Next Article in Journal
Integrated Energy Systems Modeling with Multi-Criteria Decision Analysis and Stakeholder Engagement for Identifying a Sustainable Energy Transition
Previous Article in Journal
Characterization of Architecture Bounding Surfaces in Fluvial Tight Sandstone Reservoirs and Their Influence on Remaining Gas: A Case Study from the Suzhong Block, Sulige Gas Field
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Economic and Environmental Analyses of an Integrated Power and Hydrogen Production Systems Based on Solar Thermal Energy

by
Zarif Aminov
1,2,3,*,
Khusniddin Alikulov
3 and
Tran-Dang Xuan
1,4,*
1
Graduate School of Advanced Science and Engineering, Hiroshima University, 1-5-1 Kagamiyama, Higashihiroshima 739-8529, Japan
2
Scientific Technical Center, Academy of Sciences, 32 Durmon Yoli, Mirzo Ulugbek, Tashkent 100125, Uzbekistan
3
JSC “Thermal Power Plants”, Ministry of Energy, 23, Bunyodkor Avenue, Tashkent 100097, Uzbekistan
4
Center for the Planetary Health and Innovation Science (PHIS), The IDEC Institute, Hiroshima University, 1-5-1 Kagamiyama, Higashihiroshima 739-8529, Japan
*
Authors to whom correspondence should be addressed.
Energies 2024, 17(17), 4264; https://doi.org/10.3390/en17174264
Submission received: 7 August 2024 / Revised: 19 August 2024 / Accepted: 21 August 2024 / Published: 26 August 2024
(This article belongs to the Section A5: Hydrogen Energy)

Abstract

:
This study introduces a novel hybrid solar–biomass cogeneration power plant that efficiently produces heat, electricity, carbon dioxide, and hydrogen using concentrated solar power and syngas from cotton stalk biomass. Detailed exergy-based thermodynamic, economic, and environmental analyses demonstrate that the optimized system achieves an exergy efficiency of 48.67% and an exergoeconomic factor of 80.65% and produces 51.5 MW of electricity, 23.3 MW of heat, and 8334.4 kg/h of hydrogen from 87,156.4 kg/h of biomass. The study explores four scenarios for green hydrogen production pathways, including chemical looping reforming and supercritical water gasification, highlighting significant improvements in levelized costs and the environmental impact compared with other solar-based hybrid systems. Systems 2 and 3 exhibit superior performance, with levelized costs of electricity (LCOE) of 49.2 USD/MWh and 55.4 USD/MWh and levelized costs of hydrogen (LCOH) of between 10.7 and 19.5 USD/MWh. The exergoenvironmental impact factor ranges from 66.2% to 73.9%, with an environmental impact rate of 5.4–7.1 Pts/MWh. Despite high irreversibility challenges, the integration of solar energy significantly enhances the system’s exergoeconomic and exergoenvironmental performance, making it a promising alternative as fossil fuel reserves decline. To improve competitiveness, addressing process efficiency and cost reduction in solar concentrators and receivers is crucial.

1. Introduction

The energy supply sector is the largest contributor to global greenhouse gas emissions [1]. The pollution and harmful emissions from fossil fuels have prompted a shift toward renewable energy (RE) sources [2]. Among these, solar and biomass energy stand out due to their global availability and abundance. Biomass is a carbon-neutral renewable energy source that offers several advantages for the energy sector. It can be used as a primary feedstock or in co-gasification processes due to its abundant reserves, renewability, high reactivity, and low pollutant emissions [3]. While solar energy is not consistently available and high-temperature solar technology incurs higher costs than other RE sources, it is entirely carbon-free [4]. In contrast, biomass energy is cost-effective and does not share the intermittent issues of solar energy [5]. Combining solar and biomass energy as heat sources for power generation systems is a relatively new approach. This method maximizes the advantages of both energy sources while mitigating the drawbacks of using them individually [6].
Studies indicate a growing interest among researchers in using RE sources and integrating them to reduce pollutant emissions and control climate change [7]. However, solar energy’s main drawback is its fluctuating and intermittent supply. A solution to this issue is the use of hybrid resources, such as combining solar energy with thermal energy storage (TES) [8] and hydrogen (H2) production [4,6]. In this respect, H2 can be stored and utilized for continuous power generation through systems like fuel cells [9] and internal combustion gas turbines (GT) [4,10], ensuring a steady power supply.
Recently, there has been significant attention given to hybridizing solar and biomass energies in power generation systems. For most of the investigations, concentrating solar power (CSP) plants present CO2 emissions of less than 40 gCO2/kWh [11]. Incorporating low-carbon fuels like H2 is crucial for achieving net-zero goals through clean energy technologies. H2 is also seen as a future fuel source for the transportation sector and is applicable in the petroleum and chemical industries [12]. Consequently, global spending on H2 energy research and development has increased, as it has begun to represent a major green energy opportunity in recent decades.
Numerous studies have examined different configurations of biomass–solar hybrid systems aimed at generating multiple valuable outputs in power production. The combined cycle approach is a proven and efficient method for improving the performance of traditional high-temperature power plants. Within this framework, some research has centered on proposing and exploring combined cycle systems [13] and hybrid solar and fuel-fired plants [14], particularly in Uzbekistan. Zarif et al. [15] examined the hybridization of biomass with solar energy from CSP systems to operate a syngas, H2, and power production cycle in a combined heat and power (CHP) plant. Their findings demonstrated that using a hybrid resource significantly increases energy and exergy efficiency by 37.07% and 32.52%, respectively, compared with the natural gas-fired Tashkent CHP plant.
Combining multiple energy sources can ensure stability and energy balance. The economic potential of these technologies is promising enough to meet increasing demand [16]. However, if thermal systems are optimized solely based on thermodynamic factors, the overall cost can rise significantly. Thus, it is crucial to evaluate thermal systems from thermodynamic, economic, and environmental perspectives to enhance the efficiency of energy systems [17,18].
Energetic, exergetic, exergoeconomic, and exergoenvironmental (4E) analyses [19,20] have recently emerged as valuable methods for evaluating and assessing the feasibility of energy systems. These analyses offer a comprehensive overview of performance by considering both the quantity and quality of energy, the amount of natural resources consumed, and the resulting cost of a specific product. They allocate costs to exergy and provide insights into cost distribution based on irreversibility, which is beneficial for the optimization, design, control, and diagnostics of energy systems [16]. Additionally, 4E analyses facilitate benchmarking and comparison between energy systems based on environmental aspects, using primary factors and life cycle assessment (LCA) considerations [7].
Several studies have explored the use of biomass and solar energies as simultaneous heat sources for energy conversion systems. Notably, Cao et al. [4] and Cen et al. [16] proposed hybrid systems that integrated H2 post-firing, achieving carbon dioxide (CO2) emissions of 0.475 kg/kWh and 0.45 kg/kWh, respectively. The levelized cost of electricity (LCOE) for these systems was determined to be 61.37 USD/MWh and 65.8 USD/MWh. Another study [7] that investigated solar–biomass waste poly-generation systems demonstrated energy and exergy efficiencies of 29.25% and 23.59%, respectively, with a payback period of 4.28 years.
Mohammad et al. [5] studied a hybrid heat source combining a solar power tower (SPT) and biomass gasification in a multi-generation system. Their findings showed an enhancement in total exergy efficiency of 9.04% points, attributable to waste heat recovery from the GT. The total cost rate of the proposed system was 7799 USD/h, with a unit cost of the multi-generation system of 8.26 USD/GJ. Another study proposed a multi-generation system utilizing a CO2 Brayton cycle, which resulted in a 76.9% reduction in the exergy destruction cost and a 25.9% decrease in the unit cost of the system. A poly-generation plant driven by a solid oxide fuel cell (SOFC) was found to have CO2 emissions of 10.79 kmol/MWh, with the unit cost of exergy and the total cost rate of the products being 11 USD/GJ and 223 USD/h, respectively [21].
In the CO2 transcritical cogeneration cycle [17], the product cost decreased by 9% when the H2 production rate and refrigeration power were reduced from 6.425 kW to 6.103 kW. In another system focused on H2 production and CO2 emission reduction using solar energy [6], H2 injection achieved an output power of 346.2 kW, an exergy efficiency of 24.85%, CO2 emissions of 0.257 kg/kWh, and a levelized cost of product (LCOP) of 0.0911 USD/kWh. Ehsanolah et al. [22] analyzed a solar–biomass integrated energy system that produced 7967 kW of power and 10.74 kg/h of liquid H2 with an exergy efficiency of 11.06%. The results demonstrated an LCOE of 40.16 USD/MWh, and CO2 emissions of 0.39 t/MWh. Modified exergy-based economic analyses showed a 0.24% reduction in exergy destruction, a 3.36% decrease in the total exergy destruction cost rate, a 2% reduction in the CO2 discharge rate, and a 3% decrease in the total unit product cost [23].
To boost H2 production, a plant incorporated steam reforming technology and a proton-exchange membrane (PEM) [24]. The results indicate that the unit product cost and environmental penalty cost rate were approximately 23.7 USD/GJ and 141.7 USD/h, respectively. Rahbari et al. [25] explored the production of solar fuels through supercritical water gasification (SCWG) of algae, estimating a fuel cost of 2.99 AUD/L (2.06 USD/kg), with H2 costing 2.1 AUD/kg (1.45 USD/kg).
Ding et al. [26] developed a multi-generation energy system that achieved an exergetic efficiency of 31.1% and a net product unit cost of 51.7 USD/GJ. The system’s levelized cost of hydrogen (LCOH) and LCOE were 5.67 USD/kg and 0.098 USD/kWh, respectively. Another study investigated an integrated system using biomass and solar energy for H2 production with an electrolyzer [27]. Their results indicated power generation costs of 0.014 USD/kWh and environmental impacts of 0.002 Pts/kWh for biomass. Wang et al. [9] assessed an SOFC CHP system integrated with biomass gasification and solar-assisted carbon capture and energy utilization.
In the evaluation of integrating solar thermal and alternative energy into the SCWG of microalgae [28,29], minimum fuel selling prices (MFSPs) ranged from 52 to 73 AUD/GJ (39–54.85 USD/GJ).
A study on the solar-driven steam gasification of biomass for large-scale H2 production estimated the minimum H2 prices to be 2.99 EUR/kgH2 (2.62 USD/kgH2) for all thermal processes and 2.48 EUR/kgH2 (2.17 USD/kgH2) for hybrid processes compared with 2.25 EUR/kgH2 (1.97 /kgH2) for conventional methods [30].
Karthikeyan et al. [31] investigated a system that harnesses solar power during the day and uses abundant cotton waste biomass at night. The system exhibits an energy efficiency of 19% and an exergy efficiency of 11%, with a plant cost of 37 USD/h. Another study [32] reported an LCOE and LCOH at optimal total product unit costs of 0.045 USD/kWh and 28.86 USD/kg, respectively. Results indicated that the hybrid RE system offers a levelized cost of electricity and heat at 0.38 USD/kWh. Additionally, Lin et al. [33] introduced an electricity and liquid H2 cogeneration power plant with a total cost rate of 86.61 USD/h and an optimal environmental index of 1.09 kgCO2/kWh.
Mohammad et al. [34] proposed a system integrating sorption-enhanced biomass chemical looping gasification with solar energy, waste heat recovery, and power generation subsystems. Using rice straw biomass, the system proved to be more economical and environmentally friendly than a microalgae-based system, with a fuel cost of 116.9 USD/h and CO2 emissions of 5.22 t/h.
Matin et al. [35] designed and optimized a system for generating power, heat, and H2 via electrolysis. Their optimization achieved the highest exergy efficiency of 31.15% and the lowest LCOE at 0.0211 USD/kWh. The study revealed that the main advantage of integrating a biomass gasification–SOFC system with H2 production and injection is the significant reduction in CO2 emissions and the increase in power output, with CO2 emissions of 0.257 kg/kWh and an LCOE of 0.0911 USD/kWh [36].
Rahbari et al. [37,38] conducted research on a solar fuel plant utilizing SCWG integrated with Fischer–Tropsch synthesis, employing steady-state modeling and system-level dynamic simulation. The proposed systems demonstrated a levelized cost of fuel (LCOF) that was as low as 2.44 USD/L with a 15 h syngas storage capacity and 2.50 USD/L with a 20 h syngas storage capacity.
Hashemian et al. [39] examined an assessment and multi-criteria optimization of a solar- and biomass-based multi-generation system. Their study demonstrated improvements of 82.4% in energy efficiency and 14% in exergy efficiency, a total product cost rate of 0.84 USD/s, and an exergoenvironmental impact of 0.15.
The solar-driven thermochemical conversion of empty cotton boll biomass to syngas and potassic fertilizer has been evaluated based on experimental results [40]. The study reports that the LCOF via Fischer–Tropsch synthesis was estimated at 0.90 USD/L for a 20 MW syngas plant. Additionally, an economic analysis [41] of H2 production revealed that the LCOH was reduced to approximately 3.67 USD/kg, aligning with the 2030 target value of 4000 KRW/kgH2 (3.06 USD/kgH2) set by the Korean government’s H2 economy revitalization roadmap. The study confirms that integrating various energy sources into the system enhances resource flexibility [20]. The results showed that the overall unit product cost was 21.79 USD/GJ, while the overall social–ecological factor was calculated to be 1.37.
The literature review on solar-powered plants integrated with biomass for power and hydrogen production reveals that most research has focused on specific hybrid power plants. These studies often combine different cycles, such as multi-generation and poly-generation systems, to produce a variety of fuels and products. Additionally, research has investigated the use of solar energy through PV/PVT panels or medium-temperature solar systems. However, there is a significant gap in studies that assess the techno-economic and environmental aspects of various biomass-fueled CHP systems, including hydrogen generation, using a unified energy mix approach that prioritizes biomass and solar energy sources.
Moreover, the feasibility analysis of dynamic simulations in biomass-fired GT systems remains underexplored in the existing literature, as most studies have concentrated on steady-state modeling, primarily involving solar energy. To address these gaps, this paper introduces an innovative configuration of a hybrid biomass–solar system, where solar energy is integrated as a co-feed to support the energy requirements of a biomass-fueled CHP plant. This integration aims to lower CO2 emissions while increasing product output through hydrogen production. The required solar thermal energy is generated using SPT technology equipped with a TES system that operates at high temperatures. Building on previous studies that analyzed thermodynamic performance in steady-state conditions, this research conducts exergoeconomic and exergoenvironmental analyses of hydrogen and power production from various processes using a CSP-system-level dynamic simulation.
Four scenarios were considered to evaluate the proposed models, namely conventional biomass gasification (CBG), sorption-enhanced reforming (SER) with calcium sorbent, chemical looping reforming (CLR) using iron oxygen carrier (OC), and SCWG for CO2 capture. The main objectives of this study were to:
  • Design a new solar-based cotton stalk gasification system using system-level dynamic simulation that produces renewable H2, heating, and electricity that can effectively avoid solar fluctuation and achieve efficient solar thermochemical conversion.
  • Investigate the performance of different biomass-conversion-based systems by calculation of the exergoeconomic and exergoenvironmental impact factors.
  • Analyze the impact of related parameters on an optimized H2 and power production system with a CSP plant to evaluate the overall system’s performance.

