Next Article in Journal
Sustainable Behavior among Romanian Students: A Perspective on Electricity Consumption in Households
Next Article in Special Issue
Unit Commitment for Power Generation Systems Based on Prices in Smart Grid Environment Considering Uncertainty
Previous Article in Journal
A Public Bad Game Method to Study Dynamics in Socio-Ecological Systems (Part II): Results of Testing Musa-Game in Rwanda and Adding Emergence and Spatiality to the Analysis
Previous Article in Special Issue
Numerical Evaluation of the Flow around a New Vertical Axis Wind Turbine Concept
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Development and Operation Modes of Hydrogen Fuel Cell Generation System for Remote Consumers’ Power Supply

1
Department of Electric Power Engineering, Power Supply and Power Electronics, Nizhny Novgorod State Technical University n.a. R.E. Alekseev, 603950 Nizhny Novgorod, Russia
2
Sirius University of Science and Technology, 354340 Sochi, Russia
3
Department of Applied Mathematics, Nizhny Novgorod State Technical University n.a. R.E. Alekseev, 603950 Nizhny Novgorod, Russia
4
Department of Electrical Equipment, Electric Drive and Automation, Nizhny Novgorod State Technical University n.a. R.E. Alekseev, 603950 Nizhny Novgorod, Russia
*
Author to whom correspondence should be addressed.
Sustainability 2021, 13(16), 9355; https://doi.org/10.3390/su13169355
Submission received: 30 June 2021 / Revised: 9 August 2021 / Accepted: 16 August 2021 / Published: 20 August 2021
(This article belongs to the Special Issue Design and Optimization of Renewable Energy Systems)

Abstract

:
At the present stage of electric power industry development, special attention is being paid to the development and research of new efficient energy sources. The use of hydrogen fuel cells is promising for remote autonomous power supply systems. The authors of the paper have developed the structure and determined the optimal composition of a hybrid generation system based on hydrogen fuel cells and battery storage and have conducted studies of its operating modes and for remote consumers’ power supply efficiency. A simulation of the electromagnetic processes was carried out to check the operability of the proposed hybrid generation system structure. The simulation results confirmed the operability of the structure under consideration, the calculation of its parameters reliability and the high quality of the output voltage. The electricity cost of a hybrid generation system was estimated according to the LCOE (levelized cost of energy) indicator, its value being 1.17 USD/kWh. The factors influencing the electricity cost of a hydrogen generation system have been determined and ways for reducing its cost identified.

1. Introduction

The global trends in the modern electric power industry are decarbonization and decentralization aimed at improving the environmental friendliness, reliability and quality of power supply to consumers [1]. Hydrogen is an important energy carrier of the future because it can be produced from a variety of renewable and nonrenewable resources including in remote places without electricity infrastructure [2]. Fuel cells (FC) are the most promising type of power plant for power production from hydrogen. The key advantages of fuel cells are higher energy efficiency due to reliability (no moving parts), absence of energy conversion intermediate stages (electrical efficiency up to 60%), harmful emissions and noise [3].
Hydrogen fuel cells are already widely used in the transport industry, space industry, etc. [4,5]. However, expanding the range of applications for hydrogen fuel cell power plants is essential to accelerate technology development and reduce production costs.
The paper is devoted to the development of a hydrogen fuel cell generation system (HFCGS) for power supply to remote low-power consumers (up to 3 kW). A generation system of such capacity may be needed to supply:
-
consumers along the oil and gas pipelines (for example: electrochemical protection, linear telemechanics, radio relay communication equipment, cleaning device launchers);
-
stationary railway objects (communication devices, automatic crossing signaling);
-
autonomous meteorological and telecommunication stations.
Diesel power plants are now widely used for such systems; however, diesel power plants have a number of disadvantages, such as low efficiency and reliability, the need for frequent maintenance, environmental pollution, noise and vibration.
Hydrogen fuel cells eliminate these problems, and they are promising power sources for remote low-power consumers. Meanwhile, fuel cells have disadvantages such as slow dynamic response, long start-up time and soft discharge characteristic [6]. These problems can be solved by combining fuel cells with a storage system and electricity converters in a hybrid generation system.
Many studies are devoted to the development of mathematical and simulation model or prototypes of hybrid systems based on fuel cells, storage systems and other types of energy sources (usually renewable).
Experimental studies of hybrid systems based on fuel cells and solar photovoltaic panels (PV) with a power of 500 W and 200 W are presented in [7] and [8], respectively. The electricity generated by PV is used to power a hydrogen generator. The joint operation of PV and fuel cells is not considered. The concept and prototype of a hybrid system that integrates PV, battery storage and hydrogen fuel cells for power supply of remote telecommunications is described in [9]. This system powers a 12-V load with a power of just under 10 W during the day. The PV panel is the main source while FCs are only required for relatively short timescales to charge the battery or when inclement weather and PV power is not enough. This study shows that FCs can be a useful alternative to PV or wind in case of the small loads of remote consumers.
One of the areas of research is devoted to improving the structure and development of control systems for fuel cells hybrid systems. The control system of a PV-Fuel cell hybrid system is presented in [10]. In [11], the control system of a PV-Wind-Fuel cell hybrid system was developed. The parameter coordination of renewable energy sources and fuel cells is carried out using a DC bus. Additional DC/DC and AC/DC converters are used to connect the sources to the DC bus. Fuel cells are considered as a backup power source to compensate for fluctuations in the power of PV and wind turbines.
In [12], a single-inductor dual-input single-output boost converter was used to coordinate the sources parameters of stand-alone hybrid PV-Fuel cell system. Compared to using a DC bus, this reduces the number of components and the size of the hybrid system, but it requires a high computational performance of the controller. A hybrid system based on a proton exchange membrane (PEM) fuel cell and storage batteries is presented in [13]. Two DC/DC converters are used to match PEM and battery parameters.
This paper represents a continuation of the research in [14] and presents a hydrogen fuel cell generation system (HFCGS) based on hydrogen fuel cells and battery storage. The HFCGS is designed to power remote consumers with a changing daily load schedule.
The study goal is to develop the structure and select the optimal composition of an HFCGS that provides a minimum number of converters and the size of the system and to study its operation modes and the efficiency of its use as a power supply for remote consumers.

