable rough guide for assessing the suitability of H2MHD power generation.

3. Results

This results section is divided into four subsections, which are mapped with the four possibilities arising from the two binary variables (namely, two alkali seed types and two oxidizer types). For each case of these four conditions, we present first the variation in the plasma electric conductivity ( σ ) after thermal equilibrium as a function of the total pre-ionization absolute pressure ( p ) with the mole fraction of the neutral alkali seed metal ( X s e e d ) being a parameter. Thus, the variation in ( σ ) is visualized versus ( p ) as five curves (one curve per X s e e d ), and this visualization is presented four times (for two seed types and two oxidizer types). Similarly, the estimated ideal volumetric power density ( P V ) is visualized as a function of ( p ) four times in the four coming subsections, at five values of ( X s e e d ) in each ( P V ) plot.
To aid in interpreting the electric conductivity, it is voluntarily expressed as a multiple of typical seawater electric conductivity, which is assigned here a value of 5 S/m [283,284,285]. The secondary (right) vertical axis in the electric conductivity ( σ ) plots is utilized for this purpose, while the primary (left) vertical axis is utilized for expressing the absolute (unscaled) value of ( σ ), expressed in the SI unit of S/m.
Regarding the pre-ionization speed of sound ( a ), and thus the plasma speed ( u ), because ( a ) is independent of the total pressure, it does not need to be visualized as a dependent variable of ( p ). Instead, we only tabulate its values for the five values of ( X s e e d ) at the beginning of each of the four forthcoming subsections, along with other thermochemical properties that are also independent of the total absolute pressure (such as the pre-ionization mixture’s molecular weight).
In addition to the visualized curves, we also list the numerical values for the variations in ( σ ) and ( P V ). This is useful for the sake of precision and for subsequent use of these values by interested readers in external quantitative analysis or comparisons.
For effective contrast and use of the plots, the ranges of the axes are retained the same regardless of the actual range of the data being visualized. This allows for quick identification of the relative size of the data visualized in different figures, which correspond to different operational cases or the H2MHD channel.
For effective identification of the individual influence of the seed type and the oxidizer type, we intentionally order the four forthcoming subsections such that in the transition from one subsection to the next, only one change is made. This change either alters the seed type or alters the oxidizer type (but not altering both of them at once, which makes it difficult to separate the influence of each alteration from the other).

3.1. Cesium Seed and Air Oxidizer

The first set of results corresponds to the condition of using cesium vapor as the ionizable gas, and using air as the oxidizer for the hydrogen combustion.
In Table 6, we provide our computed thermochemical characteristics for the seeded gas mixture before ionization, according to the mathematical model presented in the previous section.
It is noticed that as the mole fraction of cesium increases, the mixture’s gas constant ( R m i x ) decreases (because the mixture’s molecular weight increases). This is justified by the heavier molecular weight of cesium (Cs) compared to either water vapor (H2O) or molecular nitrogen (N2). One mole of cesium (Cs) is 4.744 times heavier than one mole of molecular nitrogen (N2), and 7.377 times heavier than one mole of water (H2O). One of the implications of the decline in the mixture’s gas constant ( R m i x ) is a decline in the speed of sound ( a ), and thus the plasma’s bulk speed (given that it is always twice the speed of sound in the current study). However, these influences of the additive cesium remain marginal up to a mole fraction of 1%. The remarkable changes at the large mole fraction of 16% are quite theoretical only, because practically it is not expected to have such a big portion of the seed.
Despite the seed content, the mole fraction ratio ( X N 2 / X H 2 O ) is always 1.881, as implied by the reaction stoichiometry indicated in Equation (13), which stipulates that this ratio should be 3.762 / 2 = 1.881 in the cases of air–hydrogen combustion.
Figure 2 shows how the electric conductivity of the hydrogen plasma drops monotonically and nonlinearly as the total pre-ionization absolute pressure increases. This trend is identical for the five seed levels of cesium vapor. This drop in the electric conductivity can be explained by the larger number density of neutral heavy particles (atoms and molecules) in the plasma at higher pressures, which retards the directional movement of the liberated electrons from the ionized cesium atoms due to more frequent collisions with these neutral particles [286,287,288].
Unlike the effect of the pressure, the effect of the cesium additive is non-monotonic. Starting from a small level of (Cs), adding more (Cs) initially improves the electric conductivity. This boost slows down and reaches a peak near a cesium mole fraction of X C s = 6 % , and then the electric conductivity starts to drop as more cesium vapor is added in place of water vapor and nitrogen. This peaking behavior of the electric conductivity ( σ ) can be attributed to the presence of two factors that impact the electric conductivity due to the free electrons produced by the ionization of a fraction of the cesium atoms. The first factor is the large number density of electrons, due to more cesium atoms; this improves the electric conductivity as more cesium is added. The other factor is the electron–ion interaction (attraction) and electron–electron interaction (repulsion) due to their electric charges. As the number of electrons (and thus the number of cesium ions) increases beyond an appreciable level, the attenuating effect on the electric conductivity caused by these interactions (Coulomb scattering) [289,290,291] overpowers the accentuating effect on the electric conductivity caused by the electrons’ number density, and thus the electric conductivity declines after reaching a peak value. This explains why the electric conductivity with a higher cesium mole fraction of 16% is less than the electric conductivity with a lower cesium mole fraction of 4%. From the figure shown, it may be inferred qualitatively that a pre-ionization cesium mole fraction of 1% or 2% is preferred, because the weak increase in the electric conductivity beyond these levels does not seem to justify such additions of cesium content.
Table 7 lists the numerical values of the plasma electric conductivity as visualized in the previous figure. We also add to the table the electric conductivity values at a pre-ionization cesium mole fraction of X C s = 6 % , which approximately corresponds to the maximized electric conductivity for this case (air–hydrogen combustion plasma with cesium seed).
Considering the five selected seed levels (0.0625%, 0.25%, 1%, 4%, and 16%), the highest obtained electric conductivity is σ = 55.0610   S / m (at X C s = 4 % and p = 0.0625   a t m ), and this is about 11 times the electric conductivity of seawater. At the higher X C s = 16 % (and the same total pressure of 0.0625 atm), the electric conductivity is slightly lower, with a value of 51.7766 S/m; at the lower X C s = 1 % (and the same total pressure of 0.0625 atm), the electric conductivity is mildly lower, with a value of 42.2099 S/m. If we take a total pressure of 1 atm (atmospheric MHD channel) as a reference, one can see that the electric conductivity can exceed 10 S/m (twice the conductivity of seawater) with only 1% cesium level.
Using the presented values of the speed of sound (see Table 6) and the values of the electric conductivity (see Table 7) in Equation (16) gives the corresponding ideal volumetric power densities ( P V ) for the current H2MHD case of cesium seeding and air oxidizer. The ( P V ) results are visualized in Figure 3 and listed in Table 8. We also add to the table the power density values at a pre-ionization cesium mole fraction of X C s = 3 % , which approximately corresponds to the maximized power density for this case (air–hydrogen combustion plasma with cesium seed).
Like the electric conductivity, the power density declines monotonically and nonlinearly as the total pressure increases. Also, like the electric conductivity, the power density exhibits a peak when the cesium level increases, and then it starts to decrease. However, unlike the case of the electric conductivity ( σ ), the cesium level for maximizing the power density ( P V ) has shifted down from being close to 6% to being close to 3%. This shift is explained by the simultaneous decline in the speed of sound as the cesium level increases. This added effect of the speed of sound (which quadratically affects the power density) favors lower cesium values for maximizing the power density.
The tabulated data are helpful in quantifying potential ranges of the power density in the current case of cesium seed and air oxidizer. Considering the five selected seed levels (0.0625%, 0.25%, 1%, 4%, and 16%), the best obtained power density is 1139.332 MW/m3 (at 4% cesium and 0.0625 atm), and the worst obtained power density is 23.854 MW/m3 (at 0.0625% cesium and 16 atm). At atmospheric pressure, a promising power density of 362.806 MW/m3 is attainable with 4% cesium, and an attractive power density of 298.752 MW/m3 is attainable with only 1% cesium. Even with 0.25% cesium, a power density of 177.106 MW/m3 can be achieved at atmospheric combustion. These are favorably intense power densities, which we found to require a minute amount of cesium seed (at 1 atm) that is not more than 0.08% Cs (at which P V = 104.932   M W / m 3 ), and all of them well exceed our voluntary threshold criterion of 100 MW/m3.

3.2. Potassium Seed and Air Oxidizer

The second set of results corresponds to the condition of using potassium vapor as the ionizable gas and using air as the oxidizer for the hydrogen combustion. Thus, compared to the previous set of results in the preceding subsection, the only change made here is changing the seed type from cesium to potassium.
Before presenting the results of this change, it is useful to mention that due to the higher ionization energy of potassium compared to cesium, potassium is harder to lose its valence electron in its outermost orbital (4s1) [292,293,294] upon heating compared to cesium, whose outermost valence electron is in (6s1) [295,296,297]. Therefore, it is expected that this change is accompanied by a drop in the H2MHD performance, primarily because of a decline in the electric conductivity. However, due to the nonlinearity of the problem and the present influence from other factors (particularly the speed of sound), the exact effect of this change in the seed type is not easy to predict analytically. Thus, our computational modeling becomes valuable in understanding quantitatively and qualitatively the implications of this single change.
In Table 9, we provide our computed thermochemical characteristics for the seeded gas mixture before ionization. Similar to the case of cesium seeding, the mixture’s gas constant ( R m i x ) decreases (and the mixture’s molecular weight increases) as the level of potassium increases. However, this decrease in the case of potassium additive is much weaker than in the case of cesium additive due to the weaker difference in the molecular weight between potassium (K) and either molecular nitrogen (N2) or water vapor (H2O). One mole of potassium (K) is only 1.396 times heavier than one mole of molecular nitrogen (N2) and 2.170 times heavier than one mole of water (H2O).
For reference, the molecular weight of the combustion products without any seeding (34.710170% H2O and 65.289830% N2 by mole) is 24.54304 kg/kmol; the unseeded mixture’s gas constant is 338.7707 J/(kg.K); the unseeded mixture’s specific heat ratio is 1.243900; and the unseeded mixture’s speed of sound (at 2300 K) is 984.486 m/s.
Figure 4 shows how the electric conductivity of the hydrogen plasma monotonically and nonlinearly drops as the total pre-ionization absolute pressure increases. This is qualitatively similar to what was observed when cesium was the additive seed. Also, the existence of a peak electric conductivity at a certain level of potassium seed is similar to what was observed in the case of cesium. This potassium level that maximizes the electric conductivity is again (as was the case in the cesium case) approximately near 6%. It is noticeable that the electric conductivity (or any given level of seeding) dropped strongly when the seed is changed from cesium to potassium. For example, the maximum obtained electric conductivity in the case of cesium was 55.0610 S/m (at X C s = 4 % and 0.0625 atm) is 2.55 times the maximum obtained electric conductivity in the case of potassium, which is 21.6201 S/m (at X K = 4 % and 0.0625 atm).
Table 10 lists the numerical values of the plasma electric conductivity as visualized in the previous figure. We also add to the table the electric conductivity values at a pre-ionization cesium mole fraction of X K = 6 % , which approximately corresponds to the maximized electric conductivity for this case (air–hydrogen combustion plasma with potassium seed).
Considering the five selected seed levels (0.0625%, 0.25%, 1%, 4%, and 16%), the highest obtained electric conductivity is σ = 21.6201   S / m (at X K = 4 % and p = 0.0625   a t m ). At the higher X K = 16 % (and the same total pressure of 0.0625 atm), the electric conductivity is slightly lower, with a value of 19.7671 S/m; at the lower X C s = 1 % (and the same total pressure of 0.0625 atm), the electric conductivity is mildly lower, with a value of 16.0343 S/m. If we take a total pressure of 1 atm (atmospheric MHD channel) as a reference, one can see that the electric conductivity is restricted to 6 S/m.
The ideal volumetric power densities ( P V ) for the current H2MHD case of potassium seeding and air oxidizer are visualized in Figure 5 and listed in Table 11. We also add to the table the power density values at a pre-ionization cesium mole fraction of X K = 5 % , which approximately corresponds to the maximized power density for this case (air–hydrogen combustion plasma with potassium seed).
Unlike the case of cesium seeding, the peak power density with potassium seeding occurs near a seed level of 5% (not 3%), which is close to the potassium seeding level that maximizes the electric conductivity. This similarity of the potassium seed level ( X K ) that maximizes the electric conductivity and the one that maximizes the power density is related to the near-neutral influence of the seed level on the speed of sound for the case of potassium seed (much weaker influence than the case of cesium seed). A nonlinear monotonic decline in the power density ( P V ) as the total pressure ( p ) increases is noted for any potassium level, and this behavior resembles the one found earlier for the case of cesium seeding.
The power density with potassium seeding is clearly smaller than the one with cesium seeding (when other settings are kept the same). For example; considering the five selected seed levels (0.0625%, 0.25%, 1%, 4%, and 16%), the largest obtained power density at 1 atm in the case of potassium seeding is 141.808 MW/m3 (with X K = 4 % ). This is 2.56 times the power density at 1 atm and X C s = 4 % , which was 362.806 MW/m3. Although potassium seeding is associated with higher speeds of sound compared to cesium seeding, this small advantage for potassium seeding is largely counteracted by the big penalty in the electric conductivity due to changing the seed vapor from cesium to potassium.
Despite the drop in the power density in the case of potassium seeding, the power density can still exceed 100 MW/m3 at 1 atm, with potassium levels as small as 1%. Thus, the H2MHD concept is still viable.