2. System Description and Simulation

The integrated CHP plant with biomass and solar energies for H2 and power production investigated in this study is based on a combination of CSP-derived thermal energy with different biomass conversion plants, a CO2 capture plant, a H2 purification unit, a steam cycle, a H2S remover unit and a CHP plant. A detailed validated steady-state physical model of the plant has been developed in Aspen Plus® [42], which was discussed in the thermodynamic modeling study by the authors [15]. In this regard, optimal parameters were obtained to maximize the H2-rich syngas production and CO2 capture in the proposed models. However, future research should also account for factors like the catalytic performance in H2 production to achieve higher activity and stability in the production process [43]. Additionally, various co-catalysts have been extensively utilized in CO2 reduction, effectively lowering the energy barrier of thermal reactions and enabling milder reaction conditions [44].
To analyze the dynamic behavior of the plant at a system level under a variable solar thermal source, an energy-based model was developed using the Ebsilon® Professional software [45], which is widely used for the modular modeling and simulation of complex dynamic systems. The proposed model utilizes polynomial performance curves and design point parameters to simulate key component metrics such as mass flows, energy flows, and efficiencies. These elements are integrated to create a system-level dynamic model of the entire plant. These dynamic models for each proposed system have been developed with the Time Series Calculation from the design and off-design simulation in Ebsilon® Professional [45].
The schematic diagrams of the proposed hybrid biomass–solar driven CHPs with H2 production are presented in Figure 1, Figure 2, Figure 3 and Figure 4. Each configuration consists of a GT cycle and a ST cycle with capacities of 30 MW and 27 MW of electricity, respectively. The systems are powered by hybrid biomass–solar resources. Cotton stalk biomass is the primary fuel, while solar energy is used for meeting the heat demand of biomass drying, preheating, syngas production via gasification and reforming, steam generation, H2S cleaning, and the regeneration of solvent.
Except for the SCWG process, the systems work by firstly drying the biomass fuel to convert it into a higher calorific biomass, followed by the production of synthesized gas via various biomass conversion procedures, namely gasification with air and steam agents (Figure 1), calcium looping reforming based on a reversible reaction between CO2 and CaO limestone (Figure 2), chemical looping reforming with an iron oxygen carrier (Figure 3), and technology that connects to a process where a metal with variable oxidative states is applied as the oxygen carrier (OC) and SCWG is carried out with a pressurized water feed (Figure 4) in the reactor. The oxygen required for the gasification is obtained from the air separation unit (ASU) in the conventional biomass gasification process (Figure 1).
After the gasification and reforming stages, the syngas is cooled to the operational temperature before being introduced into each reactor to improve the CO conversion and efficient heat recovery. H2S and other sulfur compounds are absorbed by the ZnO fixed bed [17,46].
The gaseous mixture is separated through the PSA unit after cooling to 350 °C to produce a hydrogen stream with 99.99% purity. The purification efficiency of the PSA unit is about 80%, resulting in purged gases that are recycled for combustion in the GT to produce electricity. The exhaust gas after the GT combustion is utilized by the HRSG to supply the heat. The produced steam ensures the availability of steam for the steam cycle for two turbines, one with low pressure (LP) and one with high pressure (HP), which are used to generate power. The HP ST is used only for power generation, but the LP ST is also used for heat production in addition to power generation.
In the calcination reactor, CO2 is first captured and then regenerated via a sorbent, followed by a highly endothermic calcium carbonate decomposition reaction to produce CO2. In other models, the gas is decarbonized using post-combustion (model 1 and 4) and pre-combustion (model 3) MEA-based gas–liquid absorption reactions. The heat demand for solvent regeneration by the CO2 capture unit is provided by solar energy.
The main demand for solar energy is from the gasifier reactors, with the highest consumption in models 1, 2, and 4. In the case of the iron-based chemical looping model (model 3), the heat demand of the reduction process is provided by the thermal balance of exothermic reactions in the air reactor.
To simulate the proposed integrated systems, the following assumptions were made:
-
The compositions of ambient air are 21% O2 and 79% N2 molar fractions.
-
Compressors, pumps, and turbines are simulated using isentropic efficiency values.
-
Calculation of the produced syngas is carried out using the principle of minimizing Gibbs free energy.
The main input parameters used for the modeling are listed in Table 1. The weather data source from the Meteonorm Database [47] was used to create a more detailed and accurate dynamic model of the whole system with the CSP plant, as it fills in the gaps between known data points, providing a continuous representation of the system’s performance across a range of conditions. A sub-component instance can be created within any component that requires weather or solar position data, and it is then connected to its corresponding counterpart in the weather data source component. The case study in the present work is based on an annual year of hourly weather data for Tashkent city, Uzbekistan. This site was chosen due to the high annual average of solar resources and the reference power plant in the nearby area.
The solar field utilizes central tower technology, employing an array of heliostats to reflect concentrated sunlight onto a receiver positioned at the tower’s apex. A cavity receiver was preferred in this study and a polar field layout was chosen as it is more adapted to cavity receiver designs [38]. The CSP plant was designed using the Ebsilon® Professional [45] commercial package. The EbsSolar model was used to simulate the performance of the SPT plant with the heliostat field, storage tank, and its other components [45].
Since the chosen SPT plant is operated at high temperatures (600–870 °C), which is possible in a large-scale production [49], the NaCl:KCl:ZnCl2 combination was considered in this study for the heat transfer fluid (HTF) [15]. The optimized solar field layout has 88.41% of effective reflectivity heliostats, each being 10 × 10 m2 in area size around a tower of 126.9 m height at a DNI design value of 845 W/m2, which is varied up to 1014 W/m2 in dynamic simulations of the CSP system. According to a report from the World Bank [50], the economic lifespan of a CSP is approximately 25–30 years. For this reason, the current study examined 25 years of plant life.
The field outputs a flux density distribution on a defined aperture surface that serves as an interface between the optical concentrator and the solar receiver. Each heliostat’s efficiency, defined as the ratio of reflected irradiance reaching the receiver to the irradiance accepted by its mirror surface, depends on its location within the field, which affects its cosine angle and distance to the aperture plane. In the solar receiver, the heat flux arriving at the aperture is transferred to the heat transfer medium, raising the fluid temperature or steam fraction in a steam receiver. Optical and thermal losses occur between the aperture and the fluid, which are modeled within the receiver component.
The third main unit of CSP plant consists of the thermal energy storage tanks (hot and cold tanks). Thermal energy is stored to mitigate intermittency, ensuring continuous operation during periods of low solar energy supply [20]. This approach enhances the overall system efficiency and reduces both the initial and operating costs.