2. Analysis and Selection of the Hydrogen Fuel Cell Generation System Main Units

2.1. Fuel Cells

The key element in the structure of the HFCGS is the block of fuel cells, which largely determines the efficiency and reliability of consumer’s power supply. When developing the structure and determining the composition of the HFCGS, the most important task is to select the most suitable fuel cells type.
Fuel cells are classified according to the type of electrolyte, operating temperature and fuel. Considering these parameters, five main types of fuel cells that use hydrogen as a fuel can be singled out:
  • fuel cells with a polymer or proton exchange membrane—PEMFC;
  • fuel cells with alkaline electrolyte—AFC;
  • fuel cells with phosphate electrolyte—PAFC;
  • fuel cells with molten carbonate—MCFC;
  • fuel cells with solid oxide electrolyte—SOFC.
The main structural features of the fuel cell types are shown in Table 1 [15,16,17,18,19].
PEMFCs operate within 60–120 °C temperature range and are classified as low-temperature fuel cells. PEMFC are characterized by high efficiency of electricity generation (efficiency up to 60%), with a service life of 3000 to 5000 h [20]. The important advantages of PEMFCs are their high power density, a short start and stop time, a high load change rate within a wide range, ease of operation and the ability to work in a wide range of ambient temperatures [21]. The main disadvantage of PEMFCs is the high requirements for the hydrogen quality (catalysts are susceptible to CO poisoning due to low temperatures) [22]. Therefore, additional purification of hydrogen may be required when using PEMFCs.
AFC as well as PEMFC refer to low-temperature fuel cells (operating temperature is 50–100 °C) [23]. AFC technologies are the most studied, therefore their power generation efficiency reaches 60% and the service life is estimated at 10,000 to 15,000 h. The main advantages of AFCs are similar to PEMFCs. The disadvantage of AFCs is their high sensitivity to CO2, which reacts with the electrolyte and reduces the efficiency of the fuel cell. CO2 can be contained in the air, so the use of AFCs is limited to closed spaces (space environment, submarines).
PAFCs operate at temperatures from 150 to 220 °C and are classified as medium-temperature fuel cells. The main advantages of PAFCs are the relative simplicity of their structure, their low electrolyte volatility and their high stability. As a result, PAFC service life is estimated at about 65,000 h. PAFC can use impure H2 as fuel and can tolerate up to 1.5% CO at operating temp. However, the efficiency of PAFC is about 40% and is inferior to other types of fuel cells [23].
MCFCs are high-temperature fuel cells that operate at temperatures from 600 to 800 °C. MCFCs have a fairly high efficiency (the efficiency of electricity generation is 50–60%). MCFCs are less demanding on the purity of hydrogen due to the high operating temperature (catalysts are not susceptible to CO and CO2 poisoning). The main disadvantages of MCFCs are a short service life (7000–8000 h) due to the destruction of components and corrosion, a slow start-up time and the impossibility of restarting after a stop.
SOFCs are fuel cells with the highest operating temperature from 650 to 1000 °C. SOFCs are characterized by high efficiency of electric power generation (efficiency is about 60%). The service life of SOFCs is estimated at over 20,000 h due to low material degradation. High temperatures also allow fuel to be used without special pretreatment and generate combined electric and thermal power. However, SOFCs are characterized by the significant amount of time to reach the optimal operating mode and the impossibility of changing the power generation due to high operating temperatures. This significantly complicates the use of SOFCs in power supply systems.
The results of a comparative analysis of hydrogen fuel cells are shown in Figure 1.
The type of fuel cell most suitable for use in an HFCGS is largely determined by the consumer power.
Thus, power supply sources for medium-power consumers (from 10 to 100 kW) must provide high energy efficiency and the minimum cost of electricity generated. The priority criteria for selecting a particular type of fuel cell as part of hybrid generation systems are:
  • high electricity generation efficiency;
  • low requirements for fuel quality (to reduce the cost of special preparation and large fuel volumes processing);
  • the possibility of additional use of the heat released by the fuel cell (to increase the overall source efficiency).
Meanwhile, the daily load schedule of medium-power consumers usually changes during the day, but the consumption does not go down beyond a certain level. This makes it possible to continuously stabilize the operating mode of the fuel cell (by using an energy storage system).
Thereafter, it is promising to use high-temperature fuel cells in a hybrid energy complex to supply medium-power consumers (from 10 to 100 kW), which allow the use of fuel without special pretreatment and provide combined production of electrical and thermal energy. SOFCs are considered the most efficient in terms of energy performance and service life among high-temperature fuel cells.
Low-power consumers (from 1 to 10 kW), in most cases, are characterized by a changing daily load schedule ranging from a decrease in consumption to zero and sharp peaks. Energy sources for such consumers, first of all, must ensure high stability of power supply and duration of autonomous operation. Therefore, the decisive criteria for selecting a fuel cell are a short start and stop time and the ability to quickly change generation over a wide range (for generation schedule to electrical load schedule maximum approximation).
These requirements are met by low-temperature fuel cells—PEMFCs and AFCs. They operate on high quality hydrogen fuel. At the same time, the use of AFCs (due to its high sensitivity to CO2) is limited to closed spaces (submarines, spacecraft). PEMFCs can operate in a wide temperature range without significant restrictions to environmental conditions (vehicles, low-power stationary consumers, etc.).
The developed HFCGS is designed to power a 3-kW load. Therefore, the most applicable FCs are PEMFCs, which are characterized by a high load change rate, ease of operation and the ability to work in a wide range of ambient temperatures.

2.2. Battery Storage

To increase the duration of the HFCGS autonomous operation without refueling, it seems relevant to use a combination of fuel cells with battery storage (BS). BS will ensure the operation of the fuel cells in the most efficient modes at a changing load schedule.
Lithium batteries are the most common. They are characterized by a high energy density and a minimum self-discharge rate, which significantly increases the storage period and extends the service life without deteriorating functional performance. BS must be equipped with a protective controller (BMS board) to prevent the complete loss of the battery cells potential. Lithium batteries are widely used in technology and consumer electronics, electric transport, and also serve as an uninterruptible power supply source.
The main element of any BS is a battery cell. The characteristics of the BS (capacity, reliability and functionality) depend on the parameters of the battery cell. Several types of these can be singled out, depending on the cathodes and electrolytes (Table 2). Lithium polymer batteries were excluded from this comparison as they are fire hazardous when overcharged and overheated.
Figure 2 shows the dependence of gravimetric energy density on the capacity value for various types of battery cells.
Table 2 shows that LTO battery cells are the optimal type of BS for use in stand-alone power supply systems to remote consumers. They have the highest service life (up to 20,000 cycles), allow for complete battery discharge and can operate in subzero temperatures. However, their high unit cost and low energy intensity (Figure 2) hinder their widespread distribution.
The optimal type of cells for creating an energy storage system is LiFePO4 with an HFCGS power up to 100 kW. LiFePO4 cells are available in a wide range of capacities (from 1.1 to 280 Ah), have a high efficiency (up to 95%) and a service life of more than 2000 cycles (up to 8000 cycles at standard charge–discharge currents). Their main advantages are complete safety (they have thermal and chemical stability) and a long service life. Therefore, to create the HFCGS storage system, the cylindrical cells LiFePO4 3.2 V PLB IFR26650-35A, 3500 mAh were selected [24].
In the power range from 100 to 1000 kW, LTO is preferable due to the high values of the charge–discharge current (10C) and the high charge rate. However, in order to increase the current, it will be necessary to increase the total capacity of the batteries using parallel LTO cell connections.