3.3. Potassium Seed and Oxygen Oxidizer

The third set of results corresponds to the condition of using potassium vapor as the ionizable gas and using pure oxygen as the oxidizer for the hydrogen combustion. Thus, compared to the previous set of results in the preceding subsection, the only change made here is changing the oxidizer type from air (oxygen–nitrogen mixture) to pure oxygen.
Before presenting the results of this change, it is useful to mention that due to the absence of nitrogen from the combustion products, hydrogen combustion results in pure water vapor (H2O). Water vapor has a relatively low molecular weight and thus has a relatively high specific gas constant. Therefore, the speed of sound in the current case of oxy-combustion of hydrogen is higher than the speed of sound in the previous case of air-combustion of hydrogen (for the same level of potassium seed). Therefore, it is expected that the change in the oxidizer (from air to oxygen) improves the speed of sound, and thus the plasma speed (given its fixed relation to the speed of sound in our study). It is thus the change in the electric conductivity that determines how eventually the power density is affected by the change in the oxidizer (from air to oxygen). This is difficult to predict without the numerical simulations performed in our study. This emphasizes the value and contribution made by our study to the field of MHD power generation through quantitatively and qualitatively investigating the influence of the oxidizer on the MHD power generation process.
In Table 12, we provide our computed thermochemical characteristics for the seeded gas mixture before ionization. As mentioned earlier, the specific gas constant and the speed of sound in the current case of oxy-hydrogen combustion are higher than their values in the previous case of air–hydrogen combustion. For example, with 1% potassium seed, the speed of sound with oxy-hydrogen combustion (1114.801 m/s) is 1.135 times the speed of sound with air–hydrogen combustion (982.178 m/s) with the same level of potassium seeding. Due to the larger molecular weight of potassium compared to water vapor, the more the added potassium, the higher the mixture’s molecular weight and thus the lower the mixture’s gas constant and the lower the speed of sound in that mixture of water vapor and potassium vapor. The mixture’s specific gas constant ( γ m i x ) increases as the potassium level increases, but the decrease in the mixture’s gas constant ( R m i x ) is stronger; thus, ultimately the speed of sound decreases when the potassium level increases.
Figure 6 shows the profiles of the electric conductivity of the hydrogen plasma. The monotonic and nonlinear drop as the total pre-ionization absolute pressure increases is retained, similar to the previous case of air–hydrogen combustion. However, there are two remarkable changes that can be observed from the figure when compared to its counterpart in the previous subsection that describes air–hydrogen combustion (with the same potassium seeding).
The first change worthy of attention is the overall suppression in the electric conductivity. This can be attributed to the larger effective electron-neutral momentum transfer collision cross-section for water vapor compared to molecular nitrogen [298,299,300,301]. Thus, when the molecular nitrogen in the combustion products is replaced with water vapor, electron mobility is retarded, and this suppresses the electric conductivity of the plasma.
The second change worthy of attention is that now with oxy-hydrogen combustion, the electric conductivity with 16% potassium is larger than the electric conductivity with 4% potassium. In the previous case of air–hydrogen combustion, the electric conductivity with 16% potassium was smaller than the electric conductivity with 4% potassium, due to the detrimental effect of over-seeding. In fact, while the electric conductivity with 16% potassium is larger than the electric conductivity with 4% potassium in the current case of oxy-hydrogen combustion, the peak electric conductivity has been already passed at X K = 16 % . This means that this peaking phenomenon of the electric conductivity is actually still present in the current case of oxy-hydrogen plasma. The main difference is that this peak was easy to notice in the previous electric conductivity figure corresponding to air–hydrogen plasma with potassium seeding (in the previous subsection) because the potassium mole fraction ( X K ) that maximized the electric conductivity was closer to the intermediate value of 4% than the terminal value of 16%. Changing the oxidizer from air to oxygen causes this particular ( X K ) of maximized electric conductivity ( σ ) to shift to a larger position, becoming closer to the terminal value of 16% (but below it, located nearly at X K = 13 % ). Despite the gain in the electric conductivity when the potassium seed mole fraction ( X K ) is increased from 4% to 16% (or 13%), the gain is too small to justify the need for exceeding the seed level beyond 4%.
Table 13 lists the numerical values of the plasma electric conductivity as visualized in the previous figure. We also add to the table the electric conductivity values at a pre-ionization potassium mole fraction of X K = 13 % , which approximately corresponds to the maximized electric conductivity for this case (oxy-hydrogen combustion plasma with potassium seed).
Considering the five selected seed levels (0.0625%, 0.25%, 1%, 4%, and 16%), the highest obtained electric conductivity is σ = 16.1854   S / m (at X K = 16 % and p = 0.0625   a t m ). At the lower X K = 4 % (and the same total pressure of 0.0625 atm), the electric conductivity is mildly lower, with a value of 13.9154 S/m. If we take a total pressure of 1 atm as a reference, one can see that the electric conductivity is below 5 S/m (the approximate electric conductivity of seawater) regardless of the seeding level.
The ideal volumetric power densities ( P V ) for the current H2MHD case of potassium seeding and oxygen oxidizer are visualized in Figure 7 and listed in Table 14. We also add to the table the power density values at a pre-ionization cesium mole fraction of X K = 9 % , which approximately corresponds to the maximized power density for this case (oxy-hydrogen combustion plasma with potassium seed).
The electric conductivity ( σ ) at a potassium seed level ( X K = 16 % ) was found to be larger than the electric conductivity at ( X K = 4 % ). Similarly, the volumetric power density ( P V ) at a potassium seed level ( X K = 16 % ) is larger than the power density at ( X K = 4 % ). However; because the speed of sound (and thus the plasma speed) at ( X K = 16 % ) is larger than its value at ( X K = 4 % ), the gap in ( P V ) is diminished when the seed levels of 4% and 16% are compared.
When comparing the range of ( P V ) here with those in the previous subsection (air-combustion instead of oxy-combustion); we notice that although the power density dropped, this drop is not large. For example, at 1 atm and 4% potassium, the power density here with oxy-combustion is 111.573 MW/m3, while it was 141.808 MW/m3 in the previous case of air-combustion. The difference (30.2350 MW/m3) is 23.9% of the average of both values (126.6905 MW/m3).
For exceeding the threshold power density of 100 MW/m3, the potassium seed level should not be below 3% (at which we found that a = 1103.376   m / s and σ = 3.4118   S / m ; thus, P V = 103.841   M W / m 3 )

3.4. Cesium Seed and Oxygen Oxidizer

The fourth (and final) set of results corresponds to the condition of using cesium vapor as the ionizable gas and using pure oxygen as the oxidizer for the hydrogen combustion. Thus, compared to the previous set of results in the preceding subsection, the only change made here is changing the seed type from potassium to cesium.
In Table 15, we provide our computed thermochemical characteristics for the seeded gas mixture before ionization. The speed of sound drops faster with the cesium fraction compared to the previous case of potassium seeding, due to the large influence of cesium on the mixture’s gas constant compared to potassium; this is because the cesium atom is 3.399 times heavier than the potassium atom. For example, at the same mole fraction of 4%, the speed of sound in the case of cesium seeding is 1002.552 m/s, which was larger (1097.826 m/s) in the case of potassium seeding. For comparison purposes, the speed of sound in pure water vapor (at 2300 K) is 1120.682 m/s. Thus, introducing a fraction of 4% cesium vapor causes a relative drop in this original speed of sound (1120.682 m/s) by 118.130 m/s or 10.54%. On the other hand, introducing a fraction of 4% potassium vapor causes a relative drop in the original speed of sound (1120.682 m/s) by only 22.856 m/s or 2.04%.
In addition, it may be useful to add that for unseeded water vapor (H2O), the specific gas constant is 461.5223 J/(kg.K) and the specific heat ratio is 1.183162.
Figure 8 shows how the electric conductivity of the hydrogen plasma varies with the total absolute pressure at the five selected cesium levels. The variations show a monotonical nonlinear drop, as observed in all previous cases (the three preceding subsections). However, compared to the previous subsection particularly, the electric conductivity is largely boosted here after replacing the seed from potassium (the previous subsection) to cesium (the current subsection). For example, at 1 atm and X C s = 4 % , the electric conductivity ( σ ) here is 11.0877 S/m. But in the previous case, at 1 atm and X K = 4 % , this electric conductivity was 3.7030 S/m. The ratio of the two values of ( σ ) is 2.99.
Also, as in the previous set of results (oxy-hydrogen combustion with potassium seeding), the seeding level at which the electric conductivity is maximized is larger than that level in the case of air–hydrogen combustion. That seeding level of maximum ( σ ) is approximately 13% in the current case of oxy-hydrogen combustion with cesium seeding (while it was near 6% in the case of air–hydrogen combustion with cesium seeding).
Table 16 lists the numerical values of the plasma electric conductivity as visualized in the previous figure. We also add to the table the electric conductivity values at a pre-ionization cesium mole fraction of X C s = 13 % , which approximately corresponds to the maximized electric conductivity for this case (oxy-hydrogen combustion plasma with cesium seed).
Considering the five selected seed levels (0.0625%, 0.25%, 1%, 4%, and 16%), the highest obtained electric conductivity is σ = 43.5831   S / m (at X C s = 16 % and p = 0.0625   a t m ). At the lower X C s = 4 % (and the same total pressure of 0.0625 atm), the electric conductivity is mildly lower, with a value of 37.7620 S/m (thus, 13.36% of the σ value of 43.5831 S/m at 16% cesium is lost). If we take a total pressure of 1 atm as a reference, an electric conductivity of 10 S/m is achievable with a cesium mole fraction of 3% (at which the electric conductivity is 10.2333 S/m).
The ideal volumetric power densities ( P V ) for the current H2MHD case of cesium seeding and oxygen oxidizer are visualized in Figure 9 and listed in Table 17. We also add to the table the power density values at a pre-ionization cesium mole fraction of X C s = 5 % , which approximately corresponds to the maximized power density for this case (oxy-hydrogen combustion plasma with cesium seed).
Like the case of cesium seeding with air–hydrogen combustion, but unlike the previous case of potassium seeding and oxy-hydrogen combustion, the power density here with 4% Cs is larger than the power density with 16% seed. This is a combined effect of how the electric conductivity ( σ ) and the speed of sound ( a ) respond to changes in the cesium mole fraction ( X C s ), which is different from their responses to changes in the potassium mole fraction ( X K ).
The power density with cesium seeding (the current subsection) is clearly larger than the one with potassium seeding (the previous subsection).
A power density of 100 MW/m3 or more is easily attainable at 1 atm, requiring a small mole fraction of cesium seed such as 0.25% only. Even with X C s = 0.20 % , we obtain P V = 101.838   M W / m 3 .

4. Discussion

In the current section, we discuss three topics that are related to the performed study but with a lower level of importance than the core results presented in the previous section. Therefore, skipping this discussion does not cause a deficiency in understanding the scope and outcomes of our study.
The first topic we discuss is a clarification of one of the study’s findings that can be easily misinterpreted by the reader. This results-backed finding is that using air as an oxidizer is preferred to using oxygen in terms of the plasma electric conductivity and the output power. However, it should be noted that this preference for air as an oxidizer does not mean preferring an air-combustion process over an oxy-combustion process. The reason is that we assumed that the temperature remains the same when the oxidizer is changed from air to oxygen. However, oxy-combustion leads practically to elevated temperatures [302,303,304], which strongly boosts the electric plasma conductivity and thus the electric power output. Therefore, favoring air as an oxidizer is an outcome we found based on its influence on the chemical composition of the combustion products only, while factoring out its influence on the temperature of these combustion products. When the influence of the temperature due to oxy-fuel combustion is factored in, then the use of pure oxygen becomes preferred, because the large gain in the electric conductivity due to the increase in temperature is more important than the small decline in the electric conductivity due to the altered chemical composition of the plasma.
The second topic we discuss is the contribution made by our current study to the field of hydrogen magnetohydrodynamic power generation, in comparison to other studies in the literature. Although magnetohydrodynamic power generation is not a novel topic, our contribution is expressed through proposing the use of pure hydrogen as an environment-friendly fuel and quantifying the expected electric power under this proposed configuration over a wide range of settings. Due to the high nonlinearity of the problem’s physics and the complicated relationships between various design variables and operating parameters [305,306,307,308], our study is viewed as a valuable support to the hydrogen magnetohydrodynamic power generation concept, where novel results are presented in detail as a preliminary design stage. In Table 18, we summarize the difference between our study and three recent studies (all dated 2025) in the same broad field of magnetohydrodynamic power generation. This emphasizes the distinction between our work and the work of others, and demonstrates the contributions claimed by our study.
The third topic we discuss concerns the success of magnetohydrodynamic power generation compared to some other power technologies. Although magnetohydrodynamic power generation offers the advantage of no moving parts, direct power conversion, and high power density, this technology is not deployed commercially although the concept was proven experimentally [312,313,314,315,316,317,318,319,320]. Other power technologies (particularly solar photovoltaic modules and wind turbines) are successfully deployed as clean alternatives to traditional fossil-fuel power plants. The reasons for the delay in commercializing magnetohydrodynamic power generation can be attributed to the need for specialized powerful magnets, the high capital cost of construction, and the challenge in maintaining high electric conductivity [321,322,323,324,325,326,327,328]. Despite these challenges, research in magnetohydrodynamic power is still active [329,330,331,332,333,334,335], and new ideas may make the concept more competitive and attractive for realization.

5. Conclusions

In the current study, we reported novel results about our proposed concept of hydrogen magnetohydrodynamic power generation, particularly results about potential volumetric electric power densities under different operational conditions. We forecast the performance of a Mach 2 supersonic channel, operating with hydrogen-combustion plasma at a uniform temperature of 2300 K. The influence of the pressure, seed level, seed type, and oxidizer type was presented graphically and numerically.
The following findings can be stated:
Hydrogen magnetohydrodynamic generators permit concentrated power generation, with theoretical power densities as high as 300 MW/m3 under a controlled condition of 1 atm and 1% cesium seeding (mole fraction), while using air as a conventional oxidizer.
Using cesium vapor as a seeding alkali metal for producing free electrons (as the main charge carriers in the plasma) is highly advantageous compared to potassium. The gain can more than double the output power.
Increasing the total static pressure monotonically decreases the electric conductivity and thus decreases the power density and performance of the H2MHD generator.
Increasing the alkali metal vapor seed amount monotonically increases the mixture’s molecular weight, and thus monotonically decreases the mixture’s specific gas density and in turn monotonically decreases the speed of sound in the mixture.
Increasing the alkali metal vapor seed amount increases the electric conductivity up to a certain level, where a peak electric conductivity is reached and then it declines with further seeding.
With air–hydrogen combustion and cesium seeding, the electric conductivity is maximized at a pre-ionization seeding mole fraction near 6%, while the power density is maximized (with a fixed Mach number) at a lower mole fraction near 3%.
With air–hydrogen combustion and potassium seeding, the electric conductivity is maximized at a pre-ionization seeding mole fraction near 6%, while the power density is maximized (with a fixed Mach number) at a lower mole fraction near 5%.
With oxy-hydrogen combustion and potassium seeding, the electric conductivity is maximized at a pre-ionization seeding mole fraction near 13%, while the power density is maximized (with a fixed Mach number) at a lower mole fraction near 9%.
With oxy-hydrogen combustion and cesium seeding, the electric conductivity is maximized at a pre-ionization seeding mole fraction near 13%, while the power density is maximized (with a fixed Mach number) at a lower mole fraction near 5%.
The future research direction of the current study may include (1) experimental investigation or a pilot-scale hydrogen magnetohydrodynamic generator (which was not possible for us to perform); (2) modeling the influence of plasma temperature (which was kept constant here); (3) modeling the influence of flue-gas recycle (FGR), where some of the hot steam leaving the generator is sent back to the generator and is mixed with fresh combustion products in order to exploit the high energy still contained within that exhaust steam; and (4) modeling the influence of replacing pure hydrogen as a fuel with either methane (CH4) and a blend of hydrogen and methane. Because the Mach number and the applied magnetic-field flux were shown to have a simple quadratic correlation with the power density, they do not warrant a new separate study. However, the influence of the temperature is more complicated. It not only alters the electric conductivity in a nonlinear fashion, but also alters some of the thermodynamic properties.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/hydrogen6020031/s1.

Funding

This research received no external funding.

Data Availability Statement

Supplementary Data files related to the data reported in this research article are provided as four computer files, corresponding to the four main cases presented here (air–hydrogen combustion with cesium seeding, air–hydrogen combustion with potassium seeding, oxy-hydrogen combustion with potassium seeding, and oxy-hydrogen combustion with cesium seeding). Respectively, these data files are: airCs.csv, airK.csv, oxyK.csv, and oxyCs.csv.