3. Modeling

This section is divided into five subsections, covering the methodologies for the thermodynamic investigations, economic analysis, exergoeconomic analysis, environmental assessment, and exergoenvironmental analysis. The assumptions and input data used in the exergy-based economic and environmental modeling are outlined in Table 1. Moreover, the detailed modeling performance results are included in Appendix A.

3.1. Thermodynamic Modeling

Aspen Plus® [42] and Ebsilon® Professional [45] software were used to create the thermodynamic model of the proposed models integrated with the solar-assisted system. For system feasibility assessment from the first- and second-law standpoints, thermodynamic models were developed.
According to the principle of the conventional state, knowing two independent physical properties allows for the calculation of other thermodynamic properties [27].
The energy balance principle for a control volume is applied as
i m i h i + i Q i = e m e h e + e Q e + e W e
where Q i and W e stand for the heat and electrical power, respectively, while i and e denote input and output streams of the system.
The energy efficiency information measures the value of the power, H2 production, and cooling capacity generated by the system relative to the energy used. It evaluates system performance variations based on the solar radiation and biomass flow rate in the trigeneration mode [31].
Exergy analysis is a more robust method for evaluating thermal systems, as it helps in formulating strategies for efficient energy use and in identifying the magnitude and locations of inefficiencies. This analysis requires the exergy to be determined for all the streams. Neglecting kinetic and potential exergies [24], the steady-state exergy rate balance equation is typically expressed as follows:
i m ˙ e ˙ i + E ˙ i Q = e m ˙ e ˙ e + E ˙ e W + E ˙ D
In the above relationship, E ˙ i Q , E ˙ e W , E ˙ D , and e ˙ are the exergies caused by the heat transfer flow and power exchange, exergy destruction, and specific exergy rate flow rate, respectively. The nature of the mass flow exergy is separated into physical and chemical forms that can be calculated by using references [7,19].
Therefore, the overall exergy efficiency of the proposed hybrid system, which indicates the quality of their energy performance, is defined as follows:
ε s y s = E ˙ P E ˙ F = E ˙ n e t W + E ˙ H 2 + E ˙ t h Q E ˙ b i o m a s s + E ˙ s o l a r Q + E ˙ w a t e r Q + E ˙ a i r

3.2. Economic Modeling

The National Energy Technology Laboratory (NETL) approach for the economic analysis of power plants was utilized to determine the capital expenditure (CAPEX) and operational and maintenance (O&M) costs of investment for the whole power plant construction [9]. To evaluate total investment, we need to determine the investment cost of the equipment, so the relationship used to calculate the purchase equipment cost (PEC) of each piece of equipment is given in Table A1 in Appendix B.
The CAPEX included the total direct and indirect costs (TDICs), as well as the contingency cost ( C C o n t ) and the land cost ( C L a n d ) [9,40,51]. TDICs involved direct costs (DCs) and indirect costs (IDC); the detailed breakdown of the CAPEX is shown in Table 2.
C L a n d = U P L a n d A H e l i o f L a n d
OPEX costs are estimated considering two components: fixed and variable costs. The fixed costs cover the annuity cost, maintenance, property, taxes, and insurance costs. Variable costs cover the materials directly linked to the production process (e.g., biomass, catalysts, solvent, waste disposal, make-up etc.). Table 2 summarizes the key economic assumptions employed for the OPEX calculation in the economic analysis.
The specific CAPEX was computed as the investment cost (CAPEX) reported for the production of H2 and power [52]:
S p e c i f i c   C A P E X = T o t a l   i n v e s t m e n t   c o s t m ˙ L H V H 2 + W ˙ n e t + Q h e a t i n g  
To analyze the economic performance of the whole proposed system, the LCOP was assessed. Since the proposed system in this work is the co-generation of heat, H2, and electricity, a similar definition was used, which was
L C O P = A n n u a l i z e d   C A P E X + y = 1 N O P E X y = 1 N Q ˙ e + W ˙ n e t  
where m ˙ , L H V H 2 , and W ˙ n e t are the mass flow rate and LHV of the produced H2 and the produced heating and electricity power, and r and N are the discount rate and the life of plant for y t h year.

3.3. Exergoeconomic Modeling

The performance of the energy system in this study was extensively assessed using the specific exergy costing (SPECO) method of Lazzaretto et al. [34]. The examination of all the exergy flows and irreversible processes of the system are the basis of this method for exergy analysis. The procedure is to define the production structure, including the exergy fuel ( E ˙ F ), exergy product ( E ˙ P ), and exergy destruction ( E ˙ D ) for each system.
To estimate the cost of the system, the unit cost of each stream must be determined by employing the following overall exergoeconomic balance:
c Q , k E ˙ i Q + i c i E ˙ i k + Z ˙ k = e c e E ˙ e k + c W , k E ˙ e W
where c Q , k and c W , k represent the cost rate for the thermal energy input and work output rates, respectively.
The relative difference in cost unit ( r k ) and cost factor ( f k ) of the plant is defined as below for the exergoeconomic evaluation of the system [20,35,57,58]:
r k = c P , k c F , k c F , k
f k = Z ˙ k C ˙ D , k + Z ˙ k + C ˙ L

3.4. Environmental Modeling

CO2 emissions include both direct and indirect emissions [34], as well as emissions associated with the syngas, H2, and power generation. The total CO2 emissions are calculated as follows:
C O 2   e m i s s i o n = C O 2   e m i s s i o n d i r e c t + C O 2   e m i s s i o n i n d i r e c t
C O 2   e m i s s i o n i n d i r e c t = C O 2   e m i s s i o n t h e r m a l i n d i r e c t + C O 2   e m i s s i o n e l e c t r i c a l i n d i r e c t
Here, C O 2   e m i s s i o n t h e r m a l i n d i r e c t refers to the CO2 produced from thermal energy (205.3 lbCO2/MMBtu), and C O 2   e m i s s i o n e l e c t r i c a l i n d i r e c t refers to the CO2 generated from the electricity (0.89 t/MWh) supplied to the system [59]. It is important to note that direct CO2 emissions refer to the CO2 released directly into the environment, while the captured CO2 stored in tanks is not considered as CO2 emissions.
To assess the environmental performance across different energy levels, the CO2 and CO emissions of the proposed models are estimated using Equations (12) and (13) [60,61].
ξ C O 2 = m ˙ C O 2 e ˙ H 2 m ˙ H 2 + E ˙ h e a t i n g Q + E ˙ n e t W
ξ C O = m ˙ C O e ˙ H 2 m ˙ H 2 + E ˙ h e a t i n g Q + E ˙ n e t W
where m ˙ H 2 , m ˙ C O , and m ˙ C O 2 indicate the mass flow rate of the produced H2, CO, and CO2 emissions, and e ˙ H 2 denotes the specific exergy rate of H2.
To fully understand and evaluate the CO2 emissions of a process, a life cycle assessment (LCA) should be conducted as shown in Figure A1 [62]. Figure A2 shows the life cycle boundary of the proposed power systems.
The environmental impact of different hydrogen production technologies is typically assessed using metrics like the acidification potential (AP) and the global warming potential (GWP) [12]. These modules are then integrated into the OpenLCA model [63].
The system’s environmental impact is evaluated through the social ecological factor and a CO2 emission reduction indicator using Equation (14):
S E F = 1 1 ε s y s
where S E F is the social ecological factor.
The total environmental impact points of material, Pts, can be determined by Equation (15). The total time-dependent environmental impact points of material are calculated by Equation (16), assuming that the annual operating time ( n ) and the life of the system ( N ) are about 3600 h and 25 years, respectively.
y M = m C O w C O + m O M w O M + m D I w D I
Y M = y M N × n × 3600
Here, m and w are the equipment weight (in kg) and the standard environmental impact indicator (SEI; Pts/kg). C O ,   O M and D I represent the life cycle of a system containing the construction, operation and maintenance, and disposal and dismantling stages, during which the materials (for construction) and relevant services will lead to an environmental impact [64,65].
The average yearly LCC assessment of H2 production can be given as follows:
L C C H 2 = n = 0 N C A P E X + O P E X ( 1 + r ) n + C ˙ e n v m ˙ H 2

3.5. Exergoenvironmental Modeling

The overall exergoenvironmental balance used for the studied systems is as follows [65]:
i b E ˙ i + B ˙ i Q + Y ˙ k = B ˙ e W + e b E ˙ e
In the above Equation (18), b and Y ˙ k are the environmental impacts of flow per unit of exergy and the environmental impacts of equipment during their life cycle, respectively.
A thorough exergoenvironmental balance equation is established following the calculations in Equations (19)–(22) below [27]. According to Equation (19), B ˙ D considers the inefficiency of the kth component or system in terms of the environmental impact rate, while B ˙ t o t depicts the overall environmental impact of the units [64,66,67].
B ˙ D = b F E ˙ D
B ˙ t o t = B ˙ D + Y ˙ k
r b . k = b p . k b f . k b f . k
f b . k = Y ˙ k Y ˙ k + B ˙ D . k = Y ˙ k Y ˙ k + b f . k E ˙ x D . k
where Y k ,   B ˙ D . k , rb,k, and fb,k are the total environmental impact of the system’s LCA, environmental impact destruction, exergoenvironmental coefficient, and exergoenvironmental factor of the plant are defined as below for the exergoenvironmental evaluation of the system.
Furthermore, the principal exergoenvironmental indicators are formulated as follows [20,39,58].
Exergoenvironmental impact improvement:
Ω i i = 1 Ω d e i
where Ω i i clarifies the connection between the system and the environment. Unlike Ω d e i , the Ω i i parameter should be as high as possible.
The next important metric is the exergetic sustainability index:
Ω s i = Ω s f Ω i i
In this regard, the exergetic sustainability index for the desired system should be as high as possible. The exergy stability factor ( Ω s f ) can be calculated by Equation (25):
Ω s f = E ˙ P E ˙ P + E ˙ D , t o t + E ˙ F
where E ˙ P denotes the total product exergy in the system. The Ω s f of a favorable system approaches one.