3. Materials and Methods

Operability and efficiency of the HFCGS depend on the optimal choice of its structural and circuit solutions.
The functional diagram of the developed HFCGS power circuits is shown in Figure 3. The proposed structure provides uninterrupted power supply of the combined resistive and inductive load ZL from the HFCGS or from the AC mains (if available). Switching the power supply from the HFCGS to the centralized electrical network is carried out by high-speed solid-state relays KV1–KV4 controlled by a microcontroller.
The HFCGS functional diagram contains fuel cells (FC) and battery storage (BS) based on LiFePO4 batteries. The battery functions as an electric buffer, providing together with the fuel cells the required output power level for a given time. The power circuit also contains a converter (C) and an output autonomous voltage inverter (AVI) with PWM.
The converter matches the relatively low FC voltage level with the increased AVI voltage and also realizes the galvanic isolation of the fuel cells from the remaining power circuit. It is made based on AVI with a latitude regulation of the output voltage and a transformer output.
The output AVI is made according to a single-phase bridge circuit and operates in unipolar sinusoidal PWM mode, making it possible to reduce the required power consumption of its output filter (in Figure 3, this filter and the EMC filter are included in the AVI structure). The HFCGS converters are based on MOSFETs and IGBTs.
In the developed device, the batteries are connected in series to obtain the voltage that provides the required level of AVI power supply voltage and HFCGS preset output voltage (220 V), respectively.
“Converter—fuel cell” power circuit section has a feedback system on the output FC power, so on the interval of standby power supply load FC operates in the mode of a constant rated power source.
HFCGS power circuit also contains filter capacitors C1 and C2, which block the AC components generated by the converters.
The main parameters of the HFCGS key units are shown in Table 3. The parameters of each device elements were calculated to develop a HFCGS simulation model.

3.1. Calculation of Fuel Cells Internal Resistance

The discharge characteristic of the selected 48 V FC with a power of 1 kW is shown in Figure 4.
FC internal resistance rFC according to its discharge characteristic is:
r F C = Δ U Δ I = 70 57 18.2 3.5 = 13 14.7 = 0.88 Ohm

3.2. Battery Storage Parameters Calculation

To create a HFCGS battery storage system, a series assembly of 12 LiFePO4 batteries 36 V 3.5 Ah (configuration 12S1P) with BMS boards is used. Table 4 shows the parameters of the battery storage system.
Figure 5 shows the discharge characteristics of the selected batteries.

3.3. Filter Capacitance Values Calculation

To determine the capacitances, we used the equations describing time dependence on the averaged current consumed by a single-phase bridge AVI with sinusoidal PWM:
i ˜ 1 = 1 2 μ I m · [ cos φ L cos ( 2 ω t φ L ) ] ,
where Im is amplitude of the averaged output current of a single-phase bridge AVI, μ is output voltage modulation depth, φL is load current phase angle and ω is mains voltage angular frequency.
The maximum value of the amplitude of the variable component of the averaged input current with a circular angular of 2ω (f = 100 Hz) is:
I m max = 1 2 μ max I m
The variable component of the averaged input current, i.e., the current generated by the AVI into the power circuit is closed through the battery, FC and filter capacitors C1 and C2, which causes the pulsations of AVI supply voltage and, as a consequence, distortion of HFCGS output voltage.
Maximum AVI voltage ripple amplitude is determined by the equation:
U m max = 1 2 μ max I m · r F C r B S 1 + 4 ( r F C r B S · C 1 C 2 · ω ) 2 ,
where rFC is the FCs’ internal resistance reduced to AVI source voltage; rBS is battery storage internal resistance; C1 is C1 capacitance value reduced to AVI source voltage; C2 is C2 capacitance; rFCrBS and C1C2 are total resistance and capacitance of parallel-connected elements, respectively.
For real values of the FC and battery parameters, the following relations are valid:
r F C > > r B S ; C 1 < < C 2 ,
And the Um max calculated equation takes on the form:
U m max 1 2 μ max I m · r B S max 1 + 4 ( r B S max C 2 ω ) 2 .
The ripple factor of AVI supply voltage at 100 Hz frequency is determined by the equation:
k U = U m max U BSmin
By substituting (7) into (6), we obtain an equation for calculating the required value of capacitor C2 providing a given value of the ripple factor:
C 2 = 1 2 r BSmax ω 1 4 ( μ max I m r BSmax k U U BSmin ) 2 1 .
With AVI voltage ripple factor kU = 0.045:
C 2 = 1 2 · 2.6 · 314 1 4 ( 0.95 · 24 · 2.6 0.045 · 360 ) 2 1 = 940 μ F .
The capacitance value of C2 = 1000 μF is accepted.
According to (5) an insignificant part of the variable current component generated by AVI into the power circuit is closed through the FC and C1 filter capacitor, and its value has no significant impact on HFCGS output voltage. The required value of C1 capacitance can be selected from the following approximate ratio:
C 1 r F C r B S max · C 2 .
Value of C1 capacitance:
C 1 r F C r B S max · C 2 = 0.88 2.6 · 1000 = 318 μ F .
The values of the half-bridge AVI capacitances are taken slightly lower than the calculated ones and equal to C = 2·C1 = 470 μF.