Conflicts of Interest

The author declares that there are no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Field, D. The most popular element in the universe. Phys. World 1995, 8, 24. [Google Scholar] [CrossRef]
  2. Goncharuk, V.V.; Goncharuk, D.K. Water is Everywhere. It Holds Everything, Even a Key to Understanding the Universe. D. I. Mendeleev’s Law is the Prototype of the Universe Constitution. J. Water Chem. Technol. 2019, 41, 341–346. [Google Scholar] [CrossRef]
  3. Barayeu, U.; Sawa, T.; Nishida, M.; Wei, F.-Y.; Motohashi, H.; Akaike, T. Supersulfide biology and translational medicine for disease control. Br. J. Pharmacol. 2025. [Google Scholar] [CrossRef]
  4. Krebs, R.E. The History and Use of Our Earth’s Chemical Elements: A Reference Guide; Greenwood Press: Westport, CT, USA, 2006; Available online: https://books.google.com.om/books?id=D7LOEAAAQBAJ (accessed on 4 April 2025).
  5. Ma, Y.H. Palladium Membranes for Hydrogen Separation. In Membranes for Energy Conversion; Peinemann, K.-V., Nunes, S.P., Eds.; John Wiley & Sons: Weinheim, Germany, 2008; pp. 245–261. Available online: https://books.google.com.om/books?id=KCTyy64_IO8C (accessed on 30 March 2025).
  6. Marzouk, O.A. Performance analysis of shell-and-tube dehydrogenation module. Int. J. Energy Res. 2017, 41, 604–610. [Google Scholar] [CrossRef]
  7. Vergara, D.; Fernández-Arias, P.; Lampropoulos, G.; Antón-Sancho, Á. Hydrogen Revolution in Europe: Bibliometric Review of Industrial Hydrogen Applications for a Sustainable Future. Energies 2024, 17, 3658. [Google Scholar] [CrossRef]
  8. Altaf, C.T.; Demir, O.; Colak, T.O.; Karagöz, E.; Kurt, M.; Sankir, N.D.; Sankir, M. Decarbonizing the Industry with Green Hydrogen. In Towards Green Hydrogen Generation; John Wiley & Sons, Ltd.: Hoboken, NJ, USA, 2024; pp. 1–48. [Google Scholar] [CrossRef]
  9. Sarkarzadeh, M.; Farsi, M.; Rahimpour, M.R. Modeling and optimization of an industrial hydrogen unit in a crude oil refinery. Int. J. Hydrogen Energy 2019, 44, 10415–10426. [Google Scholar] [CrossRef]
  10. Moradpoor, I.; Syri, S.; Santasalo-Aarnio, A. Green hydrogen production for oil refining—Finnish case. Renew. Sustain. Energy Rev. 2023, 175, 113159. [Google Scholar] [CrossRef]
  11. Alves, J.J.; Towler, G.P. Analysis of Refinery Hydrogen Distribution Systems. Ind. Eng. Chem. Res. 2002, 41, 5759–5769. [Google Scholar] [CrossRef]
  12. Popoola, L.T.; Nwogbu, C.C.; Taura, U.; Asmara, Y.P.; Nwobodo, L.O.; Agbo, A.O. Future prospects towards attaining zero-emission of greenhouse gases from crude oil refinery plants. Clean. Waste Syst. 2025, 11, 100290. [Google Scholar] [CrossRef]
  13. Cheng, Y.; Zheng, R.; Liu, Z.; Xie, Z. Hydrogen-based industry: A prospective transition pathway toward a low-carbon future. Natl. Sci. Rev. 2023, 10, nwad091. [Google Scholar] [CrossRef]
  14. Zhang, Q.; Liu, J.-F.; Gao, Z.-H.; Chen, S.-Y.; Liu, B.-Y. Review on the challenges and strategies in oil and gas industry’s transition towards carbon neutrality in China. Pet. Sci. 2023, 20, 3931–3944. [Google Scholar] [CrossRef]
  15. Gass, R.; Li, Z.; Outbib, R.; Jemei, S.; Hissel, D. An advanced 1D physics-based model for PEM hydrogen fuel cells with enhanced overvoltage prediction. Int. J. Hydrogen Energy 2025, 97, 1108–1125. [Google Scholar] [CrossRef]
  16. Chen, Z.; Shu, C.; Gan, Z.; Cao, J.; Qiu, P.; Sun, X.; Deng, C.; Wu, Y.; Tang, W. Research Progress and Perspectives on Anti-Poisoning Hydrogen Oxidation Reaction Electrocatalysts for Hydrogen Fuel Cells. Small 2025, 21, 2411049. [Google Scholar] [CrossRef] [PubMed]
  17. Li, Y.H.; Choi, S.S.; Rajakaruna, S. An analysis of the control and operation of a solid oxide fuel-cell power plant in an isolated system. IEEE Trans. Energy Convers. 2005, 20, 381–387. [Google Scholar] [CrossRef]
  18. Marzouk, O.A. Compilation of Smart Cities Attributes and Quantitative Identification of Mismatch in Rankings. J. Eng. 2022, 2022, 5981551. [Google Scholar] [CrossRef]
  19. Sun, L.; Wu, G.; Xue, Y.; Shen, J.; Li, D.; Lee, K.Y. Coordinated Control Strategies for Fuel Cell Power Plant in a Microgrid. IEEE Trans. Energy Convers. 2018, 33, 1–9. [Google Scholar] [CrossRef]
  20. Genovese, M.; Corigliano, O.; Piraino, F.; Fragiacomo, P. E-fuels infrastructures for heavy-duty applications: Case study of a refueling facility based on green hydrogen. Int. J. Hydrogen Energy, 2025; in press. [Google Scholar] [CrossRef]
  21. Fozer, D.; Owsianiak, M.; Hauschild, M.Z. Quantifying environmental learning and scaling rates for prospective life cycle assessment of e-ammonia production. Renew. Sustain. Energy Rev. 2025, 213, 115481. [Google Scholar] [CrossRef]
  22. Sun, S.; Jiang, Q.; Zhao, D.; Cao, T.; Sha, H.; Zhang, C.; Song, H.; Da, Z. Ammonia as hydrogen carrier: Advances in ammonia decomposition catalysts for promising hydrogen production. Renew. Sustain. Energy Rev. 2022, 169, 112918. [Google Scholar] [CrossRef]
  23. Klerke, A.; Christensen, C.H.; Nørskov, J.K.; Vegge, T. Ammonia for hydrogen storage: Challenges and opportunities. J. Mater. Chem. 2008, 18, 2304–2310. [Google Scholar] [CrossRef]
  24. 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]
  25. Amorim, J.; Baravian, G.; Sultan, G. Absolute density measurements of ammonia synthetized in N2–H2 mixture discharges. Appl. Phys. Lett. 1996, 68, 1915–1917. [Google Scholar] [CrossRef]
  26. Reuß, M.; Dimos, P.; Léon, A.; Grube, T.; Robinius, M.; Stolten, D. Hydrogen Road Transport Analysis in the Energy System: A Case Study for Germany through 2050. Energies 2021, 14, 3166. [Google Scholar] [CrossRef]
  27. 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.; Springer Nature: Singapore, 2024; pp. 261–272. [Google Scholar] [CrossRef]
  28. Dhingra, S.; Sharma, S.; Jaiswal, A.; Chadha, R.; Suneja, G.; Gupta, A. Can hydrogen fuel cell vehicles drive the future of sustainable transportation? An empirical study. Technol. Sustain. 2025, 4, 181–193. [Google Scholar] [CrossRef]
  29. Marin, G.D.; Naterer, G.F.; Gabriel, K. Rail transportation by hydrogen vs. electrification—Case study for Ontario, Canada, II: Energy supply and distribution. Int. J. Hydrogen Energy 2010, 35, 6097–6107. [Google Scholar] [CrossRef]
  30. 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; Trans Tech Publications Ltd.: Wollerau, Switzerland, 2023; pp. 89–96. [Google Scholar] [CrossRef]
  31. Galimova, T.; Fasihi, M.; Bogdanov, D.; Breyer, C. Impact of international transportation chains on cost of green e-hydrogen: Global cost of hydrogen and consequences for Germany and Finland. Appl. Energy 2023, 347, 121369. [Google Scholar] [CrossRef]
  32. 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.; Springer Nature: Cham, Switzerland, 2024; pp. 123–135. [Google Scholar] [CrossRef]
  33. Liu, Y.; Li, Z.; Huang, X.; Liu, F.; Zhou, F.; Lim, M.K. Uncovering determinants and barriers to hydrogen fuel cell vehicle adoption: Evidence from Chongqing, China. Int. J. Hydrogen Energy 2025, 106, 875–887. [Google Scholar] [CrossRef]
  34. Bianchi, F.R.; Risso, R.; Cardona, L.; Bove, D.; Cannizzaro, F.; Bonardi, L.; Palmisani, E.; Bosio, B. Feasibility analysis of e-Hydrogen, e-Ammonia and e-Methanol synthesis compared with methane to fuel production. Fuel 2025, 384, 133938. [Google Scholar] [CrossRef]
  35. Galindo Cifre, P.; Badr, O. Renewable hydrogen utilisation for the production of methanol. Energy Convers. Manag. 2007, 48, 519–527. [Google Scholar] [CrossRef]
  36. Monnerie, N.; Gan, P.G.; Roeb, M.; Sattler, C. Methanol production using hydrogen from concentrated solar energy. Int. J. Hydrogen Energy 2020, 45, 26117–26125. [Google Scholar] [CrossRef]
  37. Dalena, F.; Senatore, A.; Marino, A.; Gordano, A.; Basile, M.; Basile, A. Chapter 1—Methanol Production and Applications: An Overview. In Methanol; Basile, A., Dalena, F., Eds.; Elsevier: Amsterdam, The Netherlands, 2018; pp. 3–28. [Google Scholar] [CrossRef]
  38. Galimova, T.; Fasihi, M.; Bogdanov, D.; Lopez, G.; Breyer, C. Analysis of green e-methanol supply costs: Domestic production in Europe versus imports via pipeline and sea shipping. Renew. Energy 2025, 241, 122336. [Google Scholar] [CrossRef]
  39. Sillman, J.; Ylä-Kujala, A.; Hyypiä, J.; Kärri, T.; Tuomaala, M.; Soukka, R. Feasibility assessment of e-methanol value chains: Temporal and regional renewable energy, costs, and climate impacts. Appl. Energy 2025, 391, 125887. [Google Scholar] [CrossRef]
  40. Elmeknassi, Y.; He, W.; Adam, A.; Deng, J.; Lou, Z.; Wang, C.; Chen, L. Performance Analysis of Hydrogen-Powered Gas Turbine Engines: A Parametric Study. In Proceedings of the 2024 7th International Conference on Renewable Energy and Power Engineering (REPE), Beijing, China, 25–27 September 2024; pp. 379–385. [Google Scholar] [CrossRef]
  41. Chiesa, P.; Lozza, G.; Mazzocchi, L. Using Hydrogen as Gas Turbine Fuel. J. Eng. Gas. Turbines Power 2005, 127, 73–80. [Google Scholar] [CrossRef]
  42. Stefan, E.; Talic, B.; Larring, Y.; Gruber, A.; Peters, T.A. Materials challenges in hydrogen-fuelled gas turbines. Int. Mater. Rev. 2022, 67, 461–486. [Google Scholar] [CrossRef]
  43. Nyholm, F.; Toppinen, S.; Saxén, H. Holistic Evaluation Method for Concept-Level Study of Power-to-Liquids Technologies for e-Kerosene Production. Ind. Eng. Chem. Res. 2024, 63, 20264–20286. [Google Scholar] [CrossRef]
  44. Song, Z.; Zhao, J. Research on the Integrated Development of Nuclear Energy and Aviation Industry Under the Background of “Dual Carbon” Goals. E3S Web Conf. 2024, 573, 03008. [Google Scholar] [CrossRef]
  45. 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]
  46. Pacheco, L.S.; Hernández-Gutiérrez, L.E. Advancing Direct Air Capture Technologies: From Carbon Removal to Sustainable Aviation Fuels. In Proceedings of the Copernicus Meetings (EGU General Assembly 2025), Vienna, Austria, 27 April–2 May 2025. [Google Scholar] [CrossRef]
  47. Gao, R.; Zhang, C.; Jun, K.-W.; Kim, S.K.; Park, H.-G.; Zhao, T.; Wang, L.; Wan, H.; Guan, G. Transformation of CO2 into liquid fuels and synthetic natural gas using green hydrogen: A comparative analysis. Fuel 2021, 291, 120111. [Google Scholar] [CrossRef]
  48. Quaresma, G.; Magalhães, L.B.; Ferreira, A.F.; Silva, A. E-Kerosene Potential for Commercial Aviation Decarbonization. In Proceedings of the 34th Congress of the International Council of the Aeronautical Sciences (ICAS 2024), Florence, Italy, 9–13 September 2024; ICAS [International Council of the Aeronautical Sciences]: Florence, Italy, 2024. Available online: https://www.icas.org/icas_archive/icas2024/data/papers/icas2024_0680_paper.pdf (accessed on 28 April 2025).
  49. Nyholm, F.; Saxén, H. Evaluation Method for Power-to-Liquids Concepts for e-Kerosene Production. In Computer Aided Chemical Engineering; Manenti, F., Reklaitis, G.V., Eds.; Elsevier: Amsterdam, The Netherlands, 2024; pp. 3349–3354. [Google Scholar] [CrossRef]
  50. Ozkan, M.; Narappa, A.B.; Namboodiri, T.; Chai, Y.; Babu, M.; Jennings, J.S.E.; Gao, Y.; Tasneem, S.; Lam, J.; Talluri, K.R.; et al. Forging a sustainable sky: Unveiling the pillars of aviation e-fuel production for carbon emission circularity. iScience 2024, 27, 109154. [Google Scholar] [CrossRef]
  51. Schomakers, E.-M.; Engelmann, L.; Ziefle, M. Diversity in the acceptance of sustainable aviation fuels: Uncovering varying motivational patterns. Fuel Commun. 2024, 20, 100129. [Google Scholar] [CrossRef]
  52. De Vries, R.; Wolleswinkel, R.E.; Hoogreef, M.; Vos, R. A New Perspective on Battery-Electric Aviation, Part II: Conceptual Design of a 90-Seater. In Proceedings of the AIAA SCITECH 2024 Forum, Orlando, FL, USA, 8–12 January 2024; AIAA [American Institute of Aeronautics and Astronautics]: Orlando, FL, USA, 2024; p. 1490. [Google Scholar] [CrossRef]
  53. Marzouk, O.A. Urban air mobility and flying cars: Overview, examples, prospects, drawbacks, and solutions. Open Eng. 2022, 12, 662–679. [Google Scholar] [CrossRef]
  54. Bhapkar, S.; Sain, C.; Kazula, S. Review and Evaluation of Hydrogen and Air Heat Exchangers for Fuel Cell-Powered Electric Aircraft Propulsion. Eng. Proc. 2025, 90, 62. [Google Scholar] [CrossRef]
  55. Gao, Y.; Jausseme, C.; Huang, Z.; Yang, T. Hydrogen-Powered Aircraft: Hydrogen–electric hybrid propulsion for aviation. IEEE Electrif. Mag. 2022, 10, 17–26. [Google Scholar] [CrossRef]
  56. Gunasekar, P.; Manigandan, S.; Kumar, T.R.P. Hydrogen as the futuristic fuel for the aviation and aerospace industry—Review. Aircr. Eng. Aerosp. Technol. 2020, 93, 410–416. [Google Scholar] [CrossRef]
  57. Marzouk, O.A. A two-step computational aeroacoustics method applied to high-speed flows. Noise Control Eng. J. 2008, 56, 396. [Google Scholar] [CrossRef]
  58. Petrescu, R.V.V.; Machín, A.; Fontánez, K.; Arango, J.C.; Márquez, F.M.; Petrescu, F.I.T. Hydrogen for aircraft power and propulsion. Int. J. Hydrogen Energy 2020, 45, 20740–20764. [Google Scholar] [CrossRef]
  59. Contreras, A.; Yiğit, S.; Özay, K.; Veziroğlu, T.N. Hydrogen as aviation fuel: A comparison with hydrocarbon fuels. Int. J. Hydrogen Energy 1997, 22, 1053–1060. [Google Scholar] [CrossRef]
  60. Gerl, B.; Ronovsky-Bodisch, M.; Ferrari, N.; Berens, M. Fundamentals of Innovative Aircraft Heat Exchanger Integration for Hydrogen–Electric Propulsion. Aerospace 2025, 12, 320. [Google Scholar] [CrossRef]
  61. White, A.S.; Waddington, E.; Merret, J.M.; Ansell, P.J.; Hall, D.K. Trade-Space Assessment of Liquid Hydrogen Propulsion Systems for Electrified Aircraft. J. Aircr. 2025, 62, 270–277. [Google Scholar] [CrossRef]
  62. Karpuk, S.; Freund, Y.; Hanke-Rauschenbach, R. Potential of Hydrogen Fuel Cell Aircraft for Commercial Applications with Advanced Airframe and Propulsion Technologies. Aerospace 2025, 12, 35. [Google Scholar] [CrossRef]
  63. Rukini, A.; Rhamdhani, M.A.; Brooks, G.A.; Van den Bulck, A. Metals Production and Metal Oxides Reduction Using Hydrogen: A Review. J. Sustain. Metall. 2022, 8, 1–24. [Google Scholar] [CrossRef]