4. Results and Discussion

4.1. Energy and Exergy Evaluations

In this section, the modeling results of the presented system for 4E analyses of the proposed plants for the Tashkent CHP plant will be examined in detail. To perform 4E analyses of the hybrid power system, a comprehensive computer modeling was developed using the Aspen Plus® [42], Ebsilon® Professional [45], and openLCA [63] tools, and the results of the analyses are given in detail below.
As defined in Section 3, the overall mass, energy, and exergy balances were utilized to quantify the most relevant plant performance indexes. Table 3 shows the most significant technical indexes calculated for the assessed power and green H2 generation concepts via decarbonized solar-based biomass gasification and reforming. Multiple energy resources, such as solar and biomass, are used in the system to enable sustainable operation without intermittence issues. Also, there is a thermal energy storage tank installed for the solar cycle to provide output even during low solar energy levels. All the concepts have the same biomass feedstock with different thermal inputs as well as the same steam and gas turbine capacities.
The exergy input rate of the biomass is the highest exergetic contribution to the system, comprising 60.61–65.39% of the total exergy fuel. The highest average hourly solar exergy fuel required is 194.57 MW due to the highest energy demand in model 3. In addition, the exergy product of the system consists of four parts heat and electricity generation by the turbines, captured CO2, and H2 production. In the case of CLR and SCWG, an extra expander contributes to the net power output of models 3 and 4. Therefore, the net power output of model 3 is 66.36 MW, with a lower energy consumption.
The system-level dynamic simulation design of the proposed plants indicates that the net power output can be decreased to 5.52 MW points when the H2 and power production effectiveness reach the highest index. The H2 production energy penalty is approximately 1.67 and 7.65% exergy efficiency in models 2 and 3, respectively.
It is obvious that model 4 has the highest exergy performance with 48.67% exergy efficiency, 13.16% more than that for model 3, which illustrates the high energy quality used in the systems. This is due to the greater amount of H2 production and a lower biomass exergy fuel requirement of the system. In addition, the system shows less exergy destruction and internal irreversibility (353.5 MW) compared with the other cases (models 1–3). In this regard, the exergy efficiency was calculated to be 35.61% in the system with the CLR biomass conversion. The main reasons for this showing the lowest performance are the low H2 production and higher entropy generation in a cooler. This is because the temperature difference between steam and the ambient temperature is higher. The largest exergy destruction yield occurs in model 2, whereas the coupling system with the SER process for biomass conversion contributes 57.42% of the exergy destruction yield of the exergy fuel. This result is attributable to the significant irreversibility associated with the chemical reaction and the heat transfer across the large temperature differences between the exergy fuel and the exergy product of the reactors and combustion chambers.
An annual simulation using an hourly time step was carried out for each of the proposed models. For each operating mode, the daily average solar heat energies attributable to each energy source by month and over the entire year are presented in Figure 5. The energy supply by the CSP drops significantly in the winter due to a lower DNI by up to 25.6 W/m2 (on 20 January). Note that while this mode reduces the exergy efficiency, the efficiency of the overall system was reduced in the range of 14.58–51.7% points in all cases.
As shown in Figure 6, the monthly exergetic efficiencies for the power plant and the CSP plant are presented. The highest DNI peak was observed for the first ten days in July (Figure 5). However, when considering the average monthly solar field exergetic efficiency, this ranges from 41.25% to 57.53%, with a peak in September for all models. This is to be expected, given the lower cosine losses and larger solar resources during the summer until the end of August. The performance of CSP plant 3 results in the highest index, which reaches 57.53% of the average monthly efficiency.
These findings provide valuable insights into the exergy analysis of the systems. To enhance the exergy efficiency, it is crucial to assess and optimize the irreversibility of each unit and subsystem in future studies.

4.2. Economic and Exergoeconomic Evaluations

The primary simulation results needed for the PEC calculations are listed in Table A5 in Appendix B. Based on Equations (A4) and (A5), each cost is calculated and updated to the year 2023 using the CEPCI cost index. This allows for the calculation of each PEC for the proposed models, which is essential for determining the total investment cost.
Figure 7 shows the breakdown of the PEC for the proposed systems. The lowest PEC for hydrogen and power production is USD 159.59 million for model 2 operating for 10 h daily. This figure highlights that the heliostat cost makes up the largest proportion of the PEC, being 24.54%, 41.79%, 37.05%, and 29.15% for models 1–4, respectively. The next most costly equipment is the MEA-based CO2 capture unit, except in model 2, which does not need extra decarbonization technology. This allows the SER-based system to significantly reduce the total investment cost. The PECs are detailed in Table A3 in Appendix B. Additionally, 15.26%, 22.53%, and 15.83% of the PEC of models 1, 2 and 4, respectively, are attributed to the pressure swing absorption unit for the H2 purification process. The equipment costs for the H2S removal section amount to USD 12.75–15.67 million.
For the CAPEX, the cost correlation method was conducted. The specific CAPEX costs were calculated using Equation (5), as presented in Table 4. Figure 8a presents the results for all the proposed models. Case 2, with the SER technology, has the lowest CAPEX of about 706.55 USD/kW of H2 and power, followed by the CLR model, which presents small differences between them (1029.68 USD/kW). The model with the conventional biomass gasification process has the highest CAPEX, which is 1173.74 USD/kW. This can be explained by the fact that the solar energy demand here is the greatest compared with the other cases, which results in increased costs for the heliostats and TES units. Moreover, an extra oxygen separation cost (ASU cost) and syngas production cost (WGS cost) was required, which affected the total cost of the model.
The next relevant economic performance parameter was the OPEX costs. The main economic assumptions used for OPEX evaluation are displayed in Table 2. The OPEX costs for renewable H2 and power generation systems with CO2 capture are described in Figure 8b.
The OPEX costs are enhanced by the decarbonization of H2 and by the power production based on biomass gasification. Model 1 has the greatest OPEX, mainly due to the high O&M cost of the CSP plant and the property, taxes, and insurance costs of the overall system. However, the material costs for the operation show a lower cost than for model 4. The lowest O&M costs were observed for the case using SER (model 2).
The LCOE, LCOH, and LCOP were calculated using Equations (12)–(14) as presented in Section 4. These key economic parameters are indicated in Table 5. As observed from Table 5, CaO- and Fe2O3-based chemical looping technologies applied in biomass gasification exhibited the lowest levelized cost of electricity. Between models 3 and 4, the SCWG case using gas–liquid CO2 absorption technology had a slightly lower electricity production cost by about 3.7%. Conventional biomass gasification had the highest electricity production cost.
In the case of hydrogen production, the lowest LCOH was observed in model 2, which was 12.8 USD/MWh for hydrogen production. However, the systems with conventional gasification and gasification with SCW processes had the greatest hydrogen production costs of 19.6 USD/MWh and 20.7 USD/MWh, respectively.
The calculated solar-based levelized cost of product (hydrogen, electricity, and heating) for each model was in the range of 9.4–14.9 USD/MWh. As can be seen in Table 6, the sorption-enhanced reforming-based system offered the most cost-economic result with 9.4 USD/MWh of the LCOP. The same was seen with electricity production, with the model with conventional biomass gasification indicating the highest hydrogen and power production cost compared with that for the other cases (14.9 USD/MWh).
The LCC of hydrogen was calculated based on the cost inventory data given in Table 5, and monetization of the life cycle environmental impacts was performed using the environmental cost unit price calculated from Equation (A18). Taking into account the discounting factor (r = 8%), the annual average LCC of 1 kg of hydrogen production from cotton stalk biomass was in the range of 1.47–2.39 USD over the plant lifetime of 25 years. Hydrogen production with the SER process was about 38.5% cheaper than for model 1 and 3, which were 71 USD/MWh and 72 USD/MWh, respectively. This is attributed to the low capital investment operating costs related to syngas production (WGS and cooling units) and there being no necessity for CO2 capture in the system.
The discounting factor (r) has been reported as 10% in a study [68]. At r = 10%, the LCC of 1 kg of hydrogen production increased by 34.4–36.4% compared with the base case (r = 8%) and resulted in 0.97–1.6 USD/kg in models 1–4.
From the LCC perspective, to make biomass- and solar-based hydrogen production competitive, the CAPEX of the SCWG system technology should be reduced to 50% of its present value. And, to reduce variable OPEX costs, energy-efficient biomass pyrolysis and methane reforming are vital. Whereas in the case of SCWG, to reduce both the variable CAPEX and OPEX costs, improvements should focus on the solar energy demand and heat recovery of the system. This can be achieved by several means, but a simple approach could be to reduce the energy losses in the units and streams of the system.
The major outcomes of the exergoeconomic study for the system are presented in Table 6. These parameters are provided to validate the exergoeconomic models of the processes. The cost stream flows consist of investment, work, exergy, and heat cost flows. SPECO was used to calculate the hydrogen unit costs using the exergy-related costs in thermal systems. In this work, the unit costs of the four cases using different biomass conversion methods were obtained. To calculate the specific exergy fuel ( c F ) and product ( c P ) costs, the exergy destruction ( C ˙ D ), total operational cost rate ( C ˙ D + Z T ), relative difference ( r k ), and exergoeconomic factor ( f k ) were used.
The lowest total cost rate of the proposed system (model 3) is 11,053.28 USD/h, in which the contribution of investment, exergy destruction, biomass consumption, and environmental cost rates are 48.8%, 16.5%, 32.6%, and 21%. As can be observed, the highest contribution to the total cost rate is related to the investment cost, which is followed by the cost rate of the biomass. However, the exergoeconomic factor is 69.77%, which indicates the difference between the cost rate of exergy destruction and the total cost rate of the system. In this regard, model 4 shows the best performance, with an 82.42% exergoeconomic factor.
Exergoeconomic analysis can predict the cost rate of the product. For example, by evaluation of the unit cost of products, it is determined that in the SER case, the cost for electricity, heating, and hydrogen production is equal to 6.78 USD/GJ; hence, the unit cost of hydrogen is obtained to be 15 USD/MW, which is much cheaper than that of conventional biomass gasification (27.2 USD/MW) and supercritical water gasification (22 USD/MW) for hydrogen production. Owing to the high rate of hydrogen generation compared with in other cases, the unit cost of the whole system with the biomass gasification process is 8.8 USD/GJ. Accordingly, model 2 has the highest contribution to the exergy destruction rate and, consequently, the cost rate of the total exergy destruction (2341.16 USD/h). However, the largest cost rate of exergy loss in the system is related to the CLR process due to the greatest loss due to cooling.
Figure 9 describes the amount of exergy efficiency, unit exergy cost of hydrogen, and the summation of the exergy destruction and investment cost rates C ˙ D + Z T for the system. From a thermodynamic point of view, among the systems with an influential impact on the performance of the system, model 4, with an exergy efficiency above 48%, has the best thermodynamic performance; on the other hand, model 2, with the lowest exergy efficiency, has the worst performance. However, model 2 has the lowest value of unit exergy cost of hydrogen and C ˙ D + Z T , which is equal to 6786.6 USD/h, 3054 USD/h, and 1943.5 USD/h points lower than for case 1 and case 4, respectively.
Based on the above description, to reduce the overall system cost, efforts should focus on lowering the investment cost of high-cost components and the exergy destruction cost of low-cost components. These findings can guide future process optimization. Additionally, to enhance systems with moderate cost factors, exergoeconomic analyses of each component should be conducted.