3.4. AVI Output Voltage Modulation Depth Calculation

AVI output voltage with PWM modulation depth is determined by the equation:
μ = U m . o u t U B S ,
where Um.out is the amplitude of the HFCGS output voltage.
For a given value of the output voltage, the maximum required modulation depth μmax calculated value is:
μ max = U m . o u t U BSmin r B S max · P o u t U BSmin ,
where Pout is the power supplied by the HFCGS to the load.
At the effective output voltage value Uout = 220 V and output power Pout = 3 kW, the maximum required calculated μmax value according to (13) is:
μ max = 2 · 220 360 2.6 · 3 · 10 3 360 = 0.92 .
Taking into account voltage drop across the open transistors and dead time of switching AVI transistors, the value of μmax = 0.95 is taken.

3.5. Converter Output Voltage Boost Coefficient Calculation

This converter is made on the based on the AVI with a high frequency transformer and matches relatively low fuel cell output voltage to battery voltage. Output voltage boost coefficient kb is calculated in accordance with the equation:
k b = D max k 21 = U B S max U F C min ,
where Dmax is the maximum relative duration of switching on converter transistors and k21 is the transformation ratio of converter matching transformer.
Coefficient kb:
k b = D max k 21 = U B S max U F C min = 518.4 57 = 9.1 .
Setting the Dmax value, the required k21 transformation ratio for the AVI bridge circuit is:
k 21 = k b D max .
When using a half-bridge AVI in the converter, the transformation ratio value is calculated by the equation:
k 21 = 2 k b D max ;
k 21 = 2 k b D max = 2 · 9.1 0.8 = 22.8 .

4. Simulation Modeling of Hydrogen Fuel Cell Generation System Operation Modes

An electromagnetic processes simulation modeling was carried out in MATLAB Simulink in order to test the operability of the proposed structure of the HFCGS.
The initial data for the modeling and the parameters of the HFCGS power circuit elements are given in Table 5.
The autonomous inverter and converter simulation model structures are shown in Figure 6 and Figure 7, respectively.

5. Results and Discussion

The simulation results are presented by the timing diagrams of the autonomous inverter (Figure 8) and the voltage converter (Figure 9, Figure 10 and Figure 11).
Figure 10 shows the same converter operation curves as in Figure 9, but for a 0.5 ms time interval.
In addition to analyzing the instantaneous values of the current and voltages of the HFCGS’s major components, an analysis of the harmonic components of the autonomous inverter output voltage has been carried out (Figure 12 and Figure 13).
Figure 12 and Figure 13 show that HFCGS output voltage contains a triple frequency component of 150 Hz in addition to the first frequency harmonic 50 Hz due to the AVI supply voltage amplitude modulation. The third harmonic amplitude at the selected values of C1 and C2 capacitances does not exceed 1–2% of the first harmonic.
The simulation results confirmed the considered HFCGS structure’s operability, the reliability of its parameter calculations and the high quality of its output voltage.

6. Cost Estimation of Electric Power Generated by Hydrogen Fuel Cell Generation System

The levelized cost of energy (LCOE) was calculated to analyze the HFCGS’s competitiveness as follows:
LCOE = t = 1 n I t + M t + F t + C a r b + U t i l ( 1 + r ) t t = 1 n E t ( 1 + r ) t ,
where I t is the annual investment (capital) costs, USD; M t is the annual operating costs and maintenance costs, USD; F t is the annual fuel costs, USD; E t is the annual quantity of generated electric energy, kWh; C a r b is the annual cost of greenhouse gas emission (carbon footprint); U t i l is the decommissioning charge after the HFCGS end-of-life; r is the discount rate; and t = 1, …, n is the service life.
The following factors affect the cost per unit of electrical energy generated by a HFCGS:
Hydrogen cost. The price of hydrogen may vary depending on the country of origin and the prices of fossil fuels in it, on the method of producing hydrogen (steam reforming of hydrocarbons, gasification of solid fuels, thermochemical decomposition of water using the energy of a high-temperature gas-cooled reactor (HTGR), electrolysis of water), its transportation (by pipeline or in cylinders) and its state (compressed or liquefied). The methods to produce hydrogen differ in the amount of costs for electricity, H2O, natural gas, etc. In addition, each method of producing H2 is characterized by a different amount of CO2 emissions;
The power of the hydrogen plant, its duty factor and its fuel consumption;
Cost of greenhouse gas emissions and hydrogen plant utilization after the end of its life.
Table 6 provides a comparative analysis of various methods for producing hydrogen from the standpoint of specific energy consumption and carbon dioxide emissions.
To obtain liquefied hydrogen, it is necessary to spend three to seven times more electricity than to obtain compressed gaseous hydrogen. This also increases CO2 emissions.
Table 7 provides an estimate of the energy resources cost, the specific investments for hydrogen production and the cost of transporting hydrogen fuel.
The change in the cost of compressed hydrogen production for various technologies depending on the volume of hydrogen production is shown in Figure 14 and Figure 15 [25,29,30].
Thus, regardless of the volume of hydrogen production, the cheapest method is the method of steam methane reformation (SMR). The cost of hydrogen starts from 3 USD/kg H2 at low capacities of SMR units, but with a capacity of more than 500 tons per day, it decreases to 1.7 USD/kg H2. Hydrogen production process by coal gasification at low productivity is strongly inferior to SMR, although at high productivity the coal option is practically equal to the SMR method. The method of thermochemical decomposition of water based on HTGR is applicable only at capacities of more than 35–50 tons H2 per day. In this case, the cost of hydrogen turns out to be at least 3 USD/kg H2. The hydrogen production based on electrolysis is significantly inferior in cost indicators to the methods discussed above.
The cost of electrolytic hydrogen turns out to be much higher: from 7–8 USD/kg H2 when operating on electric energy from the power system to 9–13 USD/kg H2 when receiving electricity from wind power plants and up to 35–50 USD/kg H2 when using photovoltaic converters as electricity source.
The specific costs of hydrogen transportation is shown in Figure 16. The liquefaction of hydrogen allows for reducing transportation costs by two to five times as compared to the option of compressed hydrogen. Likewise, hydrogen production and the cost of transport will depend on the pace of technology development and cost reduction [31].
Table 8 shows hydrogen cost estimation results obtained by Russian scientists compared to data from foreign publications. The total hydrogen cost values calculated by foreign researchers appears more optimistic for most technologies, as shown by Table 8 [25,27,31].
The initial data for calculating electric energy cost from the HFCGS are given in Table 9.
The investment costs consist of the HFCGS components cost (fuel cell, batteries, converters, hydrogen generator) and the cost of installation and commissioning. A hydrogen generator based on electrolysis with a 12 L per minute capacity serves as a source of hydrogen in the HFCGS experimental sample. An average of 1000 L of H2 is produced from 1 L of distilled water according to its technical parameters.
The operating and maintenance costs of HFCGS include the cost of source services.
The costs of paying for greenhouse gas emissions and the costs of HFCGS decommissioning after the end of service life are taken as equal to 0.
LCOE = 1.17 USD/kWh is calculated substituting the values from Table 9 into Equation (20).
Figure 17 shows the LCOE for diesel, gasoline and hydrogen power plants of the equal capacity.