  64. Boretti, A. The perspective of hydrogen direct reduction of iron. J. Clean. Prod. 2023, 429, 139585. [Google Scholar] [CrossRef]
  65. Heidari, A.; Niknahad, N.; Iljana, M.; Fabritius, T. A Review on the Kinetics of Iron Ore Reduction by Hydrogen. Materials 2021, 14, 7540. [Google Scholar] [CrossRef] [PubMed]
  66. Spreitzer, D.; Schenk, J. Reduction of Iron Oxides with Hydrogen—A Review. Steel Res. Int. 2019, 90, 1900108. [Google Scholar] [CrossRef]
  67. Chen, Q.; Pan, J.; Guo, Z.; Zhu, D. Optimization of four-component furnace burdens with hydrogen-reduced metallized pellets based on blast furnace performance. Int. J. Hydrogen Energy 2025, 100, 596–607. [Google Scholar] [CrossRef]
  68. Kar, M.K.; Safarian, J. Producing High Purity Metallic Iron from Bauxite Residue Through Hydrogen Reduction Followed by Flux Smelting. ISIJ Int. 2025, 65, 372–383. [Google Scholar] [CrossRef]
  69. Adami, B.; Hoffelner, F.; Zarl, M.A.; Schenk, J. Strategic Selection of a Pre-Reduction Reactor for Increased Hydrogen Utilization in Hydrogen Plasma Smelting Reduction. Processes 2025, 13, 420. [Google Scholar] [CrossRef]
  70. Sun, M.; Pang, K.; Gu, Z.; Shao, S. Analysis of the theory and practice of hydrogen-enriched blast furnace ironmaking. Int. J. Hydrogen Energy 2025, 127, 702–716. [Google Scholar] [CrossRef]
  71. Xing, Y.; Luo, Z.; Li, H.; Wang, X.; Kou, M.; Wu, S.; Zou, Z.; Zhou, H. Modelling in-furnace phenomena in a hydrogen-rich low-carbon reduction smelting furnace: Influence of blast parameters of hydrogen-rich gas. Appl. Therm. Eng. 2025, 273, 126494. [Google Scholar] [CrossRef]
  72. Ho, T.; Karri, V. Basic tuning of hydrogen powered car and artificial intelligent prediction of hydrogen engine characteristics. Int. J. Hydrogen Energy 2010, 35, 10004–10012. [Google Scholar] [CrossRef]
  73. Wróbel, K.; Wróbel, J.; Tokarz, W.; Lach, J.; Podsadni, K.; Czerwiński, A. Hydrogen Internal Combustion Engine Vehicles: A Review. Energies 2022, 15, 8937. [Google Scholar] [CrossRef]
  74. Swain, M.R.; Pappas, J.M.; Adt, R.R.; Escher, W.J.D. Hydrogen-Fueled Automotive Engine Experimental Testing to Provide an Initial Design-Data Base. In Proceedings of the SAE International Congress and Exposition, Warrendale, PA, USA, 1 February 1981; p. 810350. [Google Scholar] [CrossRef]
  75. Wang, H.; Li, L.; Huang, J.; Wang, Z.; Wu, S.; Xu, W.; Wang, F.; Hao, L.; Bi, Y. The design and performance tests of customized water-proof welding rods for local-dry underwater welding. Eng. Res. Express 2025. [Google Scholar] [CrossRef]
  76. Aderouane, F. (Ed.) Materials and Welding. In Design, Construction, and Global Performance Analysis of Offshore Wind Turbines; Springer Nature: Cham, Switzerland, 2025; pp. 15–38. [Google Scholar] [CrossRef]
  77. Kunar, S.; Mandal, G. Advanced Welding Technologies; John Wiley & Sons: Hoboken, NJ, USA, 2025; ISBN 978-1-394-33190-1. [Google Scholar]
  78. Suban, M.; Tušek, J.; Uran, M. Use of hydrogen in welding engineering in former times and today. J. Mater. Process. Technol. 2001, 119, 193–198. [Google Scholar] [CrossRef]
  79. Zhao, Q.; Li, H.; Lv, J.; Liu, X.; Zhang, F.; Jiang, S.; Ma, L.; Wang, C.; Ni, J.; Peng, G. Adhesive-free bonding fiber optic Fabry–Perot pressure sensor based on oxy-hydrogen flame welding and spiral tube. Opt. Commun. 2020, 476, 126307. [Google Scholar] [CrossRef]
  80. Attah, J.; Mohammed, L.; Nyamful, A.; Donkor, P.; Asamoah, A.; Zainudeen, M.N.; Adjah, J.; Klutse, C.K.; Birikorang, S.A.; Agyemang, F.; et al. Oxy-hydrogen gas as a sustainable fuel for the welding industry: Alternative for oxy-acetylene gas. Clean. Energy Syst. 2024, 9, 100160. [Google Scholar] [CrossRef]
  81. Arshian, S.; Firoozbakhtian, A.; Salehnia, F.; Naghavi Sheikholeslami, M.; Hamidipanah, Y.; Gilnezhad, J.; Reza Ganjali, M.; Hosseini, M. Ti3C2/Ni/Sm-Based Screen-Printed-Electrode for the ECL Detection of Hydrogen Peroxide as a Milk Preservative. ChemistrySelect 2024, 9, e202304462. [Google Scholar] [CrossRef]
  82. Alwazeer, D. Consumption of Hydrogen-Treated Foods Provides Nutritional and Health Benefits. In Molecular Hydrogen in Health and Disease; Slezak, J., Kura, B., Eds.; Springer Nature: Cham, Switzerland, 2024; pp. 319–337. [Google Scholar] [CrossRef]
  83. Russell, G.; Nenov, A.; Hancock, J.T. How Hydrogen (H2) Can Support Food Security: From Farm to Fork. Appl. Sci. 2024, 14, 2877. [Google Scholar] [CrossRef]
  84. Keçebaş, A.; Kayfeci, M. Chapter 1—Hydrogen properties. In Solar Hydrogen Production; Calise, F., D’Accadia, M.D., Santarelli, M., Lanzini, A., Ferrero, D., Eds.; Academic Press: New York, NY, USA, 2019; pp. 3–29. [Google Scholar] [CrossRef]
  85. Allen, R.R. Hydrogenation. J. Am. Oil Chem. Soc. 1981, 58, 166–169. [Google Scholar] [CrossRef]
  86. Jovanovic, D.; Radovic, R.; Mares, L.; Stankovic, M.; Markovic, B. Nickel hydrogenation catalyst for tallow hydrogenation and for the selective hydrogenation of sunflower seed oil and soybean oil. Catal. Today 1998, 43, 21–28. [Google Scholar] [CrossRef]
  87. Anikina, I.D.; Sergeyev, V.V.; Amosov, N.T.; Luchko, M.G. Use of heat pumps in turbogenerator hydrogen cooling systems at thermal power plant. Int. J. Hydrogen Energy 2017, 42, 636–642. [Google Scholar] [CrossRef]
  88. Hara, S.; Iwami, Y.; Kawasaki, R.; Matsumoto, T.; Shirai, Y.; Shiotsu, M.; Kobayashi, H.; Naruo, Y.; Nonaka, S.; Inatani, Y.; et al. Development of Liquid Hydrogen Cooling System for a Rotor of Superconducting Generator. IEEE Trans. Appl. Supercond. 2021, 31, 5202505. [Google Scholar] [CrossRef]
  89. Snell, D.S. The hydrogen-cooled turbine generator. Electr. Eng. 1940, 59, 35–50. [Google Scholar] [CrossRef]
  90. Gray, R.F.; Montgomery, L.; Nelson, R.; Pipkin, J.; Joki-Korpel, S.; Caguiat, F. Designing the cooling systems for the world’s most powerful turbogenerator—Olkiluoto unit 3. In Proceedings of the 2006 IEEE Power Engineering Society General Meeting, Montreal, QC, Canada, 18–22 June 2006; p. 5. [Google Scholar] [CrossRef]
  91. Mah, A.X.Y.; Ho, W.S.; Bong, C.P.C.; Hassim, M.H.; Liew, P.Y.; Asli, U.A.; Kamaruddin, M.J.; Chemmangattuvalappil, N.G. Review of hydrogen economy in Malaysia and its way forward. Int. J. Hydrogen Energy 2019, 44, 5661–5675. [Google Scholar] [CrossRef]
  92. De Rinaldis, S. Green Hydrogen as Industrial Complexes Feedstock and Fuel for Maritime Transport; OnePetro: Richardson, TX, USA, 2021. [Google Scholar]
  93. Branco, J.B.; Ferreira, A.C.; Gonçalves, A.P.; Soares, C.O.; Almeida Gasche, T. Synthesis of methanol using copper–f block element bimetallic oxides as catalysts and greenhouse gases (CO2, CH4) as feedstock. J. Catal. 2016, 341, 24–32. [Google Scholar] [CrossRef]
  94. Schwengber, C.A.; Alves, H.J.; Schaffner, R.A.; da Silva, F.A.; Sequinel, R.; Bach, V.R.; Ferracin, R.J. Overview of glycerol reforming for hydrogen production. Renew. Sustain. Energy Rev. 2016, 58, 259–266. [Google Scholar] [CrossRef]
  95. 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]
  96. Espegren, K.; Damman, S.; Pisciella, P.; Graabak, I.; Tomasgard, A. The role of hydrogen in the transition from a petroleum economy to a low-carbon society. Int. J. Hydrogen Energy 2021, 46, 23125–23138. [Google Scholar] [CrossRef]
  97. 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]
  98. 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]
  99. Marzouk, O.A. Radiant Heat Transfer in Nitrogen-Free Combustion Environments. Int. J. Nonlinear Sci. Numer. Simul. 2018, 19, 175–188. [Google Scholar] [CrossRef]
  100. Frank, E.D.; Elgowainy, A.; Reddi, K.; Bafana, A. Life-cycle analysis of greenhouse gas emissions from hydrogen delivery: A cost-guided analysis. Int. J. Hydrogen Energy 2021, 46, 22670–22683. [Google Scholar] [CrossRef]
  101. 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]
  102. 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]
  103. 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]
  104. Shaner, M.R.; Atwater, H.A.; Lewis, N.S.; McFarland, E.W. A comparative technoeconomic analysis of renewable hydrogen production using solar energy. Energy Environ. Sci. 2016, 9, 2354–2371. [Google Scholar] [CrossRef]
  105. 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]
  106. Mostafaeipour, A.; Khayyami, M.; Sedaghat, A.; Mohammadi, K.; Shamshirband, S.; Sehati, M.-A.; Gorakifard, E. Evaluating the wind energy potential for hydrogen production: A case study. Int. J. Hydrogen Energy 2016, 41, 6200–6210. [Google Scholar] [CrossRef]
  107. 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]
  108. Rezaei, M.; Naghdi-Khozani, N.; Jafari, N. Wind energy utilization for hydrogen production in an underdeveloped country: An economic investigation. Renew. Energy 2020, 147, 1044–1057. [Google Scholar] [CrossRef]
  109. 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]
  110. 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]
  111. Beswick, R.R.; Oliveira, A.M.; Yan, Y. Does the Green Hydrogen Economy Have a Water Problem? ACS Energy Lett. 2021, 6, 3167–3169. [Google Scholar] [CrossRef]
  112. 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), Virtual, 25–27 November 2022; IEEE [Institute of Electrical and Electronics Engineers]: Piscataway, NJ, USA, 2022; pp. 1–6. [Google Scholar] [CrossRef]
  113. 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]
  114. Marzouk, O.A. 2030 Ambitions for Hydrogen, Clean Hydrogen, and Green Hydrogen. Eng. Proc. 2023, 56, 14. [Google Scholar] [CrossRef]
  115. George Davies, W.; Babamohammadi, S.; Yang, Y.; Masoudi Soltani, S. The rise of the machines: A state-of-the-art technical review on process modelling and machine learning within hydrogen production with carbon capture. Gas. Sci. Eng. 2023, 118, 205104. [Google Scholar] [CrossRef]
  116. Riemer, M.; Duscha, V. Carbon capture in blue hydrogen production is not where it is supposed to be—Evaluating the gap between practical experience and literature estimates. Appl. Energy 2023, 349, 121622. [Google Scholar] [CrossRef]
  117. 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]
  118. Sigel, A.; Sigel, H.; Sigel, R.K.O. (Eds.) The Alkali Metal Ions: Their Role for Life; Springer International Publishing: Cham, Switzerland, 2016. [Google Scholar] [CrossRef]
  119. Akitt, J.W. The Alkali and Alkaline Earth Metals. In Multinuclear NMR; Mason, J., Ed.; Springer: Boston, MA, USA, 1987; pp. 189–220. [Google Scholar] [CrossRef]
  120. Kennedy, J.J. The Alkali Metal Cesium and Some of Its Salts; ACS Publications: Washington, DC, USA, 2002. [Google Scholar] [CrossRef]
  121. Messerle, H.K. Magnetohydrodynamic Electrical Power Generation; Wiley: Chichester, UK, 1995. [Google Scholar]
  122. Angrist, S.W. Direct Energy Conversion, 4th ed.; Allyn and Bacon: Boston, MA, USA, 1982. [Google Scholar]
  123. Geri, A.; Salvini, A.; Veca, G.M. MHD linear generator modelling. IEEE Trans. Appl. Supercond. 1995, 5, 465–468. [Google Scholar] [CrossRef]
  124. Zauderer, B.; Tate, E. Electrical characteristics of a linear, nonequilibrium, MHD generator. Proc. IEEE 1968, 56, 1535–1547. [Google Scholar] [CrossRef]
  125. Bünde, R.; Muntenbruch, H.; Raeder, J.; Volk, R.; Zankl, G. MHD Power Generation: Selected Problems of Combustion MHD Generators; Springer: Berlin/Heidelberg, Germany, 1975. [Google Scholar]
  126. Blackman, V.H.; Jones, M.S., Jr.; Demetriades, A. MHD power generation studies in rectangular channels. In 2nd Symposium on the Engineering Aspects of Magnetohydrodynamics (EAMHD-2); Columbia University Press: New York, NY, USA, 1961; pp. 180–210. [Google Scholar]
  127. Kayukawa, N. Open-cycle magnetohydrodynamic electrical power generation: A review and future perspectives. Prog. Energy Combust. Sci. 2004, 30, 33–60. [Google Scholar] [CrossRef]
  128. 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]
  129. WAESD (Westinghouse Advanced Energy Systems Division). MHD Advanced Power Train; Westinghouse: Pittsburgh, PA, USA, 1985.
  130. Nimvari, M.E.; Hadidi, A.; Jafarian, A.; Garjasi, N. Analysis of triple combined cycle with MHD generator as a topping cycle. In Proceedings of the 3rd Conference on Thermal Power Plants, New Delhi, India, 18–19 October 2011; pp. 1–5. Available online: https://ieeexplore.ieee.org/abstract/document/6576978 (accessed on 31 March 2025).
  131. Nichols, L.D. Combined Turbine-Magnetohydrodynamic Brayton Cycle Power System for Space and Ground Use; NASA [United States National Aeronautics and Space Administration]: Cleveland, OH, USA, 1971. Available online: https://books.google.com.om/books?id=WiOTIe5zvBkC (accessed on 31 March 2025).
  132. Cicconardi, S.P.; Perna, A. Performance Analysis of Integrated Systems Based on MHD Generators. Energy Procedia 2014, 45, 1305–1314. [Google Scholar] [CrossRef]
  133. Khalili, S.; Jafarian Dehkordi, A.; Giahi, M.H. Investigating the effect of channel angle of a subsonic MHD (Magneto-Hydro-Dynamic) generator on optimum efficiency of a triple combined cycle. Energy 2015, 85, 543–555. [Google Scholar] [CrossRef]
  134. Esmaeilzadehazimi, M.A.; Manesh, M.H.K.; Heleyleh, B.B.; Modabbaer, H.V. 4E Analysis of Integrated MHD-Combined Cycle. Int. J. Thermodyn. 2019, 22, 219–228. [Google Scholar] [CrossRef]
  135. Murray, E.M. Time Fluctuations of Temperature in a Magnetohydrodynamic Plasma; Union Carbide Corporation, Nuclear Division: Oak Ridge, TN, USA, 1972. Available online: https://www.osti.gov/servlets/purl/4722548 (accessed on 4 April 2025).
  136. Park, B.-S.; Usman, M.; Imran, M.; Pesyridis, A. Review of Organic Rankine Cycle experimental data trends. Energy Convers. Manag. 2018, 173, 679–691. [Google Scholar] [CrossRef]
  137. Miao, Z.; Xu, J.; Yang, X.; Zou, J. Operation and performance of a low temperature organic Rankine cycle. Appl. Therm. Eng. 2015, 75, 1065–1075. [Google Scholar] [CrossRef]