4.3. Environmental and Exergoenvironmental Evaluations

The environmental assessment of the proposed solar–biomass system includes evaluating both direct and indirect CO2 emissions. Direct CO2 emissions can be further divided into captured (using various technologies in models 1–4) and uncaptured CO2. Indirect CO2 emissions, arising from the consumption of electricity, steam, and heat, can be calculated using Equation (11). In this study, the electrical CO2 emission is zero since the required electricity for compressors and pumps is generated by the turbine. Thermal CO2 emissions are calculated by multiplying the thermal energy needed for heating and cooling the streams by 205.3 lbCO2/MMBtu (0.088 kgCO2/MJ).
As shown in Figure 10a, the total CO2 emissions for models 1–4 are 179.84, 153.31, 161.1, and 167.93 t/h, respectively. Model 1 has the highest direct and indirect CO2 emissions among the models. Models 1 and 2 have higher direct CO2 emissions because more CO2 is produced during the production and combustion of syngas. Since CO2 is captured via a CaO-based chemical looping process, model 2 does not require an additional CO2 capture unit, such as an absorption plant, resulting in zero thermal energy consumption and, thus, zero indirect CO2 emissions. Therefore, model 2 is more environmentally friendly than the other three systems. However, if we consider that the thermal energy consumed comes from an RE source, such as solar energy, indirect CO2 emissions can be neglected. In this scenario, model 3, which utilizes a CLR process, can be considered an environmentally friendly option for hydrogen and power production.
CO2 emissions are a crucial indicator of the thermo-environmental performance of energy systems. In the proposed system, which considers the poly-generation of electricity, heating, and hydrogen, emissions range from 26.56 to 93.82 kgCO2/MWh. Among the four models, the SCWG case has the lowest specific CO2 emission at 26.56 kgCO2/MWh. According to a report by the International Renewable Energy Agency (IRENA), biomass-based power plants had an average CO2 emission of 55 kg/MWh in 2020 [61]. Therefore, the emissions from the integrated system 4, followed by models 1 and 3, are within acceptable limits. Notably, the proposed system achieves a 93% CO2 capture efficiency, significantly reducing CO2 emissions. This highlights the environmental benefits of the system and underscores its potential as a sustainable energy solution.
Figure 10b illustrates the ecological footprint of various hydrogen production technologies based on the human toxicity potential (HTP), AP, and GWP from renewable energy sources. The values of HTP, AP, and GWP are normalized using similar equations across all the models. The CaO-based reforming process for H2 production is the most environmentally destructive, with a relatively high GWP of 3996.19 kgCO2/tH2, primarily because post-combustion CO2 capture technology is not applied. GWP values can vary in the literature due to differences in system boundary assumptions. The solar-based sorption-enhanced reforming model, however, is among the best technologies for producing clean hydrogen. Biomass gasification with supercritical water has a higher AP (1.056 kgSO2/tH2). In terms of CO2 emissions, gasification with steam–oxygen and supercritical water are the most eco-friendly technologies studied, as shown in Figure 10b.
Based on the SEI value of materials with the weight of each equipment in Table A1 (Appendix B), the results of the environmental impact points of components ( Y M ) for the system are obtained and listed in Table 7. The total environmental impact points of material are in the range of 24.39–37.42 mPts/h. The first system has the highest time-dependent environmental impact points of material, which is equal to 37.42 mPts/h, followed by model 4, with a somewhat similar index (35.86 Pts/h). The utility of the system containing biomass, hydrogen, heating, and electricity, according to Equation (6), and the environmental impact points of utility ( Y U ) in Table 8 are obtained.
The environmental impact points of utility are 40, 44, 36, and 38 times that of material for models 1–4, respectively. The greatest total environmental impact points of the LCA system are 2192.42 mPts/h in the SCWG-based system.
Table 8 presents the results of the exergoenvironmental analysis for the combined cycle using different syngas and hydrogen production plants. The findings reveal that the scenario with SER (model 2) has the highest environmental impact due to significant destruction, highlighting the substantial environmental consequences of the hydrogen and power production cycle in the hybrid system. Since the destruction ratio for each piece of equipment is based on fuel and product, the relative differences in environmental impacts are detailed, with the CLR-based power plant showing a notably high value of 4.68. Furthermore, the environmental impact of the product in case 2, at 2979.28 mPts/h, underscores the environmental significance of this model. The exergoenvironmental assessment of fuel, product, and destruction across all the cases demonstrates that the conventional biomass gasification plant has the least environmental impact compared with the other three models.
As shown in Table 8, the exergoenvironmental analysis results of the combined cycle with biomass conversion processes indicate that the SER case (model 2) has the highest destructive environmental impact, reflecting the significant environmental consequences of the hydrogen and power production cycle in the hybrid system. The destruction ratio for each piece of equipment is defined by its fuel and product, and the CLR-based model exhibits a particularly high relative environmental impact. The product environmental impact of the CLR process, at 2979.28 mPts/h, highlights the environmental significance of this model. Additionally, the exergoenvironmental evaluation of fuel, product, and destruction across all systems shows that the conventional biomass gasification plant has the least environmental impact compared with the other three models.
Implementing measures to enhance exergy efficiency is crucial for achieving lower B ˙ D values and for minimizing the environmental impact. Among the three components discussed, model 3 has the highest r b value of 4.68, indicating that its environmental impact can be relatively easily reduced by increasing hydrogen production and minimizing the physical exergy loss. On the other hand, model 4, with its lower r b and already high exergy efficiency, offers limited room for improvement, making it more challenging to reduce its environmental impact compared to model 3.
Sustainability and environmental metrics were defined for analysis. Using the exergoenvironmental impact factor ( Ω i f ), the amount of exergy fuel converted into irreversibility can be calculated. Figure 11a shows that the model with the SER process has the largest Ω i f . The Ω i f values were calculated to be 56.9%, 60.5%, 64.4%, and 51.3% for cases 1–4, respectively. According to Equation (A26), by decreasing the exergy destruction rate while keeping the exergy fuel of the system constant, the Ω i f is reduced. A reduction in the impact factor leads to improved system exergoenvironmental performance.
Higher exergetic efficiency results in a lower exergoenvironmental damage effectiveness index ( Ω d e i ). Similarly, a lower Ω i f factor corresponds to a lower Ω d e i index. As shown in Figure 11a, model 3 has the largest Ω d e i index. The ecological effect factor ( θ e f ) is a function of the exergetic efficiency of the systems. Plants with the highest θ e f values should be reviewed to improve their exergetic efficiency and, thus, reduce their ecological effect factor. Figure 11b shows the largest θ e f for the models, with the system using the CLR process having a factor of 2.808.
Exergoenvironmental impact improvement ( Ω i i ) is inversely related to the θ e f . If the θ e f is higher, the Ω i i is lower. Figure 11a indicates that model 3 has the smallest Ω i i , with a value of 0.553, followed by model 2 with 0.652. To achieve a higher Ω i i for any system, the θ e f must be reduced.
As hydrogen production increases and the total exergetic inefficiency of the system decreases, the exergetic sustainability index ( Ω s i ) improves, assuming that the exergy fuel remains constant. This enhancement in the sustainability index benefits the solar- and biomass-based system, making it more efficient in reducing exergy losses by achieving higher values of the index.
The final metric for evaluating the environmental and sustainability aspects is the social ecological factor (SEF), which incorporates an exergetic efficiency index. Model 3 has the smallest SEF, determined to be 1.55. As the system’s exergetic efficiency increases, the SEF also increases. Figure 11b shows the SEF of all the systems. The overall SEF of model 3 could be improved by increasing the exergetic efficiencies of its lower subsystems, such as the cooling units.

4.4. Parametric Studies

A parametric analysis was conducted to identify the key design and operating variables and to examine their effects on system performance.
Solar energy is dynamic and varies over time, making it crucial to study its effects on solar-based energy systems. Researchers have conducted various analyses to investigate solar radiation, with one common method being the monthly average analysis of solar radiation. This approach can report the operating conditions of energy systems on a monthly average basis. In this context, the radiation conditions of solar energy were examined for the most important functional parameters of the CSP plant. In addition, the influence of working hours of the CSP plant and the overall system on the techno-economic and environmental performance of the system was evaluated.
Figure 12a illustrates how time and exergy efficiency variations influence the ecological effect factor of the CSP plant, which is dependent on the system’s exergetic efficiency. This environmental indicator is derived from the average monthly exergetic performance of the CSP plant. During the monthly peak radiation period, the CSP plant in model 3, with its higher exergy efficiency, achieved the lowest ecological effect factor of approximately 1.738. The highest ecological effect factor, 1.974, was observed in the CSP plant of model 2. Conversely, CSP plant 1 had the highest ecological effect factor of 2.423 during a month with lower radiation, indicating that the availability of solar radiation significantly impacts this environmental indicator.
Figure 12b shows another environmental metric, the social ecological factor, which is based on the exergy efficiency of the CSP plant. The social ecological factor demonstrated a linear relationship with exergy efficiency and solar radiation, showing higher values during the average peak radiation month. The CSP plant in model 3, with its higher exergy utilization, had the highest social ecological factor value, ranging from 1.78 to 2.35, compared with other cities. The other CSP plants also maintained adequate values throughout the year, ranging from 1.70 to 2.25, indicating that this CSP system has exergy sustainability.
Figure 13a shows the variations in the exergy destruction rate and exergy destruction cost rate with the operation time of the system. It is seen that operation time has the same effects on the exergy destruction rate and the exergy destruction cost rate. Increasing operation hours increases the exergy destruction rate in the cycle as well as the exergy cost rate. On other hand, model 4 has the lowest cost of exergy destruction (1482.2–16,665.5 USD/h), especially in the 10 h, 12 h, and 14 h modes, followed by model 3, which can be attributed to the lower irreversibility due to the lower solar energy demand of the system. As indicated in Figure 13b, enhancement of the operating life of the CSP plant results in opposite influences on the specific CAPEX and investment cost rate of the system. As time increases, the investment cost rate decreases in all cases until it reaches a minimum; then, it rapidly increases due to the high cost of heliostats in models 3 and 4.
Referring to Figure 14, as the daily operation of the CSP plant increases, the LCOE firstly decreases via enhancement of the output power, and then decreases, resulting in the achievement of an optimum value for heat storage time (14 h), which minimizes the LCOE in the systems with CLR and SCWG processes. Such a trend for Wnet can be attributed to its direct relation with the operation time and investment cost of the system (Equation (A8)). The LCOH and LCOP have similar trends, according to Equations (A9) and (6). Thus, the lowest LCOE of the models was observed as 83.4, 49.2, 55.4, and 88.4 USD/MWh and as 13.9, 10.7, 19.5, and 16.1 USD/MWh for the LCOH for models 1–4, respectively. It shows the same results as the work by [41]. Based on this, the LCOH can be reduced by decreasing the CAPEX and investment cost rate of the system, suggesting that economic feasibility can be secured.
Figure 15 shows the impact of operation time variation on the unit cost of exergy products, relative differences, and exergoeconomic factors. As operation hours of the CSP plant increase, all the unit costs decrease, resulting in reductions of 8.2, 3.1, 1.4, and 3.3 USD/MW points for hydrogen costs in each case. Models 3 and 4 exhibit the lowest hydrogen costs in the 14 h mode. Regarding the exergoeconomic factor of the system, it also decreases with longer operation time. However, the exergoeconomic factor improves due to the high solar energy consumption at zero cost in the 16 h mode for models 3 and 4.
The next significant parameter is the environmental index, which is evaluated based on the exergoenvironmental criteria of the proposed solar- and biomass-based system, as shown in Figure 16. With a longer operation time, the impact factor and damage effectiveness index increase, while the impact improvement, sustainability index, ecological effect factor, and social ecological factor decrease. This is due to the reduced exergy efficiency and constant destroyed exergy, leading to worsened environmental impacts for the hybrid systems. Ultimately, it can be concluded that improving system performance can lead to a decrease in adverse environmental parameters.
According to Figure 17, as the working hours of the system increase, exergy destruction grows, and the environmental impact of this exergy destruction slightly increases, leading to a decrease in the exergoenvironmental factor. Additionally, increasing the capacity of the CSP plant results in an increase in both its weight and environmental impact rate. Notably, the exergoenvironmental factors of models 3 and 4 dramatically decrease in the 14 h mode.