7. Conclusions

The paper presents results of the structure development and choice of the optimal composition of a hydrogen generation system based on fuel cells and an accumulation system, as well as the research results of its operation modes.
The structural and functional diagrams of the hydrogen fuel cell generation system have been developed. One of the main elements of the HFCGS is fuel cells, the PEMFC type being the most suitable ones. They are characterized by a high rate of load change, ease of operation and the ability to work in a wide range of ambient temperatures. A sequential assembly of 12 LiFePO4 36 V 3.5 Ah batteries is defined as a buffer storage system for HFCGS, providing the required level of output power together with the fuel cell. The HFCGS power circuit also includes a converter and an output autonomous voltage inverter with PWM. The proposed scheme eliminates the need to install an additional converter for matching batteries and HFCGS voltages, which significantly simplifies the HFCGS, stabilizes its operation and improves the weight-size parameters.
The parameters calculation of the device elements has been performed—filter capacitors, converter, autonomous voltage inverter, fuel cells and batteries. The HFCGS simulation model was built to test the operability of the considered structure using the obtained parameters. The operation diagrams of the output autonomous voltage inverter, converter, as well as the spectral analysis of HFCGS output voltage curve have been constructed. It was found that the HFCGS output voltage contains the third harmonic and its value does not exceed 1–2% of the amplitude of the fundamental harmonic. The simulation results confirmed the operability of the considered HFCGS structure, the reliability of its parameter calculations and a high quality of output voltage.
The cost estimation of the electric power generation by the hybrid energy complex has been carried out. It exceeds the electric energy cost from power plants running on gasoline and diesel. It has been established that the electric energy cost is influenced by both the cost of energy source itself and the hydrogen cost. This allows us to identify two important directions in reducing the electric energy cost—improving the elements that make up a hydrogen source (fuel cells, converters, batteries, etc.) and development methods of more efficient ways of hydrogen production.

Author Contributions

Conceptualization, A.K. (Aleksandr Kulikov) and A.L.; methodology, A.K. (Andrey Kurkin) and A.K. (Andrey Kozelkov); software, V.V.; validation, A.S. (Andrey Shahov), A.S. (Andrey Shalukho), R.B., A.K. (Andrey Kozelkov) and E.K.; formal analysis, A.S. (Andrey Shalukho) and E.K.; investigation, V.V.; resources, I.L.; data curation, R.B., A.D. and I.L.; writing—original draft preparation, A.K. (Aleksandr Kulikov), A.L., A.S. (Andrey Shalukho), E.K. and I.L.; writing—review and editing, A.L.; visualization, R.B.; supervision, A.L. and A.D.; project administration, A.K. (Aleksandr Kulikov) and A.L.; funding acquisition, A.K. (Aleksandr Kulikov) and A.K. (Andrey Kurkin). All authors have read and agreed to the published version of the manuscript.

Funding

The reported study was funded by RFBR, Sirius University of Science and Technology, JSC Russian Railways and Educational Fund “Talent and success”, project number 20-38-51016 and Council of the grants of President of the Russian Federation for the state support of Leading Scientific Schools of the Russian Federation (Grant No. NSH-2485.2020.5).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