  138. He, Z.; Zhang, Y.; Dong, S.; Ma, H.; Yu, X.; Zhang, Y.; Ma, X.; Deng, N.; Sheng, Y. Thermodynamic analysis of a low-temperature organic Rankine cycle power plant operating at off-design conditions. Appl. Therm. Eng. 2017, 113, 937–951. [Google Scholar] [CrossRef]
  139. Esmaeilzadehazimi, M.A.; Khoshgoftar Manesh, M.H.; Majidi, M.; Nourpour, M. Evaluation of a Novel Quadruple Combined Cycle with the Magnetohydrodynamic Generator Based on 6E Analysis. J. Energy Resour. Technol. 2021, 143, 1–28. [Google Scholar] [CrossRef]
  140. 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]
  141. Nabil, T.; Khairat Dawood, M.M. Enabling efficient use of oxy-hydrogen gas (HHO) in selected engineering applications; transportation and sustainable power generation. J. Clean. Prod. 2019, 237, 117798. [Google Scholar] [CrossRef]
  142. Paparao, J.; Murugan, S. Oxy-hydrogen gas as an alternative fuel for heat and power generation applications—A review. Int. J. Hydrogen Energy 2021, 46, 37705–37735. [Google Scholar] [CrossRef]
  143. NCBI [National Center for Biotechnology Information] PubChem. Compound Summary for CID 11430, Potassium Carbonate. Available online: https://pubchem.ncbi.nlm.nih.gov/compound/11430 (accessed on 31 March 2025).
  144. Dixit, N.S.; Venkatramani, N.; Rohatgi, V.K. Measurement of temperature, electrical conductivity and ion density of seeded combustion plasmas. Energy Convers. Manag. 1987, 27, 103–109. [Google Scholar] [CrossRef]
  145. Avdeev, K.A.; Aksenov, V.S.; Ivanov, V.S.; Medvedev, S.N.; Frolov, S.M.; Frolov, F.S.; Shamshin, I.O. Magnetohydrodynamic effects of heterogeneous spray detonation. Russ. J. Phys. Chem. B 2015, 9, 637–643. [Google Scholar] [CrossRef]
  146. Lee, K.-T.; Gabriela, S.; Chen, W.-H.; Ong, H.C.; Rajendran, S.; Tran, K.-Q. Co-torrefaction and synergistic effect of spent coffee grounds and tea waste for sustainable waste remediation and renewable energy. Renew. Energy 2024, 233, 121181. [Google Scholar] [CrossRef]
  147. Tepper, F.; Murchison, A.; Zelenak, J. Thermophysical Properties of Rubidium and Cesium. In Proceedings of the USAF Aerospace Fluids and Lubricants Conference, San Antonio, TX, USA, 16–19 April 1963; Adamczak, R.L., Ed.; Defense Documentation Center for Scientific and Technical Information: Dayton, OH, USA, 1964; pp. 368–387. Available online: https://books.google.com.om/books?id=UWRUAAAAYAAJ (accessed on 31 March 2025).
  148. Liao, Q.; Tan, Z. Numerical investigations of cold gas dynamic spray with a novel convergent-divergent nozzle. AIP Conf. Proc. 2013, 1558, 2333–2336. [Google Scholar] [CrossRef]
  149. 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]
  150. Deshpande, O.N.; Narappanawar, N. Space optimization through the use of de-laval nozzle and bell nozzle and its theory. In IAENG Transactions on Engineering Sciences; World Scientific: Singapore, 2016; pp. 432–445. [Google Scholar] [CrossRef]
  151. Kanhukamwe, T.N.; Zincume, P.N. Global Landscape of Photovoltaic Module Manufacturing: A Systematic Literature Review. In Proceedings of the 2024 IEEE International Conference on Engineering, Technology, and Innovation (ICE/ITMC), Madeira Island, Portugal, 24–28 June 2024; pp. 1–7. [Google Scholar] [CrossRef]
  152. 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]
  153. Marzouk, O.A. Thermoelectric generators versus photovoltaic solar panels: Power and cost analysis. Edelweiss Appl. Sci. Technol. 2024, 8, 406–428. [Google Scholar] [CrossRef]
  154. Li, G.; Fan, Y.; Li, Q.; Zheng, Y.; Zhao, D.; Wang, S.; Dong, S.; Guo, W.; Tang, Y. A review on micro combustion powered thermoelectric generator: History, state-of-the-art and challenges to commercialization. Renew. Sustain. Energy Rev. 2025, 207, 114897. [Google Scholar] [CrossRef]
  155. Chundawat, J.S.; Kumar, A.; Saini, M. Applications of institutionstic and dual hesitant fuzzy numbers in the reliability evaluation of turbogenerators in thermal power plants. Int. J. Inf. Technol. 2025, 1–9. [Google Scholar] [CrossRef]
  156. 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]
  157. 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, 10–15 June 2007; ASME [American Society of Mechanical Engineers]: San Diego, CA, USA, 2009; pp. 741–752. [Google Scholar] [CrossRef]
  158. 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]
  159. Panchenko, V.P. Preliminary Analysis of the “Sakhalin” World Largest Pulsed MHD Generator. In Proceedings of the 33rd Plasmadynamics and Lasers Conference, Maui, HI, USA, 20–23 May 2002; AIAA [American Institute of Aeronautics and Astronautics]: Maui, HI, USA, 2002; p. 2147. [Google Scholar] [CrossRef]
  160. Panchenko, V.P. 55. Preliminary analysis of the “Sakhalin” world largest pulsed MHD generator. In Proceedings of the 4th Workshop on Magnetoplasma Aerodynamics for Aerospace Applications, Moscow, Russia, 20–23 May 2002; pp. 322–331. [Google Scholar]
  161. Velikhov, E.P.; Pismenny, V.D.; Matveenko, O.G.; Panchenko, V.P.; Yakushev, A.A.; Pisakin, A.V.; Blokh, A.G.; Tkachenko, B.G.; Sergienko, N.M.; Zhukov, B.B.; et al. Pulsed MHD power system “Sakhalin”—The world largest solid propellant fueled MHD generator of 500MWe electric power output. In Proceedings of the 13th International Conference on MHD Electrical Power Generation and High Temperature Technologies, Beijing, China, 12–15 October 1999; pp. 387–398. [Google Scholar]
  162. Veefkind, A.; Houben, J.W.M.A.; Blom, J.H.; Rietjens, L.H.T. High-power density experiments in a shock-tunnel MHD generator. AIAA J. 1976, 14, 1118–1122. [Google Scholar] [CrossRef]
  163. Kirkpatrick, A.T. Energy Flows in Engines. Available online: https://www.engr.colostate.edu/~allan/heat_trans/page3/page3.html (accessed on 31 March 2025).
  164. Hiereth, H.; Prenninger, P. Charging the Internal Combustion Engine; Springer Science & Business Media: Vienna, Austria, 2007. [Google Scholar]
  165. Chen, H.; Lee, C.H.T. Parametric Sensitivity Analysis and Design Optimization of an Interior Permanent Magnet Synchronous Motor. IEEE Access 2019, 7, 159918–159929. [Google Scholar] [CrossRef]
  166. Li, J.; Cheng, L.; Wan, N.; Ma, J.; Hu, Y.; Wen, J. Hybrid harvesting of wind and wave energy based on triboelectric-piezoelectric nanogenerators. Sustain. Energy Technol. Assess. 2023, 60, 103466. [Google Scholar] [CrossRef]
  167. Zhang, X.; Cheng, S.; Wang, X.; Huang, X.; Logan, B.E. Separator Characteristics for Increasing Performance of Microbial Fuel Cells. Environ. Sci. Technol. 2009, 43, 8456–8461. [Google Scholar] [CrossRef]
  168. Kong, X.; Sun, Y.; Li, L.; Li, Y.; Yuan, Z.; Kong, X. Electricity Generation Comparison of Two-Chamber Microbial Fuel Cells with Different Membranes. In Proceedings of the 2010 4th International Conference on Bioinformatics and Biomedical Engineering, Chengdu, China, 18–20 June 2010; pp. 1–4. [Google Scholar] [CrossRef]
  169. Karra, U.; Manickam, S.S.; McCutcheon, J.R.; Patel, N.; Li, B. Power generation and organics removal from wastewater using activated carbon nanofiber (ACNF) microbial fuel cells (MFCs). Int. J. Hydrogen Energy 2013, 38, 1588–1597. [Google Scholar] [CrossRef]
  170. Borghi, C.A.; Massarini, A.; Mazzanti, G. Multidimensional models for the analysis of linear MHD generator channel plasma flows. IEEE Trans. Plasma Sci. 1992, 20, 473–476. [Google Scholar] [CrossRef]
  171. Xue, W.; Miao, L.; Qie, L.; Wang, C.; Li, S.; Wang, J.; Li, J. Gravimetric and volumetric energy densities of lithium-sulfur batteries. Curr. Opin. Electrochem. 2017, 6, 92–99. [Google Scholar] [CrossRef]
  172. Zhai, T.; Lu, X.; Wang, H.; Wang, G.; Mathis, T.; Liu, T.; Li, C.; Tong, Y.; Li, Y. An Electrochemical Capacitor with Applicable Energy Density of 7.4 Wh/kg at Average Power Density of 3000 W/kg. Nano Lett. 2015, 15, 3189–3194. [Google Scholar] [CrossRef]
  173. Janovec, M.; Babčan, V.; Kandera, B.; Šajbanová, K.; Škultéty, F.; Halvoň, Ľ. Performance and Weight Parameters Calculation for Hydrogen- and Battery-Powered Aircraft Concepts. Aerospace 2023, 10, 482. [Google Scholar] [CrossRef]
  174. Blinov, V.N.; Vavilov, I.S.; Kositsin, V.V.; Lukyanchik, A.I.; Ruban, V.I.; Shalay, V.V. Study of power-to-weight ratio of the electrothermal propulsion system of nanosatellite maneuvering satellite platform. J. Phys. Conf. Ser. 2018, 944, 012020. [Google Scholar] [CrossRef]
  175. 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]
  176. Ma, Y.; Karpuk, S.; Elham, A. Conceptual design and comparative study of strut-braced wing and twin-fuselage aircraft configurations with ultra-high aspect ratio wings. Aerosp. Sci. Technol. 2022, 121, 107395. [Google Scholar] [CrossRef]
  177. Cinar, G.; Cai, Y.; Chakraborty, I.; Mavris, D.N. Sizing and Optimization of Novel General Aviation Vehicles and Propulsion System Architectures. In Proceedings of the 2018 Aviation Technology, Integration, and Operations Conference, Atlanta, GA, USA, 25–29 June 2018; AIAA [American Institute of Aeronautics and Astronautics]: Atlanta, GA, USA, 2018; p. 3974. [Google Scholar] [CrossRef]
  178. Mikhaylik, Y.; Kovalev, I.; Xu, J.; Schock, R. Rechargeable Li2S Battery with Specific Energy 350 Wh/kg and Specific Power 3000 W/kg. ECS Trans. 2008, 13, 53. [Google Scholar] [CrossRef]
  179. Yang, Y.; McDowell, M.T.; Jackson, A.; Cha, J.J.; Hong, S.S.; Cui, Y. New Nanostructured Li2S/Silicon Rechargeable Battery with High Specific Energy. Nano Lett. 2010, 10, 1486–1491. [Google Scholar] [CrossRef]
  180. Son, Y.; Cha, H.; Jo, C.; Groombridge, A.S.; Lee, T.; Boies, A.; Cho, J.; De Volder, M. Reliable protocols for calculating the specific energy and energy density of Li-Ion batteries. Mater. Today Energy 2021, 21, 100838. [Google Scholar] [CrossRef]
  181. Park, J.O.; Kim, M.; Kim, J.-H.; Choi, K.H.; Lee, H.C.; Choi, W.; Ma, S.B.; Im, D. A 1000 Wh kg−1 Li–Air battery: Cell design and performance. J. Power Sources 2019, 419, 112–118. [Google Scholar] [CrossRef]
  182. Seikel, G.R. Coal-Fired Open-Cycle MHD Plants. In The Science and Technology of Coal and Coal Utilization; Cooper, B.R., Ellingson, W.A., Eds.; Springer: Boston, MA, USA, 1984; pp. 307–337. [Google Scholar] [CrossRef]
  183. Ishikawa, M.; Fujita, M.; Fujita, Y.; Umoto, J. Performance of coal-fired MHD generators with large leakage current. Energy Convers. Manag. 1993, 34, 607–617. [Google Scholar] [CrossRef]
  184. 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]
  185. Ishikawa, M.; Kyogoku, A.; Umoto, J. Stability of large-scale MHD channels designed for coal-fired MHD power generation. Energy Convers. Manag. 1996, 37, 31–41. [Google Scholar] [CrossRef]
  186. 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]
  187. Sugita, H.; Matsuo, T.; Inui, Y.; Ishikawa, M. Two-dimensional behavior of gas-particle two-phase flow under strong MHD interaction. In Proceedings of the 13th International Conference on MHD Electrical Power Generation and High Temperature Technologies, Beijing, China, 12–15 October 1999; pp. 453–462. [Google Scholar]
  188. 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]
  189. Matsuo, T.; Sugita, H.; Ishikawa, M.; Zeigarnik, V.A. Boundary-layer separation and generator performance of self-excited pulsed MHD channel with strong MHD interaction. In Proceedings of the 13th International Conference on MHD Electrical Power Generation and High Temperature Technologies, Beijing, China, 12–15 October 1999; pp. 399–408. [Google Scholar]
  190. 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]
  191. 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), Las Vegas, NV, USA, 4–7 September 2007; ASME [American Society of Mechanical Engineers]: Las Vegas, NV, USA, 2009; pp. 1755–1774. [Google Scholar] [CrossRef]
  192. 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]
  193. Sugita, H.; Matsuo, T.; Inui, Y.; Ishikawa, M. Two-dimensional analysis of gas-particle two phase flow in pulsed MHD channel. In Proceedings of the 30th Plasmadynamic and Lasers Conference, Norfolk, VA, USA, 1–28 June 1999; American Institute of Aeronautics and Astronautics: Norfolk, VA, USA, 1999; p. 3483. [Google Scholar] [CrossRef]
  194. Marzouk, O.A. Flow control using bifrequency motion. Theor. Comput. Fluid. Dyn. 2011, 25, 381–405. [Google Scholar] [CrossRef]
  195. Hardianto, T.; Sakamoto, N.; Harada, N. Three-Dimensional Flow Analysis in a Faraday-Type MHD Generator. IEEE Trans. Ind. Appl. 2008, 44, 1116–1123. [Google Scholar] [CrossRef]
  196. 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]
  197. Hynes, A.J.; Steinberg, M.; Schofield, K. The chemical kinetics and thermodynamics of sodium species in oxygen-rich hydrogen flames. J. Chem. Phys. 1984, 80, 2585–2597. [Google Scholar] [CrossRef]
  198. 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]
  199. Cheng, T.S.; Chao, Y.-C.; Wu, C.-Y.; Li, Y.-H.; Nakamura, Y.; Lee, K.-Y.; Yuan, T.; Leu, T.S. Experimental and numerical investigation of microscale hydrogen diffusion flames. Proc. Combust. Inst. 2005, 30, 2489–2497. [Google Scholar] [CrossRef]
  200. Aoyagi, M.; Ito, S.; Hashizume, H.; Muroga, T. MHD pressure drop characteristics in a three-surface-multi-layered channel under a strong magnetic field. Fusion. Eng. Des. 2010, 85, 1181–1184. [Google Scholar] [CrossRef]
  201. Huang, X.; Cao, Y.; Li, J.; Yi, W.; Qu, Y. Enhanced reversible magnetocaloric effect in Ni-Co-Mn-Sb-based magnetic shape memory alloy achieved by Ti substitution for Ni. J. Alloys Compd. 2025, 1010, 178112. [Google Scholar] [CrossRef]
  202. Hurley, M.E.; Bollineni, R.K.; Donald, A.M.; Flynn, S.; Hamlin, J.J.; Kesler, M.S.; Manuel, M.V.; Meisel, M.W.; Li, L.; Miller, V.M. Microstructural Evolution of Steel During Magnetic Field-Assisted Processing. JOM 2025, 77, 2862–2874. [Google Scholar] [CrossRef]
  203. Bityurin, V.A.; Borghi, C.A.; Ribani, P.L. High enthalpy extraction numerical experiment in a plasma vane MHD generator. IEEE Trans. Plasma Sci. 1995, 23, 844–851. [Google Scholar] [CrossRef]