4.5. Comparison with Other Solar- and Biomass-Based Power Plants

To better understand the performance of the proposed hybrid biomass–solar system, a comparison is made with various hybrid systems for multi- and poly-generation explored in previous studies. Each of these earlier studies investigated different processes and cycles for producing multiple products, such as electricity, fresh water, heating, cooling, CO2, hydrogen, and more. However, all these systems share the same goals: to reduce the cost of electricity and hydrogen and to minimize the system’s environmental impact. Table A5 presents a comparison of these previous works under optimal operating conditions. The comparison considers factors such as exergetic performance, levelized costs of electricity, hydrogen, and products, unit CO2 emissions, total costs, environmental impact rates, unit hydrogen costs, and exergoeconomic and exergoenvironmental factors.
In terms of thermodynamic performance, most of the previously studied systems have lower overall exergy efficiencies that range from 11% to 33%. In contrast, the models in the current study achieve exergy efficiencies of 34.2% to 44%. Model 4, with the highest exergy efficiency of 44%, also has the lowest CO2 emissions (26.56 kgCO2/MWh) compared with the other models in this study and those in previous works.
Referring to this table, it is evident that the proposed models 2 and 3, which utilize chemical looping reforming processes, demonstrate superior performance in terms of the LCOE, with values of 49.2 USD/MWh and 55.4 USD/MWh, respectively, compared with previous studies [4,26,61]. Although solar energy is an expensive technology, using it as the primary resource increases the system’s LCOE. A solar–biomass integrated system [22] and a multi-generation system [69] studied previously reported electricity production costs of 40.16 USD/MWh and 49.28 USD/MWh for the LCOE, respectively. On the other hand, other systems [27,35,66] have reported significantly lower LCOE values. It is suggested that these systems primarily produce electricity and other products through various cycles, with hydrogen production achieved via water electrolysis technology. Additionally, the proposed systems in this study include cost units for syngas production, hydrogen purification, and CO2 capture.
Due to the innovative nature of the proposed systems, which focus on producing green hydrogen by decarbonizing conventional CHP plants, it is important to provide insights into hydrogen production costs using different technologies and systems, such as biomass gasification and water electrolysis (utilizing solid oxide and proton-exchange membrane technologies). The proposed models have proven to be the most cost-effective processes to date, with LCOHs ranging from 10.7 to 19.5 USD/MWh and a unit exergy cost of hydrogen between 3.3 USD/GJ and 5.28 USD/GJ. Among the reviewed research, one plant [57] achieved the lowest unit exergy cost of hydrogen at 4.29 USD/GJ, while others [5,20,70,71] reported hydrogen unit costs between 15.74 and 45 USD/GJ. These cost differences are primarily due to three factors: the cost of the primary resource, the plant’s capacity, and the hydrogen production technology employed. Most of the reviewed studies utilized hydrogen production based on water electrolysis technology, which generally incurs higher production costs than other methods. This study proposes a system for syngas and hydrogen production using mature, commercialized technologies that offer higher efficiency and lower product costs [72]. Despite this, natural gas, mainly methane, is predominantly used for hydrogen production through steam reforming and partial oxidation processes, followed by coal gasification technology.
The total investment cost of a system is significantly impacted by the various processes, equipment, and units required. Due to these differences, the proposed model 2, which uses a CaO-based chemical looping process, has the lowest total cost rate at 6076.8 USD/h, with an LCOP of 7.8 USD/MWh and an LCC of 1.23 USD/kgH2. This cost is competitive with another system [68] that reports costs ranging from 0.47 to 3.48 USD/kg. Additionally, the biomass gasification method has the highest eco-cost at 11.82 USD/kgH2 [73].
Another important outcome of the study is the environmental impact indexes, which help to differentiate the environmental impact between the LCA and energy consumption. Table A5 provides a detailed comparison of the exergoenvironmental impact factor, with positive values from this study directly illustrating the relationship between the LCA impact and the exergy loss contribution.
A cradle-to-gate life cycle assessment was conducted by Chisalita et al. [74] to evaluate the environmental impact of producing 1 kg of hydrogen using innovative steam methane reforming and chemical looping technologies. They reported a GWP ranging from 0.31 to 9.65 kgCO2/kgH2. Ahmed et al. [8] proposed an enhancement in steam-only gasification, achieving the lowest economic annualized cost and environmental emissions per hydrogen production capacity of approximately 0.12 USD/kgH2 and 20 kgCO2/kgH2. In this study, model 2 demonstrates a relatively lower GWP of 3.99 kgCO2/kgH2.
The exergoenvironmental impact factor ranged from 66.2% to 73.9%, indicating that the exergy and environmental aspects are comparable in terms of environmental impact, consistent with industrial conditions. A small exergoenvironmental factor value for chemical looping reforming-based models indicates that the exergy loss contributes to the main environmental impact and should be given priority in future improvements. According to the results [69], the LCA values for the proposed multi-generation plant were estimated at 4.377 Pts/GJ (15.73 Pts/MWh), whereas the proposed models achieved LCA values of 5.4–7.1 Pts/MWh. However, another integrated solar- and biomass-driven energy system [66] reported an environmental impact of 1.96 Pts/MWh, suggesting that the proposed models need improvement not only thermodynamically but also in reducing environmental impacts. Given that the exergetic performance shows better results, the social ecological factors (1.52–1.78) are comparable to some studies [20,31,75]. Although the environmental impacts of the proposed systems are lower than those of current systems, their range appears appropriate compared with other values.

5. Conclusions

This study presents a poly-generation renewable system that uses solar and biomass energy to produce electricity, heating, carbon dioxide, and hydrogen. The primary innovation of this analysis lies in the comprehensive integration of technical, economic, and environmental assessments for various green hydrogen production pathways utilizing decarbonized biomass gasification.
The key findings from the design and analysis are as follows:
-
Comprehensive dynamic simulations have been developed for CO2 capture and hydrogen production units.
-
The most significant exergy destruction occurs in the biomass conversion and syngas production processes, with the highest associated costs and environmental impacts seen in hydrogen production.
-
The dynamic results indicate that extending the operation of CSP systems enhances hydrogen production by reducing the levelized costs of electricity, hydrogen, and total products.
Overall, the optimized system achieves an exergy efficiency of 48.67%, an exergoeconomic factor of 80.65%, and an exergoenvironmental factor of 77.89%. It can produce 51.5 MW of electricity, 23.3 MW of heat, and 8334.4 kg/h of hydrogen through the supercritical water gasification of 87,156.4 kg/h of biomass.
In terms of thermodynamic performance, the models in the current study achieve exergy efficiencies of 34.2% to 44%. Model 4, with the highest exergy efficiency of 44%, also has the lowest CO2 emissions (26.56 kgCO2/MWh) compared with other models in this study and those in previous works.
The proposed systems 2 and 3, which utilize chemical looping reforming processes, demonstrate superior performance in terms of the LCOE, with values of 49.2 USD/MWh and 55.4 USD/MWh, respectively. Moreover, the proposed models have proven to be the most cost-effective processes to date, with LCOHs ranging from 10.7 to 19.5 USD/MWh and a unit exergy cost of hydrogen between 3.3 USD/GJ and 5.28 USD/GJ.
The exergoenvironmental impact factor ranged from 66.2% to 73.9% with an environmental impact rate of 5.4–7.1 Pts/MWh, indicating that exergy and environmental aspects are comparable in terms of their environmental impact, consistent with industrial conditions. An extensive comparison demonstrated that the designed hybrid system with solar energy integration can outperform other similar plants in terms of exergoeconomic and exergoenvironmental performance and, sometimes, even be competitive with them.
Finally, it is important to note that the systems studied have high irreversibility, necessitating analysis of the units and components in each model. To make solar-driven hybrid power plants more competitive, it is essential to address issues relating to improving process efficiency and reducing the costs associated with CSP plants (heliostats, solar receivers, etc.). Solar concentrators contribute substantially to the investment, while receivers must endure high temperatures and corrosive chemicals.

Author Contributions

Conceptualization, K.A. and Z.A.; methodology, K.A. and Z.A.; investigation, Z.A.; writing—original draft preparation, Z.A.; visualization, Z.A.; project administration, Z.A. and T.-D.X.; writing—review and editing, K.A. and T.-D.X.; resources, T.-D.X.; supervision, T.-D.X.; funding acquisition, T.-D.X. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

The data used in this study are available upon request from the corresponding author. The data are not publicly available due to restrictions set by each plant owner.