There are no publicly available data that should be provided.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Voropay, N.I. Prospects and problems of electric power system transformations. Electrichestvo 2020, 7, 12–21. [Google Scholar] [CrossRef]
  2. Pethaiah, S.S.; Sadasivuni, K.K.; Jayakumar, A.; Ponnamma, D.; Tiwary, C.S.; Sasikumar, G. Methanol electrolysis for hydrogen production using polymer electrolyte membrane: A mini-review. Energies 2020, 13, 5879. [Google Scholar] [CrossRef]
  3. Sosnina, E.; Shalukho, A.; Veselov, L. Application of SOFCs on biogas in power supply systems for agricultural enterprises. Smart Electr. Eng. 2020, 4, 27–41. [Google Scholar] [CrossRef]
  4. Hemmes, K. A personal retrospect on three decades of high temperature fuel cell research; ideas and lessons learned. Int. J. Hydrogen Energy 2021, 46, 14962–14976. [Google Scholar] [CrossRef]
  5. Zhou, S.; Cui, Q.; Zhang, M.; Xia, G.; Wang, K. Study on the management of fuel cell vehicle energy system using hybrid fuzzy logic controller. Power Gener. Technol. 2018, 39, 554–560. [Google Scholar] [CrossRef]
  6. Li, G.; Chen, J.; Zheng, X.; Xiao, C.; Zhou, S. Research on energy management strategy of hydrogen fuel cell vehicles. In Proceedings of the 2020 Chinese Automation Congress (CAC), Shanghai, China, 6–8 November 2020; pp. 7604–7607. [Google Scholar] [CrossRef]
  7. Choi, H.-J.; Park, S.-J.; Choi, J.-S.; Cha, I.-S.; Yoon, J.-P.; Suh, J.-S.; Gun, S.-D. An analysis of PEMFC & photovoltaic 500W hybrid system. In Proceedings of the 7th International Conference on Power Electronics, Daegu, Korea, 22–26 October 2007; pp. 522–524. [Google Scholar] [CrossRef]
  8. Khurshid, O.; Saher, S.; Qamar, A. Power generation by hybrid approach solar PV/battery power/hydrogen generation/fuel cell. In Proceedings of the 2019 International Conference on Electrical, Communication, and Computer Engineering (ICECCE), Swat, Pakistan, 24–25 July 2019; pp. 1–4. [Google Scholar] [CrossRef]
  9. Wilson, P.R.; Wilcock, R. Hybrid hydrogen fuel cell and photo-voltaic system for remote telecommunications applications. In Proceedings of the 2013 4th IEEE International Symposium on Power Electronics for Distributed Generation Systems (PEDG), Rogers, AR, USA, 8–11 July 2013; pp. 1–4. [Google Scholar] [CrossRef] [Green Version]
  10. Taoufikl, M.; Lassad, S. Hybrid photovoltaic-fuel cell system with storage device control. In Proceedings of the 2017 International Conference on Green Energy Conversion Systems (GECS), Hammamet, Tunisia, 23–25 March 2017; pp. 1–6. [Google Scholar] [CrossRef]
  11. Rao, T.E.; Elango, S.; Swamy, G.G. Power management strategy between PV-Wind fuel hybrid system. In Proceedings of the 2021 7th International Conference on Electrical Energy Systems (ICEES), Chennai, India, 11–13 February 2021; pp. 101–107. [Google Scholar] [CrossRef]
  12. Wang, B.; Xian, L.; Manandhar, U.; Ye, J.; Ukil, A.; Gooi, H.B. A stand-alone hybrid PV-fuel cell power system using single-inductor dual-input single-output boost converter with model predictive control. In Proceedings of the 2017 Asian Conference on Energy, Power and Transportation Electrification (ACEPT), Singapore, 24–26 October 2017; pp. 1–5. [Google Scholar] [CrossRef]
  13. Wang, Y.-X.; Ou, K.; Qin, F.-F.; Kim, Y.-B. Proton exchange membrane fuel cell protection control for its hybrid power system application. In Proceedings of the 2015 IEEE PES Asia-Pacific Power and Energy Engineering Conference (APPEEC), Brisbane, Australia, 15–18 November 2015; pp. 1–5. [Google Scholar] [CrossRef]
  14. Loskutov, A.; Sosnina, E.; Chivenkov, A.; Kryukov, E. The development of hybrid power source based on SOFC for distant electricity consumers power supply. In Proceedings of the 2015 IEEE Innovative Smart Grid Technologies—Asia (ISGT ASIA), Bangkok, Thailand, 3–6 November 2015; pp. 1–6. [Google Scholar] [CrossRef]
  15. Filippov, S.; Golodnitsky, A.; Kashin, A. Fuel cells and hydrogen energy. Energy Policy 2020, 11, 28–39. [Google Scholar] [CrossRef]
  16. Faizan, M.; Ali, S.; Ahmad, D.A. An overview of fuel cell based distribution generation integration. In Proceedings of the 2019 International Conference on Power Electronics, Control and Automation (ICPECA), New Delhi, India, 16–17 November 2019; pp. 1–6. [Google Scholar] [CrossRef]
  17. Giorgi, L.; Leccese, F. Fuel cells: Technologies and applications. Open Fuel Cells J. 2013, 6, 1–20. [Google Scholar] [CrossRef] [Green Version]
  18. Jayakumar, A. An assessment on polymer electrolyte membrane fuel cell stack components. In Applied Physical Chemistry with Multidisciplinary Approaches, 1st ed.; Haghi, A.K., Balkose, A., Thomas, S., Eds.; Apple Academic Press: Boca Raton, FL, USA, 2018; pp. 23–49. [Google Scholar] [CrossRef]
  19. Jayakumar, A. A comprehensive assessment on the durability of gas diffusion electrode materials in PEM fuel cell stack. Front. Energy 2019, 13, 325–338. [Google Scholar] [CrossRef]
  20. Chaudhary, S.; Chauhan, Y.K. Studies and performance investigations on fuel cells. In Proceedings of the 2014 International Conference on Advances in Engineering & Technology Research (ICAETR—2014), Unnao, India, 1–2 August 2014; pp. 1–6. [Google Scholar] [CrossRef]
  21. Lee, S.C.; Kwon, O.; Lee, D. Fuel cell simulation: Steady-state and dynamic case. In Proceedings of the 2012 7th International Conference on Computer Science & Education (ICCSE), Melbourne, Australia, 14–17 July 2012; pp. 974–979. [Google Scholar] [CrossRef]
  22. Sohn, S.; Oh, J.; Lee, Y.; Park, D.; Oh, I. Design of a fuel-cell-powered catamaran-type unmanned surface vehicle. IEEE J. Ocean. Eng. 2015, 40, 388–396. [Google Scholar] [CrossRef]
  23. Guaitolini, S.V.M.; Yahyaoui, I.; Fardin, J.F.; Encarnação, L.F.; Tadeo, F. A review of fuel cell and energy cogeneration technologies. In Proceedings of the 2018 9th International Renewable Energy Congress (IREC), Hammamet, Tunisia, 20–22 March 2018; pp. 1–6. [Google Scholar] [CrossRef]
  24. Dongguan Power Long Battery Technology Co, Ltd. (PLB). Available online: http://www.plb-battery.com/battery-cells/3-2v-3500mah-lifepo4-lfp-battery-cell-for.html (accessed on 2 June 2021).