  204. Ishikwa, M.; Yuhara, M.; Fujino, T. Three-dimensional computation of magnetohydrodynamics in a weakly ionized plasma with strong MHD interaction. J. Mater. Process. Technol. 2007, 181, 254–259. [Google Scholar] [CrossRef]
  205. Ishikawa, M.; Koshiba, Y.; Matsushita, T. Effects of induced magnetic field on large scale pulsed MHD generator with two phase flow. Energy Convers. Manag. 2004, 45, 707–724. [Google Scholar] [CrossRef]
  206. Ishikawa, M.; Koshiba, Y. 52. Preliminary analysis of large pulsed MHD generator. In Proceedings of the 3th Workshop on Magnetoplasma Aerodynamics for Aerospace Applications, Moscow, Russia, 24–26 April 2001; pp. 276–281. Available online: https://apps.dtic.mil/sti/pdfs/ADA407842.pdf (accessed on 8 February 2025).
  207. 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]
  208. Hustad, C.-W.; Coleman, D.L.; Mikus, T. Technology Overview for Integration of an MHD Topping Cycle with the CES Oxyfuel Combustor; CO2-Global. 2009. Available online: https://co2.no/wp-content/uploads/2020/07/MHD_Report-Final.pdf (accessed on 12 June 2011).
  209. Marzouk, O.A. Evolutionary Computing Applied to Design Optimization. In Proceedings of the ASME 2007 International Design Engineering Technical Conferences and Computers and Information in Engineering Conference (IDETC-CIE 2007), Las Vegas, NV, USA, 4–7 September 2007; ASME [American Society of Mechanical Engineers]: Las Vegas, NV, USA, 2009; pp. 995–1003. [Google Scholar] [CrossRef]
  210. Miller, E.K. Exploring the behavior of a matched load in the time domain. IEEE Antennas Propag. Mag. 2014, 56, 103–111. [Google Scholar] [CrossRef]
  211. McCune, J. Non-Linear Effects of Fluctuations on MHD Performance. In Proceedings of the 6th Symposium on Engineering Aspects of Magnetohydrodynamics, Pittsburgh, PA, USA, 21–22 April 1965; AIAA [American Institute of Aeronautics and Astronautics]: Pittsburgh, PA, USA, 1965; pp. 89–103. [Google Scholar] [CrossRef]
  212. 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]
  213. 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]
  214. Frost, L.S. Conductivity of Seeded Atmospheric Pressure Plasmas. J. Appl. Phys. 1961, 32, 2029–2036. [Google Scholar] [CrossRef]
  215. Raeder, J. Chapter 2—Theory. In MHD Power Generation—Selected Problems of Combustion MHD Generators; Springer: Berlin/Heidelberg, Germany, 1975; pp. 5–85. [Google Scholar]
  216. Khomkin, A.L.; Shumikhin, A.S. Equation of state and conductivity of aluminum dense vapor plasma. In Proceedings of the 35th EPS Conference on Plasma Physics (2008-Europhysics), Crete, Greece, 9–13 June 2008; pp. 720–723. Available online: https://citeseerx.ist.psu.edu/document?repid=rep1&type=pdf&doi=624078677a5a7acac621fa9651ac08e77a1839b9 (accessed on 31 March 2025).
  217. Ebeling, W.; Fortov, V.E.; Klimontovich, Y.L.; Kovalenko, N.P.; Kraeft, W.D.; Krasny, Y.P.; Kremp, D.; Kulik, P.P.; Riaby, V.A.; Röpke, G.; et al. Transport Properties of Dense Plasmas; Springer: Basel, Switzerland, 2013; Available online: https://books.google.com.om/books?id=xUigBgAAQBAJ (accessed on 31 March 2025).
  218. 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]
  219. Gavioso, R.M. Determination of the universal gas constant R and other metrological applications of speed of sound measurements. In Recent Advances in Metrology and Fundamental Constants; IOS Press: Amsterdam, Netherlands, 2001; pp. 195–215. [Google Scholar] [CrossRef]
  220. Marzouk, O.A. Assessment of Three Databases for the NASA Seven-Coefficient Polynomial Fits for Calculating Thermodynamic Properties of Individual Species. Int. J. Aeronaut. Sci. Aerosp. Res. 2018, 5, 150–163. [Google Scholar] [CrossRef]
  221. Šafarík, P.; Nový, A.; Jícha, D.; Hajšman, M. On the Speed of Sound in Steam. Acta Polytech. 2015, 55, 422–426. [Google Scholar] [CrossRef]
  222. McBride, B.J.; Zehe, M.J.; Gordon, S. NASA Glenn Coefficients for Calculating Thermodynamic Properties of Individual Species; NASA [United States National Aeronautics and Space Administration]: Cleveland, OH, USA, 2002. Available online: https://ntrs.nasa.gov/api/citations/20020085330/downloads/20020085330.pdf (accessed on 11 October 2011).
  223. Smirnov, V.N. Calculation of strong-collision dissociation rate constants from NASA thermodynamic polynomials. Int. J. Chem. Kinet. 2020, 52, 559–579. [Google Scholar] [CrossRef]
  224. Oliva, A.A.; Jemcov, A. Method for Efficient Evaluation of Temperature Using the NASA Polynomials. AIAA J. 2024, 62, 405–408. [Google Scholar] [CrossRef]
  225. Euclides, H.O.; Barreto, P.R.P. APUAMA: A software tool for reaction rate calculations. J. Mol. Model. 2017, 23, 176. [Google Scholar] [CrossRef]
  226. Laney, C.B. Computational Gasdynamics, 1st ed.; Cambridge University Press: Cambridge, UK, 1998. [Google Scholar] [CrossRef]
  227. 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).
  228. Akasaka, R.; Huber, M.L.; Simoni, L.D.; Lemmon, E.W. A Helmholtz Energy Equation of State for trans-1,1,1,4,4,4-Hexafluoro-2-butene [R-1336mzz(E)] and an Auxiliary Extended Corresponding States Model for the Transport Properties. Int. J. Thermophys. 2023, 44, 50. [Google Scholar] [CrossRef]
  229. Bespyatov, M.A. Low-Temperature Heat Capacity and Thermodynamic Functions of Europium(III) Heptafluorodimethyloctanedionate. J. Chem. Eng. Data 2023, 68, 3222–3227. [Google Scholar] [CrossRef]
  230. 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]
  231. MacMillan, D. NIST Data Gateway. Ref. Rev. 2012, 26, 43–44. [Google Scholar] [CrossRef]
  232. Engel, T.; Gasteiger, J. Chemoinformatics: Basic Concepts and Methods; John Wiley & Sons: Weinheim, Germany, 2018; Available online: https://books.google.com.om/books?id=X7paDwAAQBAJ (accessed on 31 March 2025).
  233. Chan, B. High-Level Quantum Chemistry Reference Heats of Formation for a Large Set of C, H, N, and O Species in the NIST Chemistry Webbook and the Identification and Validation of Reliable Protocols for Their Rapid Computation. J. Phys. Chem. A 2022, 126, 4981–4990. [Google Scholar] [CrossRef] [PubMed]
  234. NIST [United States National Institute of Standards and Technology]. NIST Chemistry WebBook—Water. 2023. Available online: https://webbook.nist.gov/cgi/cbook.cgi?Name=h2o (accessed on 15 August 2023).
  235. NIST [United States National Institute of Standards and Technology]. NIST Chemistry WebBook—Nitrogen. 2022. Available online: https://webbook.nist.gov/cgi/cbook.cgi?Name=n2 (accessed on 7 May 2022).
  236. NIST [United States National Institute of Standards and Technology]. NIST Chemistry WebBook—Cesium. 2023. Available online: https://webbook.nist.gov/cgi/cbook.cgi?Name=cesium (accessed on 15 August 2023).
  237. NIST [United States National Institute of Standards and Technology]. NIST Chemistry WebBook—Potassium. 2025. Available online: https://webbook.nist.gov/cgi/cbook.cgi?Name=potassium (accessed on 31 March 2025).
  238. Kuo, K.K. Principles of Combustion, 2nd ed.; John Wiley: Hoboken, NJ, USA, 2005. [Google Scholar]
  239. Poinsot, T.; Veynante, D. Theoretical and Numerical Combustion, 2nd ed.; Edwards: Spring Hope, NC, USA, 2005. [Google Scholar]
  240. Ram, N.K.; Singh, N.R.; Raman, P.; Kumar, A.; Kaushal, P. A detailed experimental analysis of air–steam gasification in a dual fired downdraft biomass gasifier enabling hydrogen enrichment in the producer gas. Energy 2019, 187, 115937. [Google Scholar] [CrossRef]
  241. Taniguchi, T.; Kawamura, K.; Sakamoto, S.; Tagashira, H. Three-body attachment in oxygen and an air-like nitrogen and oxygen mixture. J. Phys. Appl. Phys. 1982, 15, 1187. [Google Scholar] [CrossRef]
  242. Elyamani, M.A. The Role of Oxygen-Enriched Air in Sustainable Hydrogen Production and Combustion Performance in Fired Heaters. In Proceedings of the Mediterranean Offshore Conference, Alexandria, Egypt, 20–22 October 2024; OnePetro: Alexandria, Egypt, 2024. [Google Scholar] [CrossRef]
  243. Mariani, M. Post-combustion CO2: Separation And Stocking. WIT Trans. Ecol. Environ. 2006, 86, 785–791. [Google Scholar]
  244. Lu, X.; Hu, E.; Li, X.; Ku, J.; Huang, Z. Non-monotonic behaviors of laminar burning velocities of H2/O2/He mixtures at elevated pressures and temperatures. Int. J. Hydrogen Energy 2017, 42, 22036–22045. [Google Scholar] [CrossRef]
  245. Cooper, C.D.; Naff, W.T.; Compton, R.N. Negative ion properties of p-benzoquinone: Electron affinity and compound states. J. Chem. Phys. 1975, 63, 2752–2757. [Google Scholar] [CrossRef]
  246. NCBI [National Center for Biotechnology Information] PubChem. Ionization Energy in the Periodic Table of Elements. Available online: https://pubchem.ncbi.nlm.nih.gov/periodic-table/ionization-energy (accessed on 2 April 2025).
  247. Compton, R.N.; Cooper, C.D. Molecular electron affinities from collisional ionization of cesium. II. SF6 and TeF6. J. Chem. Phys. 1973, 59, 4140–4144. [Google Scholar] [CrossRef]
  248. Softusvista Unitsconverters. eV/Particle to KJ/mol (Electron Volt Per Particle to KiloJoule Per Mole). Available online: https://www.unitsconverters.com/ru/D-%D0%9A-Millenium/Utu-97-7744 (accessed on 4 April 2025).
  249. Nizzi, K.E.; Pommerening, C.A.; Sunderlin, L.S. Gas-Phase Thermochemistry of Polyhalide Anions. J. Phys. Chem. A 1998, 102, 7674–7679. [Google Scholar] [CrossRef]
  250. Gibb, T.C. Principles of Mössbauer Spectroscopy; Science Paperbacks: New York, NY, USA, 1976; Available online: https://books.google.com.om/books?id=6vjtCAAAQBAJ (accessed on 4 April 2025).
  251. Civiš, S.; Ferus, M.; Kubelík, P.; Jelinek, P.; Chernov, V.E. Potassium spectra in the 700–7000 cm−1 domain: Transitions involving f-, g-, and h-states. Astron. Astrophys. 2012, 541, A125. [Google Scholar] [CrossRef]
  252. Kotarba, A.; Adamski, G.; Sojka, Z.; Djega-Mariadassou, G. Potassium surface stability and electronic promotion in K-NbN0.9O0.1 catalysts. Appl. Surf. Sci. 2000, 161, 105–108. [Google Scholar] [CrossRef]
  253. Schmidt, L.; Gomer, R. Adsorption of Potassium on Tungsten. J. Chem. Phys. 1965, 42, 3573–3598. [Google Scholar] [CrossRef]
  254. Englert, B.-G. Statistical Atom: Ionization Energies. Z. Für Naturforschung A 1987, 42, 825–834. [Google Scholar] [CrossRef]
  255. Crawford, J.R.; Kunz, P.; Yang, H.; Schaffer, P.; Ruth, T.J. e211Rn/e211At and e209At production with intense mass separated Fr ion beams for preclinical 211At-based α-therapy research. Appl. Radiat. Isot. 2017, 122, 222–228. [Google Scholar] [CrossRef]
  256. RSC [Royal Society of Chemistry]. Francium—Element Information, Properties and Uses. Available online: https://periodic-table.rsc.org/element/87/francium (accessed on 2 April 2025).
  257. NCBI [National Center for Biotechnology Information] PubChem. Francium (Fr). Available online: https://pubchem.ncbi.nlm.nih.gov/element/87 (accessed on 2 April 2025).
  258. Venetskii, S. Francium. Metallurgist 1978, 22, 61–65. [Google Scholar] [CrossRef]
  259. Pocsai, M.A.; Barna, I.F.; Tökési, K. Photoionisation of rubidium in strong laser fields. Eur. Phys. J. D 2019, 73, 74. [Google Scholar] [CrossRef]
  260. Roman, V.; Kupliauskienė, A.; Borovik, A. Excitation and ionization of outer shells in Rb by electron impact. J. Phys. B At. Mol. Opt. Phys. 2015, 48, 205204. [Google Scholar] [CrossRef]
  261. RSC [Royal Society of Chemistry]. Rubidium—Element Information, Properties and Uses. Available online: https://periodic-table.rsc.org/element/37/Rubidium (accessed on 2 April 2025).
  262. Xing, P.; Wang, C.; Chen, Y.; Ma, B. Rubidium extraction from mineral and brine resources: A review. Hydrometallurgy 2021, 203, 105644. [Google Scholar] [CrossRef]
  263. Sharma, S.K.; Truong, D.Q.; Guo, J.; An, A.K.; Naidu, G.; Deka, B.J. Recovery of rubidium from brine sources utilizing diverse separation technologies. Desalination 2023, 556, 116578. [Google Scholar] [CrossRef]
  264. Mackay, K.K.; Freund, J.B.; Johnson, H.T. Enhancement of hydrogen microcombustion via field-emission dielectric barrier discharge. Plasma Sources Sci. Technol. 2018, 27, 085007. [Google Scholar] [CrossRef]
  265. Choi, K.C.; Kim, S.H.; Shin, B.J.; Kang, J.; Choi, K.-Y.; Yoo, E.-H. Effects of Kr,hboxN2, and Ar on Address Discharge Time Lag in AC Plasma-Display Panel with High Xenon Content. IEEE Trans. Electron. Devices 2006, 53, 2410–2413. [Google Scholar] [CrossRef]
  266. Mohandas, S.; Ramabhadran, R.O.; Kumar, S.S. Theoretical Investigation of a Vital Step in the Gas-Phase Formation of Interstellar Ammonia NH2+ + H2 → NH3+ + H. J. Phys. Chem. A 2020, 124, 8373–8382. [Google Scholar] [CrossRef] [PubMed]
  267. Pia, M.G.; Seo, H.; Batic, M.; Begalli, M.; Kim, C.H.; Quintieri, L.; Saracco, P. Evaluation of Atomic Electron Binding Energies for Monte Carlo Particle Transport. IEEE Trans. Nucl. Sci. 2011, 58, 3246–3268. [Google Scholar] [CrossRef]
  268. Wiles, D.M. The Reactions of Active Nitrogen with Phosphine and Hydrogen Chloride. Ph.D. Thesis, Department of Chemistry, McGill University, Montreal, QC, Canada, 1957. Available online: https://escholarship.mcgill.ca/concern/theses/qr46r4819 (accessed on 4 April 2025).
  269. Mitra, S.K. Atomic Nitrogen in Auroras. Nature 1951, 167, 897. [Google Scholar] [CrossRef] [PubMed]
  270. Kumar, A.; Kołaski, M.; Lee, H.M.; Kim, K.S. Photoexcitation and Photoionization Dynamics of Water Photolysis. J. Phys. Chem. A 2008, 112, 5502–5508. [Google Scholar] [CrossRef]
  271. Misaizu, F.; Sanekata, M.; Tsukamoto, K.; Fuke, K.; Iwata, S. Photodissociation of size-selected aquamagnesium (Mg+(H2O)n) ions for n = 1 and 2. J. Phys. Chem. 1992, 96, 8259–8264. [Google Scholar] [CrossRef]
  272. Wang, C.; Liu, M.; Li, B.; Liu, Y.; Yan, J. Thermodynamic analysis on the transient cycling of coal-fired power plants: Simulation study of a 660 MW supercritical unit. Energy 2017, 122, 505–527. [Google Scholar] [CrossRef]