Acknowledgments

The authors express their gratitude to the JDS project for providing funding to implement this study.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviation

ABScarbon dioxide capture unitOCoxygen carrier
APacidification potentialO&Moperational and maintenance cost
ARair reactorOPEXoperational and maintenance expenditure
ASUair separation unitPECpurchased equipment cost
CAPEXcapital expenditure costPSApressure swing adsorption
CCcombustion chamber PVphotovoltaic
CHPcombined heat and powerRErenewable energy
CLRchemical looping reformingSCWGsupercritical water gasification
CompcompressorSEFsocial ecological factor
CSPconcentrated solar powerSERsorption-enhanced reforming
CondcondenserSPECOspecific exergy costing method
DNIdirect normal irradiationSRsteam reactor
EFecological effect factor STsteam turbine
ESIexergetic sustainability indexTESthermal energy storage
ExpexpanderSubscript
FRfuel reactorCcarbon element
GengeneratorCO2carbon dioxide
GTgas turbineHhydrogen element
GWPglobal warming potentialH2hydrogen
HEheat exchangerH2Owater
HRSGheat-recovery steam generator Nnitrogen element
IIexergoenvironmental impact improvementN2nitrogen
LCAlife cycle assessmentOoxygen element
LCOElevelized cost of electricityO2oxygen
LCOHlevelized cost of hydrogenSsulfur element
LCOPlevelized cost of productH2Shydrogen sulfide
MEAmonoethanolamine

Appendix A

Appendix A.1. Thermodynamic Modeling

Thermodynamic modeling can be performed as follows:
Mass balance:
m ˙ i = m ˙ e
The first law of thermodynamics can be expressed uniformly as shown in Equation (1), which serves as the foundation for the thermodynamic modeling of the energy system. The goal of the thermodynamic analyses is to determine the physical properties of each substance based on its operating conditions.
There are some methods in systems thermal studies containing energy and exergy investigations. Applying the second law of thermodynamics is a suitable solution to analyze the entropy generation of energy systems and their irreversibility and to measure the deviation of energy systems in their operational state from their ideal operational state.
The exergy destruction of each component that represents its irreversibility is defined by the following equation [27]:
E ˙ D = i E ˙ F e E ˙ P
y D = E ˙ D E ˙ F
In Equations (A2) and (A3), E ˙ F , E ˙ P , and y D are the fuel exergy, the product exergy, and the exergy destruction yield of the system, respectively.
Exergy analysis compares the effectiveness of a low-grade energy system to an ideal one based on the maximum energy used relative to the ambient temperature [2]. It considers the solar and biomass exergy input, as well as the exergy generated by heating, power, and H2 production.

Appendix A.2. Economic Modeling

Regarding the economic indicators of the assessed biomass and solar-based renewable H2 production concepts with decarbonization features, investment and operating costs were calculated considering the purchased equipment costs of the components, e.g., biomass processing, oxygen supply unit, gasification island, syngas treatment line, acid gas removal, hydrogen purification, carbon processing, etc.
DC has installation costs such as the labor cost for installing the equipment, piping, electrical, yard improvement, and building. In the CSP plant, the purchase cost of the heliostats, central tower, and thermal energy storage tank is already included in the installation cost [40]. A further breakdown of the DC is shown in Table 2. The installation factors for the biomass conversion and syngas production plants were taken from the literature [9,40,51]. The indirect cost (IDC) accounts for the engineering and construction, which is equal to 10% of the total PEC. Typically for high-risk projects, to take care of unexpected expenses, a contingency fund of 15% of the DIC is considered [40,51]. To estimate the CCont, a land utilization factor ( f L a n d ) of 0.35 and a unit land price ( U P L a n d ) of 1 USD/m2 were assumed.
In the sub-systems, the capital cost ( C i ) for a certain production capacity ( Q i ) was assessed using the reference cost data ( C r e f ), reference sizes ( Q r e f ), and an exponent ( M ), as shown in Equation (A4). Reference data for the CAPEX calculation used the Chemical Engineering Plant Cost Index (CEPCI) [51], updated to the year 2023 according to Table A2:
C i = C r e f Q i Q r e f m
The original equipment cost ( Z k ) is summarized in Table A3. The conversion equation for the component cost in different years was proposed in the literature [9] as shown in Equation (A5).
C i , 2023 = C r e f C E P C I 2023 C E P C I r e f ,   y e a r
The capital annuity cost (CAnn) is the yearly installment of the initial investment plus the annual interest on the initial investment to be paid over the lifetime of the plant. The CAnn can be determined by multiplying the CAPEX with the capital recovery factor (CRF). The CRF is calculated according to Equation (A6):
C R F = i r 1 + i r n 1 + i r n 1
where i r and n denote the interest rate and the economic life of the system.
The annual operation and maintenance cost for the heliostat field was 9.33 USD/m2/y [40] and 3% of the purchase equipment cost for the rest of the equipment [40]. Property, taxes, and insurance costs ( C P T I ) were estimated to be 2% of the CAPEX.
Economic aspects play a major role in the feasibility analysis and profitability evaluation of a novel power plant [4]. The LCOE is the other key economic parameter that is valuable for comparing the cost of different systems per generated power unit [35].
Lastly, the profitability of the power plant over its life cycle was determined using net present values (NPVs) [76,77]. To determine the present value, all future cash flows are estimated using the interest rate ( i r ), and the power system becomes feasible only when the NPV is positive. The NPV calculation is shown in Equation (A7):
N P V = A I × 1 + r n 1 r 1 + r n T C
Here, A I is the annual income of the system, and T C responds to the total investment cost.
The LCOE was computed using Equation (12):
L C O E = A n n u a l i z e d   C A P E X + y = 1 N O P E X y = 1 N W ˙ n e t  
For the feasibility analysis of an innovative H2 production plant, the LCOH is a broadly used indicator for the profitability assessment of power systems, while it also helps engineers to compare different power production systems. This parameter is defined as:
L C O H = A n n u a l i z e d   C A P E X + y = 1 N O P E X y = 1 N m ˙ L H V H 2 + W ˙ n e t  

Appendix A.3. Exergoeconomic Modeling

The maximum accessible work achieved through a process that brings the system into equilibrium with its surroundings is known as exergy [32]. Exergoeconomics is a tool used to evaluate thermodynamic systems from a cost perspective that is related to exergy flows [20].
As assessed in the previous sub-section, once a positive NPV is reached, it can be concluded that the energy system is profitable. Exergoeconomic analysis is meaningful only if the energy system is expected to be profitable within a 10-year period [9]. The purpose of exergoeconomics is to identify which part of the plant contributes to the high costs of the power system.
The exergoeconomic investigation simultaneously considers both thermodynamic and economic characteristics to evaluate the overall cost rate of the integrated system. Exergoeconomic analysis is a complementary assessment that deepens the understanding of the system performance.
Z ˙ k represents the investment cost rate, and it includes the capital cost rate and the operating and maintenance cost, which is obtained as:
Z ˙ k = Z ˙ k C I + Z ˙ k O M = Z k C R F φ N
where φ shows the maintenance factor, which is assumed to be 1.06.
A thorough cost-balance equation will be established following the calculations in Equations (A11)–(A14) below:
c F , k = C ˙ F , k E ˙ F , k
c P , k = C ˙ P , k E ˙ P , k
C ˙ D , t o t = c F , t o t E ˙ D , t o t = c F , t o t k E ˙ D , k
c P ,   s y s t e m = i = 1 k Z ˙ i + i = 1 k c F u e l E ˙ F u e l i = 1 k E ˙ P r o d u c t
where C ˙ L denotes the cost rate for exergy loss and can be expressed as:
C ˙ L , k = c F E ˙ L , k
Finally, the summation of cost rates ( C ˙ t o t ), which includes the cost of investment, exergy destruction, environmental impacts, and biomass, is computed as follows [5]:
C ˙ t o t = Z ˙ k + C ˙ D , k + C ˙ e n v + C ˙ b i o m a s s
C ˙ b i o m a s s = c b i o m a s s m ˙ b i o m a s s E ˙ b i o m a s s
C ˙ e n v = m C O c C O + m C O 2 c C O 2
where c C O and c C O 2 are the associated damage costs of CO and CO2 emissions, which are 0.02 USD/kg and 0.024 USD/kg, respectively [6,17,36].
Finally, the following index is utilized to evaluate the extent of enhancement in the system’s environmental performance through decarbonization [70]:
λ = C ˙ e n v E ˙ P

Appendix A.4. Environmental Modeling

In addition to its techno-economic feasibility, the environmental impact of energy systems is another issue that should be addressed, as the global warming phenomenon caused by CO2 from greenhouse gas emissions is one of the most important international challenges. Although fossil fuel-fired power remains a predominant source of energy and heat in many countries, the shift toward low-carbon renewable energy technologies is imperative for reaching carbon neutrality as soon as possible. Therefore, this parameter is considered here as a proper criterion to estimate the system’s environmental performance.
The biomass gasification system integrated with solar energy, as described in this study, achieves reduced CO2 emissions through carbon capture. Additionally, biomass is a carbon-neutral RE source with minimal sulfur and nitrogen content, which results in a negligible environmental impact compared with CO2. This research evaluates and compares the environmental impact of power generation systems by calculating the CO2 and CO emissions from syngas combustion.
H2 has significant potential to substantially reduce CO2 emissions when produced from sustainable, renewable, and clean energy sources.
LCA offers a holistic perspective for a complete environmental assessment following a standardized method that uses goal and scope definitions, functional unit and system boundaries, inventory data acquisition, and impact assessment. The GWP measures CO2 emissions in terms of kg CO2 equivalents, while the AP evaluates changes in acidity by indicating the extent of sulfur dioxide (SO2) release into soil and water in g SO2 equivalents.
These sustainability indexes vary, relying on the system’s exergy efficiency in relation to ambient temperature. A lower index suggests a smaller contribution to sustainability, whereas a higher index points to greater environmental benefits [31]. Moreover, utilization of renewable energy within the system significantly reduces CO2 emissions compared with a system powered solely by electricity.
D e c l i n e   i n   C O 2   e m i s s i o n s = E ˙ n e t W
where is the C O 2 emission factor of electricity (0.98 kgCO2/kWh), and E ˙ n e t W is the net exergy output by the operated system.
The Aspen Plus® [42] and Ebsilon® Professional [45] models of the integrated systems in this study are utilized to develop custom hydrogen and power production modules for each design option based on our simulation results. The energy input and GHG emission credits associated with the electricity co-product are included in the models.
In the LCA analysis, the environmental impact rate of each piece of equipment is determined by multiplying its weight by the environmental impact coefficient per unit mass. Table A1 provides the equipment weight of the system. The environmental impact of the equipment per unit mass is calculated using the Eco-indicator 99, as referenced in designated studies for the specific equipment [7,27]. In Table A4, more detailed information from the LCA of the components and how to calculate their environmental impacts are described.
Once the environmental rate of each piece of equipment is determined, the environmental balance for the equipment can be established. By using auxiliary equations and solving the system of balances, the environmental parameters of each equipment can be determined. With these values, additional environmental metrics such as the environmental destruction rate and environmental factors can be calculated.
In addition to the previously mentioned points, the system’s operation requires utility services, namely fuel, electricity, heating, water, and so on), which also contribute to the environmental impact, denoted as Y U and calculated using Equation (A21). The sum of the environmental impact of material and the utility gives the total environmental impact of the LCA system ( Y t o t ), as shown in Equation (A22):
Y U = w U m U
Y t o t = Y M + Y U
The study employs a gate-to-gate life cycle costing (LCC) together with life cycle assessment (LCA) methodology, which is guided by the ISO 14040–44 and ISO15686-5 standards, respectively [68,78,79]. To maintain consistency with the LCA system boundary, a steady-state LCC assessment model was applied, which accounts for both financial and environmental costs. H2 storage was not considered, as the production rate remains unchanged in both cases, and it is assumed that the LCA and LCC of H2 storage would be identical. By excluding hydrogen storage, the analysis specifically targets the H2 production process, allowing for a more focused evaluation of its environmental and economic performance.