  25. Sinyak, Y.V. Modeling the cost of Hydrogen fuel in the conditions of its centralized production. In Hydrogen Energy Technologies: Materials of the Seminar of the Laboratory of Hydrogen Energy Technologies of the Joint Institute for High Temperatures of the Russian Academy of Sciences; Leontiev, A.I., Dunikov, D.O., Eds.; Federal State Budgetary Institution of Science Joint Institute for High Temperatures of the Russian Academy of Sciences (JIHT RAS): Moscow, Russia, 2017; Volume 1, pp. 39–56. [Google Scholar]
  26. Mitrova, T.; Melnikov, Y.; Chugunov, D.; Glagoleva, A. The Hydrogen Economy—A Path towards Low Carbon Development; SKOLKOVO Energy Centre, Moscow School of Management SKOLKOVO: Moscow, Russia, 2019. [Google Scholar]
  27. Christensen, A. Assessment of Hydrogen Production Costs from Electrolysis: United States and Europe; International Council on Clean Transportation: Berlin, Germany, 2020. [Google Scholar]
  28. Glenk, G.; Reichelstein, S. Economics of converting renewable power to hydrogen. Nat. Energy 2019, 4, 216–222. [Google Scholar] [CrossRef]
  29. Albrecht, U.; Barth, F.; Bünger, U.; Fraile, D.; Lanoix, J.-C.; Pschorr-Schoberer, E.; Vanhoudt, W.; Weindorf, W.; Zerta, M.; Zittel, W. Study on Hydrogen from Renewable Resources in the EU; Ludwig-Bölkow-Systemtechnik GmbH (LBST): Munich, Germany, 2015. [Google Scholar]
  30. Grube, T.; Höhlein, B. Costs of making hydrogen available in supply systems based on renewables. In Hydrogen and Fuel Cell; Töpler, J., Lehmann, J., Eds.; Springer: Berlin/Heidelberg, Germany, 2016. [Google Scholar] [CrossRef]
  31. ACIL Allen Consulting for ARENA. Opportunities for Australia from Hydrogen Exports. Available online: https://acilallen.com.au/projects/energy/opportunities-for-australia-from-hydrogen-exports (accessed on 2 June 2021).
Figure 1. Comparative analysis of hydrogen fuel cells.
Figure 1. Comparative analysis of hydrogen fuel cells.
Sustainability 13 09355 g001
Figure 2. The diagram of the gravimetric energy density dependence on the capacity value for various types of battery cells.
Figure 2. The diagram of the gravimetric energy density dependence on the capacity value for various types of battery cells.
Sustainability 13 09355 g002
Figure 3. Functional diagram of the HFCGS power circuits.
Figure 3. Functional diagram of the HFCGS power circuits.
Sustainability 13 09355 g003
Figure 4. Fuel cell discharge characteristic.
Figure 4. Fuel cell discharge characteristic.
Sustainability 13 09355 g004
Figure 5. Discharge characteristics of 3.2 V PLB IFR26650-35A batteries: (a) for various values of ranging from 0.5C to 10.0C; (b) when discharged with a constant current of 0.5C at various ambient temperatures.
Figure 5. Discharge characteristics of 3.2 V PLB IFR26650-35A batteries: (a) for various values of ranging from 0.5C to 10.0C; (b) when discharged with a constant current of 0.5C at various ambient temperatures.
Sustainability 13 09355 g005
Figure 6. Autonomous inverter simulation model.
Figure 6. Autonomous inverter simulation model.
Sustainability 13 09355 g006
Figure 7. Converter simulation model.
Figure 7. Converter simulation model.
Sustainability 13 09355 g007
Figure 8. Output AVI operation diagrams: (a) AVI output voltage; (b) Battery current; (c) HFCGS output voltage; (d) AVI output current.
Figure 8. Output AVI operation diagrams: (a) AVI output voltage; (b) Battery current; (c) HFCGS output voltage; (d) AVI output current.
Sustainability 13 09355 g008
Figure 9. Converter operation diagrams: (a,b) Converter output current and voltage; (c) FC output current; (d) Battery voltage.
Figure 9. Converter operation diagrams: (a,b) Converter output current and voltage; (c) FC output current; (d) Battery voltage.
Sustainability 13 09355 g009
Figure 10. Converter operation diagrams: (a,b) Converter output current and voltage; (c) FC output current; (d) Battery voltage.
Figure 10. Converter operation diagrams: (a,b) Converter output current and voltage; (c) FC output current; (d) Battery voltage.
Sustainability 13 09355 g010
Figure 11. Converter operation diagrams: (a) Converter output power; (b) Battery voltage.
Figure 11. Converter operation diagrams: (a) Converter output power; (b) Battery voltage.
Sustainability 13 09355 g011
Figure 12. Spectral analysis of the HFCGS output voltage curve at C2 = 1000 μF.
Figure 12. Spectral analysis of the HFCGS output voltage curve at C2 = 1000 μF.
Sustainability 13 09355 g012
Figure 13. Spectral analysis of the HFCGS output voltage curve at C2 = 2000 μF.
Figure 13. Spectral analysis of the HFCGS output voltage curve at C2 = 2000 μF.
Sustainability 13 09355 g013
Figure 14. Cost of hydrogen production for steam methane reformation (SMR) technologies, coal gasification and thermochemical water decomposition using HTGR.
Figure 14. Cost of hydrogen production for steam methane reformation (SMR) technologies, coal gasification and thermochemical water decomposition using HTGR.
Sustainability 13 09355 g014
Figure 15. Cost of producing compressed hydrogen for water electrolysis technologies based on electric energy from the network and renewable energy sources.
Figure 15. Cost of producing compressed hydrogen for water electrolysis technologies based on electric energy from the network and renewable energy sources.
Sustainability 13 09355 g015
Figure 16. Specific cost of hydrogen transportation in different forms by various transport.
Figure 16. Specific cost of hydrogen transportation in different forms by various transport.
Sustainability 13 09355 g016
Figure 17. LCOE for diesel, gasoline and hydrogen power plants of the equal capacity.
Figure 17. LCOE for diesel, gasoline and hydrogen power plants of the equal capacity.
Sustainability 13 09355 g017
Table 1. Structure features of hydrogen fuel cells.
Table 1. Structure features of hydrogen fuel cells.
Fuel Cell TypeElectrolyteElectrode MaterialAnodic ReactionCathodic ReactionCritical CO Content in Fuel
PEMFCPerfluoro-sulfonic acid (Nafion)Anode: Pt, PtRu
Cathode: Pt
H2→2H+ + 2e0.5O2 + 2H+ + 2e→H2OCO > 10–100 ppm
AFCPotassium hydroxide, soaked in a matrix KOHAnode: Ni
Cathode: Ag
H2 + 2OH→2H2O + 2e0.5O2 + H2O + 2e→2OHCO > 10 ppm
PAFCPhosphoric acid soaked in a matrixAnode: Pt, PtRu
Cathode: Pt
H2→2H+ + 2e0.5O2 + 2H+ + 2e→H2OCO > 0.5%
MCFCCarbonate solutionAnode: Ni-5Cr