  273. Boonnasa, S.; Namprakai, P. Sensitivity analysis for the capacity improvement of a combined cycle power plant (100–600 MW). Appl. Therm. Eng. 2008, 28, 1865–1874. [Google Scholar] [CrossRef]
  274. Chen, W.; Zhang, G.; Li, B.; Liu, M.; Liu, J. Simulation study on 660 MW coal-fired power plant coupled with a steam ejector to ensure NOx reduction ability. Appl. Therm. Eng. 2017, 111, 550–561. [Google Scholar] [CrossRef]
  275. Mason, J.E.; Archer, C.L. Baseload electricity from wind via compressed air energy storage (CAES). Renew. Sustain. Energy Rev. 2012, 16, 1099–1109. [Google Scholar] [CrossRef]
  276. Massarini, A.; Borghi, C.A. Time-dependent quasi-one-dimensional flow models for linear magnetohydrodynamic generator channels. Phys. Fluids B Plasma Phys. 1992, 4, 2823–2829. [Google Scholar] [CrossRef]
  277. Borghi, C.A.; Cristofolini, A.; Ribani, P.L. Analysis of magneto-plasma dynamic transients in a combustion gas magnetohydrodynamic generator. Phys. Plasmas 1997, 4, 3082–3090. [Google Scholar] [CrossRef]
  278. Cristofolini, A.; Borghi, C.A. A difference method for the solution of the electrodynamic problem in a magnetohydrodynamic field. IEEE Trans. Magn. 1995, 31, 1749–1752. [Google Scholar] [CrossRef]
  279. Borghi, C.A.; Massarini, A.; Mazzanti, G.; Ribani, P.L. Steady State Descriptions of MHD Plasma Flows; International Academic Publishers: Lausanne, Switzerland, 1992; pp. 770–775. [Google Scholar]
  280. Wang, X.; Barnett, A. The Evolving Value of Photovoltaic Module Efficiency. Appl. Sci. 2019, 9, 1227. [Google Scholar] [CrossRef]
  281. 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]
  282. Dubey, S.; Sarvaiya, J.N.; Seshadri, B. Temperature Dependent Photovoltaic (PV) Efficiency and Its Effect on PV Production in the World—A Review. Energy Procedia 2013, 33, 311–321. [Google Scholar] [CrossRef]
  283. Saiki, T.; Takizawa, Y.; Miyahara, K.; Arima, M. Utilizing conductivity of seawater for bioelectric measurement of fish. Sci. Rep. 2020, 10, 16363. [Google Scholar] [CrossRef]
  284. Saiki, T.; Takizawa, Y.; Murai, K.; Okuno, R.; Arima, M. A novel method for noninvasive bioelectric measurement utilizing conductivity of seawater. Sci. Rep. 2021, 11, 7073. [Google Scholar] [CrossRef]
  285. Zheng, Z.; Fu, Y.; Liu, K.; Xiao, R.; Wang, X.; Shi, H. Three-stage vertical distribution of seawater conductivity. Sci. Rep. 2018, 8, 9916. [Google Scholar] [CrossRef]
  286. Zambrano, G.; Riascos, H.; Prieto, P.; Restrepo, E.; Devia, A.; Rincón, C. Optical emission spectroscopy study of r.f. magnetron sputtering discharge used for multilayers thin film deposition. Surf. Coat. Technol. 2003, 172, 144–149. [Google Scholar] [CrossRef]
  287. Jitschin, W.; Reich, G. Molecular velocity distribution at large Knudsen numbers. J. Vac. Sci. Technol. A 1991, 9, 2752–2756. [Google Scholar] [CrossRef]
  288. Li, J.; Kuipers, J.A.M. Effect of pressure on gas–solid flow behavior in dense gas-fluidized beds: A discrete particle simulation study. Powder Technol. 2002, 127, 173–184. [Google Scholar] [CrossRef]
  289. Lynch, G.R.; Dahl, O.I. Approximations to multiple Coulomb scattering. Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. At. 1991, 58, 6–10. [Google Scholar] [CrossRef]
  290. Gámiz, F.; López-Villanueva, J.A.; Jiménez-Tejada, J.A.; Melchor, I.; Palma, A. A comprehensive model for Coulomb scattering in inversion layers. J. Appl. Phys. 1994, 75, 924–934. [Google Scholar] [CrossRef]
  291. Johnson, W.R.; Weber, T.A.; Mullin, C.J. Coulomb Scattering of Polarized Electrons. Phys. Rev. 1961, 121, 933–939. [Google Scholar] [CrossRef]
  292. RSC [Royal Society of Chemistry]. Potassium—Element Information, Properties and Uses. Available online: https://periodic-table.rsc.org/element/19/potassium (accessed on 2 April 2025).
  293. Sarkar, A. Momentum-space properties for the S-states of the valence electron of potassium atom. Eur. Phys. J. D 2022, 76, 118. [Google Scholar] [CrossRef]
  294. Brown, J.M. A New Limit on Lorentz- and CPT-Violating Neutron Spin Interactions Using a Potassium-Helium Comagnetometer. Ph.D. Thesis, Department of Physics, Princeton University, Princeton, NJ, USA, 2011. Available online: https://www.proquest.com/openview/da8c7ee94b771c5271825badfd66f968/1 (accessed on 3 April 2025).
  295. Fritzsche, S.; Jänkälä, K.; Huttula, M.; Urpelainen, S.; Aksela, H. Photoelectron satellite structure from the 3d and 4d inner-shell ionization of rubidium and cesium: Role of atomic relaxation. Phys. Rev. A 2008, 78, 032514. [Google Scholar] [CrossRef]
  296. Webelements. Caesium—Properties of Free Atoms. Available online: https://www.webelements.com/caesium/atoms.html (accessed on 2 April 2025).
  297. Sherson, J.; Krauter, H.; Olsson, R.K.; Julsgaard, B.; Polzik, E.S. Quantum memory and teleportation using macroscopic gas samples. J. Phys. B At. Mol. Opt. Phys. 2008, 41, 223001. [Google Scholar] [CrossRef]
  298. Mori, Y.; Ohtake, K.; Yamamoto, M.; Imani, K. Thermodynamic and Electrical Properties of Combustion Gas and Its Plasma: 1st Report, Theoretical Calculation. Bull. JSME 1968, 11, 241–252. [Google Scholar] [CrossRef]
  299. 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]
  300. Yang, W. A tutorial overview of the angular scattering models of electron–neutral, ion–neutral, neutral–neutral, and Coulomb collisions in Monte Carlo collision modeling on low-temperature plasma. Plasma Sources Sci. Technol. 2024, 33, 023001. [Google Scholar] [CrossRef]
  301. Banks, P. Collision frequencies and energy transfer electrons. Planet. Space Sci. 1966, 14, 1085–1103. [Google Scholar] [CrossRef]
  302. Hjärtstam, S.; Andersson, K.; Johnsson, F.; Leckner, B. Combustion characteristics of lignite-fired oxy-fuel flames. Fuel 2009, 88, 2216–2224. [Google Scholar] [CrossRef]
  303. Bagheri, G.; Ranzi, E.; Pelucchi, M.; Parente, A.; Frassoldati, A.; Faravelli, T. Comprehensive kinetic study of combustion technologies for low environmental impact: MILD and OXY-fuel combustion of methane. Combust. Flame 2020, 212, 142–155. [Google Scholar] [CrossRef]
  304. Dobó, Z. Heat radiation measurement method for high pressure oxy-fuel combustion. Measurement 2018, 124, 191–196. [Google Scholar] [CrossRef]
  305. Davidson, R. Methods in Nonlinear Plasma Theory; Elsevier: New York, NY, USA, 2012; Available online: https://books.google.com.om/books?hl=en&lr=&id=8iW0MDOVr0oC (accessed on 28 April 2025).
  306. Filbet, F.; Negulescu, C.; Yang, C. Numerical study of a nonlinear heat equation for plasma physics. Int. J. Comput. Math. 2012, 89, 1060–1082. [Google Scholar] [CrossRef]
  307. 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, 15–20 June 2008; ASME [American Society of Mechanical Engineers]: Estoril, Portugal, 2009; pp. 901–912. [Google Scholar] [CrossRef]
  308. Kono, M.; Skoric, M. Nonlinear Physics of Plasmas; Springer Science & Business Media: Berlin/Heidelberg, Germany, 2010; Available online: https://books.google.com.om/books?hl=en&lr=&id=fHDM99M49xIC (accessed on 28 April 2025).
  309. Masuda, R.; Matsumoto, M.; Kawasaki, A.; Okuno, Y. Power Generation Characteristics of Disk-Shaped Magnetohydrodynamic Generator Driven by Rotating Detonation. J. Propuls. Power 2025, 41, 125–134. [Google Scholar] [CrossRef]
  310. Wang, S.; Liu, Z.; Huang, H.; Lu, P. Performance investigations of the two-phase mixer for liquid metal magnetohydrodynamic generator. Theor. Comput. Fluid. Dyn. 2025, 39, 18. [Google Scholar] [CrossRef]
  311. Chakradhar, K.; Nandagopal, K.; Prashanthi, V.; Parandhama, A.; Somaiah, T.; Thrinath, B.V.S.; Tarakaramu, N.; Rasool, G.; Abduvalieva, D. MHD effect on peristaltic motion of Williamson fluid via porous channel with suction and injection. Partial. Differ. Equ. Appl. Math. 2025, 13, 101103. [Google Scholar] [CrossRef]
  312. Muehlhauser, J.W. UTSI/CFFF MHD Program Completion and Related Activity (Quarterly Technical Progress Report). 2004. Available online: https://www.osti.gov/servlets/purl/822682 (accessed on 20 September 2023).
  313. Inoue, I.; Lineberry, J.T.; Ishikawa, M.; Umoto, J. Numerical study of the electrodynamics behavior of the CDIF MHD generator. Energy Convers. Manag. 1998, 39, 785–795. [Google Scholar] [CrossRef]
  314. Daniel, V.; Lineberry, J.; Wu, Y. Performance evaluations of MHD generator tests at CDIF. In Proceedings of the 30th Aerospace Sciences Meeting and Exhibit, Reno, NV, USA, 6–9 January 1992; AIAA [American Institute of Aeronautics and Astronautics]: Reno, NV, USA, 1992. [Google Scholar] [CrossRef]
  315. Pian, C.C.P.; Petty, S.W.; Schmitt, E.W.; Farrar, L.C. Prototypic magnetohydrodynamic anode designs and test results. J. Propuls. Power 1995, 11, 1330–1337. [Google Scholar] [CrossRef]
  316. Norton, O.; Wilson, W. Laser Doppler velocity measurements in the diffuser at CDIF. In Proceedings of the 30th Aerospace Sciences Meeting and Exhibit, Reno, NV, USA, 6–9 January 1992; AIAA [American Institute of Aeronautics and Astronautics]: Reno, NV, USA, 1992. [Google Scholar] [CrossRef]
  317. Wang, S.-T.; Genens, L.; Elmer, J.; Brown, E.; McHenry, H. Measurements of mechanical properties of superstructures materials for the CFFF syperconducting MHD magnet. IEEE Trans. Magn. 1981, 17, 2308–2311. [Google Scholar] [CrossRef]
  318. Lineberry, J.; Galanga, F.; Frazier, J. CFFF low mass flow DCW generator operation. In Proceedings of the 24th Aerospace Sciences Meeting, Reno, NV, USA, 6–9 January 1986; AIAA [American Institute of Aeronautics and Astronautics]: Reno, NV, USA, 1986. [Google Scholar] [CrossRef]
  319. Niemann, R.C.; Wang, S.-T.; Dawson, J.W.; Genens, L.; Smith, R.P.; Turner, L.R.; Gonczy, J.D.; Hoffman, J.; Mataya, K.F.; Smelser, P.; et al. Cryogenic Aspects of the UTSI-CFFF Superconducting Dipole Magnet for MHD Research. In Advances in Cryogenic Engineering; Timmerhaus, K.D., Snyder, H.A., Eds.; Springer: Boston, MA, USA, 1980; pp. 30–38. [Google Scholar] [CrossRef]
  320. Lee, J. Three-dimensional predictions of reactive turbulent recirculating flow of a cylindrical MHD type combustor. In Proceedings of the 28th Aerospace Sciences Meeting, Reno, NV, USA, 8–11 January 1990; AIAA [American Institute of Aeronautics and Astronautics]: Reno, NV, USA, 1990. [Google Scholar] [CrossRef]
  321. Bera, T.K. A Magnetohydrodynamic (MHD) Power Generating System: A Technical Review. IOP Conf. Ser. Mater. Sci. Eng. 2020, 955, 012075. [Google Scholar] [CrossRef]
  322. Ovcharenko, V.A. Commercial Realization of MHD—A Challenge for Superconducting Magnets. In Advances in Cryogenic Engineering; Timmerhaus, K.D., Ed.; Springer: Boston, MA, USA, 1978; pp. 1–8. [Google Scholar] [CrossRef]
  323. Malghan, V.R. History of MHD power plant development. Energy Convers. Manag. 1996, 37, 569–590. [Google Scholar] [CrossRef]
  324. Bera, T.K.; Bohre, A.K.; Ahmed, I.; Bhattacharya, A.; Bhowmik, P.S. Magnetohydrodynamic (MHD) Power Generation Systems. In Planning of Hybrid Renewable Energy Systems, Electric Vehicles and Microgrid: Modeling, Control and Optimization; Bohre, A.K., Chaturvedi, P., Kolhe, M.L., Singh, S.N., Eds.; Springer Nature: Singapore, 2022; pp. 905–929. [Google Scholar] [CrossRef]
  325. Macheret, S.; Shneider, M.; Miles, R. Potential performance of supersonic MHD power generators. In Proceedings of the 39th Aerospace Sciences Meeting and Exhibit, Reno, NV, USA, 8–11 January 2001; AIAA [American Institute of Aeronautics and Astronautics]: Reno, NV, USA, 2001; p. 0795. [Google Scholar] [CrossRef]
  326. Domínguez-Lozoya, J.C.; Cuevas, S.; Domínguez, D.R.; Ávalos-Zúñiga, R.; Ramos, E. Laboratory Characterization of a Liquid Metal MHD Generator for Ocean Wave Energy Conversion. Sustainability 2021, 13, 4641. [Google Scholar] [CrossRef]
  327. Ali, H.K.; Braun, R.D. Application of Magnetohydrodynamic Energy Generation to Planetary Entry Vehicles. In Proceedings of the 12th International Energy Conversion Engineering Conference, Cleveland, OH, USA, 28–30 July 2014; AIAA [American Institute of Aeronautics and Astronautics]: Cleveland, OH, USA, 2014; p. 3556. [Google Scholar] [CrossRef]
  328. Panchadar, K.; West, D.; Taylor, J.A.; Krupenkin, T. Mechanical energy harvesting using a liquid metal vortex magnetohydrodynamic generator. Appl. Phys. Lett. 2019, 114, 093901. [Google Scholar] [CrossRef]
  329. Domínguez-Lozoya, J.C.; Domínguez-Lozoya, D.R.; Cuevas, S.; Ávalos-Zúñiga, R.A. MHD Generation for Sustainable Development, from Thermal to Wave Energy Conversion: Review. Sustainability 2024, 16, 10041. [Google Scholar] [CrossRef]
  330. Zhao, L.; Peng, A. Review of Conductive Reciprocating Liquid Metal Magnetohydrodynamic Generators. Energies 2025, 18, 959. [Google Scholar] [CrossRef]
  331. Noda, K.; Kawasaki, A.; Higashino, K. An Experimental Study on Magnetohydrodynamic Power Generation Using a Rotating Detonation Combustor. In Proceedings of the AIAA SCITECH 2025 Forum, Orlando, FL, USA, 6–10 January 2025; AIAA [American Institute of Aeronautics and Astronautics]: Orlando, FL, USA, 2025; p. 203. [Google Scholar] [CrossRef]
  332. Wang, Y.; Huang, H.; Han, X.; Lu, P. The influence of the gas–liquid metal inlet angles on the performance of a magnetohydrodynamic generator. Chem. Eng. J. 2024, 499, 156316. [Google Scholar] [CrossRef]
  333. Wang, Y.; Cheng, K.; Xu, J.; Jing, W.; Huang, H.; Qin, J. A rapid performance prediction method for Two-Phase liquid metal MHD generators based on Quasi-One-Dimensional model. Therm. Sci. Eng. Prog. 2024, 47, 102258. [Google Scholar] [CrossRef]
  334. Pacheco, C.C.; Verissimo, G.L.; Colaço, M.J.; Leiroz, A.J.K.; Cruz, M.E.C.; Santos, H.F.L.; DeFilippo, M.; Quirino, T.S. A computational study on square and helical magnetohydrodynamic generators including applications to a combined power cycle. J. Braz. Soc. Mech. Sci. Eng. 2024, 46, 604. [Google Scholar] [CrossRef]