Cathode: NiO(Li)
H2 + CO32−→H2O + CO2+2e0.5O2 + CO2 + 2e→CO32−does not poison the catalyst
SOFCYttria-stabilized zirconia, or more recently, lanthanide-doped ceriaAnode: Ni-YSZ
Cathode: Lanthanum strontium manganite (LSM)
H2 + O2−→H2O + 2e0.5O2 + 2e→O2−does not poison the catalyst
Table 2. Lithium battery cell parameters.
Table 2. Lithium battery cell parameters.
TypeLi-IonLiFePO4LTO
Type (form factor)cylindricalcylindricalprismaticcylindricalprismatic
Capacity C, Ah0.7–71.1–5010–2801.3–402.9–23
Gravimetric energy density, Wh/kg120–27080–140110–16570–9646–90
Cell voltage, V3.63.23.22.42.3–2.4
Standard charge current, C∙A0.50.50.50.51.0
Maximum charge current, C∙A1.01.01.05.04.0
Long-term discharge current, C∙A2.01.0–3.01.0–3.01.0–5.01.0–3.0
Maximum discharge current, C∙A3.03.0–5.03.0–5.05.0–10.04.0–10.0
Service life, cycles (not less)10002000350010,00015,000
Charge-discharge depth, %SOC20–8020–8020–800–1000–100
Operating temperature (discharge), °C−20–+60−20–+60−20–+55−30–+60−30–+60
Specific cost, USD/Ah1.0–2.50.7–3.00.5–0.90.75–3.01.0–2.1
Table 3. HFCGS key units and parameters.
Table 3. HFCGS key units and parameters.
UnitSymbol (Figure 3)TypeNominal Parameters
Fuel cellFCPEMFC1 kW, 46 V
Battery storageBCLiFePO4, cylindrical3.5 Ah, 12 × 36 V
ConverterCsingle-phase bridge, MOSFET/IGBTsee Section 3.5
Autonomous voltage inverterAVIsingle-phase bridge, MOSFET/IGBTsingle-phase, 220 V
FiltersC1, C2capacitorssee Section 3.3
RelaysKV1-KV4solid-state-
LoadZL-3 kW, 220 V
Electrical mainsUif applicable220 V
Table 4. HFCGS battery storage system parameters.
Table 4. HFCGS battery storage system parameters.
ParameterValue
Capacity C, Ah/kWh3.5/1.6
Nominal voltage, V460.8
Maximum voltage UBSmax, V518.4
Minimum voltage UBSmin, V360
Continuous charge current, A3.5
Maximum discharge current, A28
Internal resistance AC 1 kHz rBSmax, Ω2.6
Operating temperature t, °C−20–+50
Volumetric Energy Density, Wh/L306
Gravimetric Energy Density, Wh/kg125
Table 5. Parameters of the HFCGS power circuit elements.
Table 5. Parameters of the HFCGS power circuit elements.
ParameterValue
Load rated power P, kW3
Power factor cos φL0.8
The effective value of the HFCGS output voltage (single phase) Uout, V220
Output voltage frequency fout, Hz50
FC rated power PFC, kW1
Battery voltage UBSmin/UBSmax, V360/500
Converters clock frequency fCL, kHz20
Output L-C filter L, μH/C, μF500/2
Capacities C1, μF/C2 and C3, mF235/1.0 and 2.0
Table 6. Comparative analysis of various methods for producing compressed hydrogen [25,26].
Table 6. Comparative analysis of various methods for producing compressed hydrogen [25,26].
H2 Production TechnologyEnergy Consumption per 1 kg H2CO2 Emission per 1 kg H2, kg
Natural Gas, m3Coal, kgH2O, kgElectrical Energy, kWhThermal Energy, kWh
Steam methane reformation5–5.5-4–4.50.7–0.9-9.5
Coal gasification-7–7.590.7–0.8-21
Thermochemical water splitting based on HTGR--9–202–2.560–651.7
Electrolysis (from an electrical network)--955–60-41.1
Electrolysis (from a wind turbine)--955–60--
Electrolysis (from a solar energy station)--955–60--
Table 7. Cost value estimation of hydrogen production [25,26,27,28].
Table 7. Cost value estimation of hydrogen production [25,26,27,28].
ParametersValue
Energy prices
Natural gas, thousand m3111
Coal, USD/ton112
Electric energy from a centralized source, USD/kWh0.05
Specific capital investment in auxiliary energy sources
HTGR, USD/kW (thermal power)500
Solar power plants (SPP), USD/kW1000
Wind power plants (WPP), USD/kW500
Specific investment in elements of hydrogen production technology
Electrolyser, USD/kW740
Methane reformer, USD/kg H225–30
Hydrogen compressor, USD/kW2000
H2 liquefaction unit, USD/kg H2/day1100
The cost of transporting hydrogen per 100 km
Compressed pipeline, USD/kg0.09
Liquefied by auto transportation in cryogenic tanks, USD/kg0.04
Table 8. Comparison of the hydrogen cost estimation results by Russian specialists with global data.
Table 8. Comparison of the hydrogen cost estimation results by Russian specialists with global data.
Hydrogen Production TechnologyHydrogen Cost, USD/kg H2
Estimates of Russian ExpertsEstimation of Global Experts
compressedSMR (natural gas)1.2–2.71.8–3.5
Coal gasification1.9–2.31.6
Thermochemical water splitting with HTGR3.3–7.51.0–1.6
Electrolysis (from an electrical network)4.3–9.34.7
Electrolysis (from a wind turbine)4.4–25.93.9–7.1
Electrolysis (from a solar energy station)7.1–506.4–25.8
liquefiedSMR (natural gas)2.9–4.43.8
Coal gasification3.8–4.14.5–5.1
Thermochemical water splitting with HTGR5.7–10.21.4–2.1
Electrolysis (from an electrical network)5.7–11.67.8
Electrolysis (from a wind turbine)6.0–31.34.5–9.5
Electrolysis (from a solar energy station)9.1–60.07.5
Table 9. Initial data for electric energy cost calculation.
Table 9. Initial data for electric energy cost calculation.
ParameterValue
Fuel cell power, kW1
Hydrogen consumption, liters per minute12
Hydrogen source service life, years10
The total cost of hydrogen source experimental sample (with hydrogen generator), thousand USD50
Distilled water cost, USD/m318
Inflation rate, %5.1
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kulikov, A.; Loskutov, A.; Kurkin, A.; Dar’enkov, A.; Kozelkov, A.; Vanyaev, V.; Shahov, A.; Shalukho, A.; Bedretdinov, R.; Lipuzhin, I.; et al. Development and Operation Modes of Hydrogen Fuel Cell Generation System for Remote Consumers’ Power Supply. Sustainability 2021, 13, 9355. https://doi.org/10.3390/su13169355

AMA Style

Kulikov A, Loskutov A, Kurkin A, Dar’enkov A, Kozelkov A, Vanyaev V, Shahov A, Shalukho A, Bedretdinov R, Lipuzhin I, et al. Development and Operation Modes of Hydrogen Fuel Cell Generation System for Remote Consumers’ Power Supply. Sustainability. 2021; 13(16):9355. https://doi.org/10.3390/su13169355

Chicago/Turabian Style

Kulikov, Aleksandr, Aleksey Loskutov, Andrey Kurkin, Andrey Dar’enkov, Andrey Kozelkov, Valery Vanyaev, Andrey Shahov, Andrey Shalukho, Rustam Bedretdinov, Ivan Lipuzhin, and et al. 2021. "Development and Operation Modes of Hydrogen Fuel Cell Generation System for Remote Consumers’ Power Supply" Sustainability 13, no. 16: 9355. https://doi.org/10.3390/su13169355

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

Article Metrics

Back to TopTop