  335. Kimsor, O.; Okuno, Y. Numerical study of plasma behavior in a disk-shaped noble gas MHD generator. Electron. Commun. Jpn. 2024, 107, e12445. [Google Scholar] [CrossRef]
Figure 1. The hydrogen magnetohydrodynamic generator power system.
Figure 1. The hydrogen magnetohydrodynamic generator power system.
Hydrogen 06 00031 g001
Figure 2. Electric conductivity in the case of cesium seed and air oxidizer.
Figure 2. Electric conductivity in the case of cesium seed and air oxidizer.
Hydrogen 06 00031 g002
Figure 3. Ideal power density in the case of cesium seed and air oxidizer.
Figure 3. Ideal power density in the case of cesium seed and air oxidizer.
Hydrogen 06 00031 g003
Figure 4. Electric conductivity in the case of potassium seed and air oxidizer.
Figure 4. Electric conductivity in the case of potassium seed and air oxidizer.
Hydrogen 06 00031 g004
Figure 5. Ideal power density in the case of potassium seed and air oxidizer.
Figure 5. Ideal power density in the case of potassium seed and air oxidizer.
Hydrogen 06 00031 g005
Figure 6. Electric conductivity in the case of potassium seed and oxygen oxidizer.
Figure 6. Electric conductivity in the case of potassium seed and oxygen oxidizer.
Hydrogen 06 00031 g006
Figure 7. Ideal power density in the case of potassium seed and oxygen oxidizer.
Figure 7. Ideal power density in the case of potassium seed and oxygen oxidizer.
Hydrogen 06 00031 g007
Figure 8. Electric conductivity in the case of cesium seed and oxygen oxidizer.
Figure 8. Electric conductivity in the case of cesium seed and oxygen oxidizer.
Hydrogen 06 00031 g008
Figure 9. Ideal power density in the case of cesium seed and oxygen oxidizer.
Figure 9. Ideal power density in the case of cesium seed and oxygen oxidizer.
Hydrogen 06 00031 g009
Table 1. Example uses of hydrogen.
Table 1. Example uses of hydrogen.
Serial NumberCharacteristicsReferences
1.Oil refinery[7,8,9,10,11,12,13,14]
2.Fuel cell power units[15,16,17,18,19]
3.Synthesizing ammonia[20,21,22,23,24,25]
4.Electrified-type hydrogen-powered transport[26,27,28,29,30,31,32,33]
5.Synthesizing methanol[34,35,36,37,38,39]
6.Gas turbines powered by hydrogen or hydrogen-blended gas[40,41,42]
7.Synthesizing hydrocarbon fuels[43,44,45,46,47,48,49,50,51,52]
8.Aerospace and rocket propulsion[53,54,55,56,57,58,59,60,61,62]
9.Reduction processes to extract a metal from its ore[63,64,65,66,67,68,69,70,71]
10.Combustion-type hydrogen-powered transport[72,73,74]
11.Specialized welding operations[75,76,77,78,79,80]
12.Food industry[81,82,83,84,85,86]
13.Cooling of turbogenerator winding[87,88,89,90]
14.Feedstock for chemical industries[91,92,93,94]
Table 2. Three fixed parameters in the current study.
Table 2. Three fixed parameters in the current study.
Fixed ParameterValueReferences
Temperature2300 K = 2026.85 °C[197,198,199]
Magnetic-field flux density5 T (5 teslas) = 5000 G (50,000 gausses)[200,201,202,203]
Mach number2[204,205,206]
Table 3. Molecular weights used in the current study.
Table 3. Molecular weights used in the current study.
Gaseous SpeciesMolecular Weight [kg/kmol]Reference
Water vapor (H2O)18.0153[234]
Nitrogen (N2)28.0134[235]
Cesium vapor (Cs)132.9054519[236]
Potassium vapor (K)39.0983[237]
Table 4. Fitting coefficients of the specific heat capacities at constant pressure.
Table 4. Fitting coefficients of the specific heat capacities at constant pressure.
Coefficient for (Cp,i)Species Index (i)
H2ON2CsK
b 1 , i 1.034972096 × 1065.877124060 × 1056.166040900 × 106–3.56642236 × 106
b 2 , i –2.412698562 × 103–2.239249073 × 103–1.896175522 × 1041.085289825 × 104
b 3 , i 4.6461107806.0669492202.483229903 × 101–1.054134898 × 101
b 4 , i 2.291998307 × 10–3–6.139685500 × 10–4–1.251977234 × 10–28.009801350 × 10–3
b 5 , i –6.836830480 × 10–71.491806679 × 10–73.309017390 × 10–6–2.696681041 × 10–6
b 6 , i 9.426468930 × 10–11–1.923105485 × 10–11–3.354012020 × 10–104.715294150 × 10–10
b 7 , i –4.822380530 × 10–151.061954386 × 10–159.626500908 × 10–15–2.976897350 × 10–14
Table 5. Varied quantities in the current study.
Table 5. Varied quantities in the current study.
CounterVaried Quantity
p [atm] X s e e d [%]Seed MetalOxidizer
1.0.06250.0625%, (1/16)%Cesium (Cs)Air (79% N2, 21% O2)
2.0.1250.25%Potassium (K)Oxygen (100% O2)
3.0.251%
4.0.54%
5.116%
6.2
7.4
8.8
9.16
Table 6. Pre-ionization properties of cesium-seeded air–hydrogen combustion mixture.
Table 6. Pre-ionization properties of cesium-seeded air–hydrogen combustion mixture.
Property Different Pre-Ionization Cesium Mole Fraction ( X C s )
0.0625%0.25%1%4%16%
X C s [%]0.062500%0.250000%1.000000%4.000000%16.000000%
X H 2 O [%]34.688476%34.623395%34.363068%33.321763%29.156543%
X N 2 [%]65.249024%65.126605%64.636932%62.678237%54.843457%
M m i x [kg/kmol]24.61124.81425.62728.87841.881
R m i x [J/(kg.K)]337.8384335.0721324.4457287.9214198.5258
γ m i x [-]1.2439911.2442651.2453651.2498691.269668
a [m/s]983.166979.241964.014909.773761.408
u [m/s]1966.331958.481928.031819.551522.82
Table 7. Electric conductivity [S/m] of hydrogen plasma (cesium seed, air oxidizer).
Table 7. Electric conductivity [S/m] of hydrogen plasma (cesium seed, air oxidizer).
Pressure [atm] Plasma Electric Conductivity [S/m] (at the Different Cesium Levels)
0.0625%0.25%1%4%6%16%
0.062514.011625.838342.209955.061056.060551.7766
0.12510.285819.242632.175242.640743.401039.6267
0.257.470214.123724.037932.253532.818829.6901
0.55.383010.255117.682923.948324.361621.8837
13.85737.387812.858917.533517.832115.9340
22.75305.29239.272912.705012.918911.4995
41.95933.77626.64709.13799.29048.2470
81.39172.68694.74476.53786.64615.8882
160.98711.90813.37684.66034.73714.1912
Table 8. Power density [MW/m3] of hydrogen plasma (cesium seed and air oxidizer).
Table 8. Power density [MW/m3] of hydrogen plasma (cesium seed and air oxidizer).
Pressure [atm] Volumetric Power Density [MW/m3] (at the Different Cesium Levels)
0.0625%0.25%1%3%4%16%
0.0625338.596619.417980.6651146.8281139.332750.427
0.125248.561461.299747.529886.885882.329574.332
0.25180.520338.585558.474670.107667.395430.315
0.5130.082245.844410.828497.150495.543317.173
193.213177.106298.752363.771362.806230.940
266.527126.871215.438263.484262.894166.669
447.34790.526154.430189.454189.083119.528
833.63164.413110.234135.521135.28185.341
1623.85445.74378.45396.59196.43260.745
Table 9. Pre-ionization properties of potassium-seeded air–hydrogen combustion mixture.
Table 9. Pre-ionization properties of potassium-seeded air–hydrogen combustion mixture.
Property Different Pre-Ionization Potassium Mole Fraction ( X K )
0.0625%0.25%1%4%16%
X K [%]0.062500%0.250000%1.000000%4.000000%16.000000%
X H 2 O [%]34.688476%34.623395%34.363068%33.321763%29.156543%
X N 2 [%]65.249024%65.126605%64.636932%62.678237%54.843457%
M m i x [kg/kmol]24.55224.57924.68925.12526.872
R m i x [J/(kg.K)]338.6451338.2691336.7734330.9206309.4112
γ m i x [-]1.2439941.2442781.2454181.2500871.270686
a [m/s]984.341983.906982.178975.430950.935
u [m/s]1968.681967.811964.361950.861901.87
Table 10. Electric conductivity [S/m] of hydrogen plasma (potassium seed, air oxidizer).
Table 10. Electric conductivity [S/m] of hydrogen plasma (potassium seed, air oxidizer).
Pressure [atm] Plasma Electric Conductivity [S/m] (at the Different Potassium Levels)
0.0625%0.25%1%4%6%16%
0.06254.92179.344316.034321.620121.982719.7671
0.1253.52826.740511.689515.885616.150414.4434
0.252.51894.83348.445911.542811.734210.4518
0.51.79313.45136.06308.31958.45687.5108
11.27392.45704.33255.96176.05975.3707
20.90371.74563.08594.25464.32433.8271
40.64051.23832.19313.02773.07722.7205
80.45360.87761.55612.15022.18531.9307
160.32110.62151.10291.52501.54991.3686
Table 11. Power density [MW/m3] of hydrogen plasma (potassium seed and air oxidizer).
Table 11. Power density [MW/m3] of hydrogen plasma (potassium seed and air oxidizer).
Pressure [atm] Volumetric Power Density [MW/m3] (at the Different Potassium Levels)
0.0625%0.25%1%4%5%16%
0.0625119.219226.149386.697514.268518.972446.874
0.12585.464163.132281.914377.864381.364326.521
0.2561.016116.977203.689274.564277.128236.283
0.543.43583.528146.221197.892199.751169.796
130.85859.464104.486141.808143.144121.415
221.89042.24774.422101.202102.15886.519
415.51529.96952.89172.01972.69861.502
810.98821.23937.52851.14651.63143.647
167.77815.04126.59836.27536.61830.940
Table 12. Pre-ionization properties of potassium-seeded oxy-hydrogen combustion mixture.
Table 12. Pre-ionization properties of potassium-seeded oxy-hydrogen combustion mixture.
Property Different Pre-Ionization Potassium Mole Fraction ( X K )
0.0625%0.25%1%4%16%
X K [%]0.0625%0.25%1%4%16%
X H 2 O [%]99.9375%99.75%99%96%84%
M m i x [kg/kmol]18.02818.06818.22618.85921.389
R m i x [J/(kg.K)]461.1850460.1760456.1837440.8839388.7337
γ m i x [-]1.1832441.1834901.1844791.1885431.206766
a [m/s]1120.3111119.2011114.8011097.8261038.727
u [m/s]2240.622238.402229.602195.652077.45
Table 13. Electric conductivity [S/m] of hydrogen plasma (potassium seed, oxygen oxidizer).
Table 13. Electric conductivity [S/m] of hydrogen plasma (potassium seed, oxygen oxidizer).
Pressure [atm] Plasma Electric Conductivity [S/m] (at the Different Potassium Levels)
0.0625%0.25%1%4%13%16%
0.06252.38744.67358.675313.915416.253716.1854
0.1251.70273.34086.238510.085711.810011.7558
0.251.21122.38034.46357.25638.51198.4704
0.50.86011.69203.18215.19326.09936.0683
10.60991.20082.26283.70304.35274.3300
20.43220.85131.60632.63353.09743.0809
40.30600.60301.13891.86962.19972.1878
80.21660.42690.80681.32561.56011.5516
160.15330.30210.57120.93911.10541.0993
Table 14. Power density [MW/m3] of hydrogen plasma (potassium seed, oxygen oxidizer).
Table 14. Power density [MW/m3] of hydrogen plasma (potassium seed, oxygen oxidizer).
Pressure [atm] Volumetric Power Density [MW/m3] (at the Different Potassium Levels)
0.0625%0.25%1%4%9%16%
0.062574.910146.352269.538419.279458.894436.583
0.12553.426104.618193.827303.888333.446317.099
0.2538.00474.540138.679218.636240.334228.479
0.526.98852.98598.866156.474172.221163.685
119.13737.60370.304111.573122.907116.797
213.56126.65949.90779.34987.46383.104
49.60118.88335.38556.33262.11659.013
86.79613.36825.06739.94144.05441.853
164.8109.46017.74728.29631.21529.652
Table 15. Pre-ionization properties of cesium-seeded oxy-hydrogen combustion mixture.
Table 15. Pre-ionization properties of cesium-seeded oxy-hydrogen combustion mixture.
Property Different Pre-Ionization Cesium Mole Fraction ( X C s )
0.0625%0.25%1%4%16%
X C s [%]0.0625%0.25%1%4%16%
X H 2 O [%]99.9375%99.75%99%96%84%
M m i x [kg/kmol]18.08718.30319.16422.61136.398
R m i x [J/(kg.K)]459.6900454.2795433.8539367.7191228.4336
γ m i x [-]1.1832421.1834821.1844491.1884191.206172
a [m/s]1118.4931112.0041087.1611002.552796.065
u [m/s]2236.992224.012174.322005.101592.13
Table 16. Electric conductivity [S/m] of hydrogen plasma (cesium seed, oxygen oxidizer).
Table 16. Electric conductivity [S/m] of hydrogen plasma (cesium seed, oxygen oxidizer).
Pressure [atm] Plasma Electric Conductivity [S/m] (at the Different Cesium Levels)
0.0625%0.25%1%4%13%16%
0.06256.987813.550124.428237.762043.671043.5831
0.1255.06339.870018.051428.408433.027132.9308
0.253.64277.127313.171621.007124.518424.4295
0.52.60735.11469.522315.334217.948017.8732
11.85953.65396.838211.087713.003412.9440
21.32272.60234.88767.96299.35149.3060
40.93921.84923.48185.69136.68996.6560
80.66611.31212.47464.05414.76844.7435
160.47190.93001.75592.88123.39013.3721
Table 17. Power density [MW/m3] of hydrogen plasma (cesium seed, oxygen oxidizer).
Table 17. Power density [MW/m3] of hydrogen plasma (cesium seed, oxygen oxidizer).
Pressure [atm] Volumetric Power Density [MW/m3] (at the Different Cesium Levels)
0.0625%0.25%1%4%5%16%
0.0625218.548418.885721.804948.874947.117690.486
0.125158.358305.119533.382713.839714.060521.722
0.25113.928220.332389.194527.861528.886387.036
0.581.545158.112281.365385.314386.524283.165
158.157112.956202.055278.609279.721205.072
241.36880.447144.419200.090201.008147.435
429.37457.166102.880143.010143.725105.451
820.83340.56273.119101.870102.40975.151
1614.75928.75051.88372.39872.79353.424
Table 18. Comparison between our study and three recent studies in a similar field.
Table 18. Comparison between our study and three recent studies in a similar field.
ReferenceComparison with Our Study
[309]The study in this reference is limited to potassium seeding and does not explore the effect of using cesium seeding (which we did).
[310]The study in this reference uses liquid metal as a working fluid for magnetohydrodynamic power generators, not hydrogen (which we used).
[311]The study in this reference describes a specific flow problem and a special channel that is porous. The exact fluid type is not specified, making the study have probable great theoretical value but not necessarily applicable to commercial magnetohydrodynamic power generation (which we considered).
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. Power Density and Thermochemical Properties of Hydrogen Magnetohydrodynamic (H2MHD) Generators at Different Pressures, Seed Types, Seed Levels, and Oxidizers. Hydrogen 2025, 6, 31. https://doi.org/10.3390/hydrogen6020031

AMA Style

Marzouk OA. Power Density and Thermochemical Properties of Hydrogen Magnetohydrodynamic (H2MHD) Generators at Different Pressures, Seed Types, Seed Levels, and Oxidizers. Hydrogen. 2025; 6(2):31. https://doi.org/10.3390/hydrogen6020031

Chicago/Turabian Style

Marzouk, Osama A. 2025. "Power Density and Thermochemical Properties of Hydrogen Magnetohydrodynamic (H2MHD) Generators at Different Pressures, Seed Types, Seed Levels, and Oxidizers" Hydrogen 6, no. 2: 31. https://doi.org/10.3390/hydrogen6020031

APA Style

Marzouk, O. A. (2025). Power Density and Thermochemical Properties of Hydrogen Magnetohydrodynamic (H2MHD) Generators at Different Pressures, Seed Types, Seed Levels, and Oxidizers. Hydrogen, 6(2), 31. https://doi.org/10.3390/hydrogen6020031

Article Metrics

Back to TopTop