Appendix A.5. Exergoenvironmental Modeling

Given that exergy and economic assessments of solar- and biomass-based hybrid systems are often insufficient, this system has also been evaluated from an exergoenvironmental perspective. In the sub-section, the environmental performance of the proposed system has been investigated using exergoenvironmental analysis.
This analysis quantifies the environmental damage ( B ˙ i ) resulting from the resource use that generates the flows. The definition of each flow’s exergoenvironmental impact is similar to that used for calculating the exergoeconomic and exergy costs. The second proposition is adapted to include the unit exergoenvironmental impacts of the incoming resources, represented by their unit CO2 emissions [16]. The exergoenvironmental impacts of the fuel and product ( B ˙ F , k and B ˙ P , k ) are determined using Equations (A23) and (A24):
b F , k = B ˙ F , k E ˙ F , k
b P , k = B ˙ P , k E ˙ P , k
This analysis is derived from the life cycle approach. The LCA accounts for the entire environmental impact of equipment throughout its life cycle, encompassing manufacturing, transportation, operation, and disposal. It evaluates all detrimental environmental effects during the equipment’s lifespan [7]. As outlined in the previous sub-section, the LCA integrates these harmful effects into a single comprehensive unit using the Eco-indicator 99, making it the most thorough environmental analysis compared with other available methods.
Exergoenvironmental analysis examines the system’s environmental impact by analyzing the correlation between the exergy destruction rate and its environmental consequences. When this correlation increases, the system’s performance deteriorates, necessitating the supply of more harmful fuel. During operation, solar- and biomass-based systems do not directly pollute the environment. However, increasing the system’s size requires more energy consumption for components and products, which can indirectly harm the environment [58].
The ecological effect factor ( θ e f ) and the exergoenvironmental impact factor ( Ω i f ) are related to the exergy destruction and exergy efficiency of the system, as shown in Equations (A25) and (A26):
θ e f = 1 ε s y s
Ω i f = E ˙ D , t o t E ˙ F
where E ˙ D , t o t and E ˙ F are the total exergy destruction and the total exergy fuel of the system, respectively.
The exergoenvironmental impact factor illustrates the environmental damage resulting from system inefficiencies. By reducing the total exergy destruction within the system, the extent of environmental harm is diminished.
Exergoenvironmental damage effectiveness index:
Ω d e i = Ω i f × 100 % ε s y s
The Ω d e i index specifies if the system harms the environment or not. A favorable system has a lower damage effectiveness index.

Appendix B

Figure A1. Framework of the LCA [22].
Figure A1. Framework of the LCA [22].
Energies 17 04264 g0a1
Figure A2. Boundaries of the hydrogen and power production process.
Figure A2. Boundaries of the hydrogen and power production process.
Energies 17 04264 g0a2
Table A1. Calculation of the cost and weight of the equipment in the system.
Table A1. Calculation of the cost and weight of the equipment in the system.
ComponentCost EquationRef.CEPCIWeight Equation [7,27,80]
Solar tower P E C r e c = A r e c ( 79 T r e c 42,000 ) [22]381.7 W e i g h t R e c = 0.019 × Q r e c
Heliostat P E C h e l i o = 100 A h e l i o N h e l i o [40]567.3 W e i g h t H e l i o = 6 × N h e l i o
Solar HE P E C S H E = 12,000 A h e l i o 100 0.6 [80]541.7 W e i g h t S H E = 2.14 × ( Q ˙   S H E ) 0.7
TES P E C T S E = 113 V T E S [20]596.2 W e i g h t R e a c t o r = 8.050 × V R e c t o r
Water HE P E C W H E = 6.534 Q H E [9]468.2 W e i g h t H E = 2.14 × ( Q ˙   H E 1000 ) 0.7
HEX P E C H E X = 235 Q ˙ H E X 0.75 [27]567.5 W e i g h t H E = 2.14 × ( Q ˙   H E X 1000 ) 0.7
Gasifier P E C g a s i f i e r = 1600 ( m ˙ d r y , b i o m a s s ) 0.67 [4]525.4 W e i g h t = 8.050 × V R e c t o r
Reformer P E C R e f = 4.78 × 10 6 m ˙ C H 4 , i n 44,910 0.67 [34]556.8 W e i g h t R e f = 8.050 × V R e f
Calciner P E C C a l = 0.59 × 10 6 m ˙ C a C O 3 , i n 59,875 0.67 [34]556.8 W e i g h t C a l = 8.050 × V C a l
Carbonator P E C C a r = 0.59 × 10 6 m ˙ C O 2 , o u t 59,875 0.67 [34]556.8 W e i g h t C a r = 8.050 × V C a r
Fuel reactor P E C F R = 2.15 × 10 6 m ˙ F e 2 O 3 , i n 1,009,800 0.67 [34]541.7 W e i g h t R e a c t o r = 8.050 × V R e a c t o r
Steam reactor P E C S R = 0.63 × 10 6 m ˙ s t e a m , i n 213,840 0.67 [34]541.7 W e i g h t R e a c t o r = 8.050 × V R e a c t o r
Air reactor P E C A R = 0.63 × 10 6 m ˙ a i r , i n 213,840 0.67 [34]541.7 W e i g h t R e a c t o r = 8.050 × V R e a c t o r
Water pump P E C W P = 1120 ( W ˙ W P ) 0.8
P E C S C W , P = 3450 ( W ˙ W P ) 0.71
[32]
[4]
607.5
596.2
W e i g h t W P = 0.0061 × W ˙ P 0.95
W e i g h t W P = 0.175 ln W
Solar pump P E C S P = 3540 ( W ˙ S P ) 0.67 [24]607.5 W e i g h t S P = 0.0061 × W ˙ P 0.95
Air compressor P E C A C o m p = 71.1 m ˙ a i r 1 0.9 η A C o m p P e P i ln P e P i [70]567.5 W e i g h t A c o m p = 4.9 × W ˙ G T 0.73
Cooler P E C c o o l e r = 12,000 m ˙ 1 h 1 h 2 100 U T L M 0.6 [81]607.5 W e i g h t H E = 2.14 × ( Q ˙   H E 1000 ) 0.7
Flash P E C s e p = 0 [21]596.2 W e i g h t = 8.050 × V f l a s h
CC P E C C C = 46.08 0.995 ( P e / P i ) 1 + e x p 0.018 T e 26.4 [80]567.5
GT P E C G T = 479.34 m ˙ g a s 3600 0.92 η G T ln P i P e 1 + e x p 0.036 T i 54.4 [6]402 W e i g h t S T = 4.9 × W ˙ G T 0.73
Generator P E C G e n = 26.18 W ˙ G e n 0.95 [21]607.5 W e i g h t W P = 0.125 ln W 0.0415
HRSG P E C H R S G = 6570 Q ˙ E C O T E C O 0.8 + Q ˙ E V A T E V A 0.8 + Q ˙ S H T S H 0.8 + 21,276 m ˙ w a t e r + 1184.4 m ˙ f g 1.2 [27]567.3 W e i g h t H R S G = 8.42 Q ˙   S H 1000 0.87 + 13.91 Q ˙   E V A 1000 0.68 + 2.989 × Q ˙   E C O 1000 0.98
ST P E C S T = 2210 × W ˙ S T 0.7 [80]541.7 W e i g h t S T = 4.9 × W ˙ S T 0.73
Compressor P E C C o m p = 91,562 W ˙ C o m p 455 0.67   [32]402 W e i g h t c o m p = 4.9 × W ˙ c o m p 0.73
Expander P E C E X P = 479.34 m ˙ g a s 3600 0.92 η E X P ln P i P e 1 + e x p 0.036 T i 54.4 [6]402 W e i g h t e x p = 4.9 × W 0.73
Condenser P E C C o n d = 10,000 + 324 A C o n d 0.91 [82]556.8 W e i g h t C o n d = 0.073 × Q ˙ C o n d 1000 0.99
Valve P E C V a l v e = 114.5 m ˙ f l o w [70]607.5
Boiler P E C B o i l e r = 130 A B o i l e r 0.093 0.78 [32]499.6 W e i g h t b o i l e r = Q ˙   H E 1000 0.7
Cyclone P E C C y c l = 0.31 × 10 6 m C a C O 3 47.85 × 3.6 × 10 6 0.8 [51]567.3 W e i g h t c y c l = 8.050 × V C o l u m n
Dryer P E C D r y e r = 0.01 × 10 6 m i 119,047 0.67 [51]395.6 W e i g h t D r y e r = 8.050 × V C o l u m n
ASU P E C A S U = 29.5 × 10 6 24 m O 2 432 × 10 3 0.65 [54]521.9 W e i g h t A S U = 8.050 × V C o l u m n
WGS, kmol/h P E C W G S = 17 × 10 6 m i 33,415 0.6 [83]541.7 W e i g h t W G S = 8.050 × V C o l u m n
H2S remover P E C H 2 S = 46.24 × 10 6 m H 2 S 2400 0.75 [52] 1402 W e i g h t H 2 S = 8.050 × V C o l u m n
PSA, kmol/h P E C P S A = 38.3 × 10 6 m i 17,069 0.6 [53] 2525.4 W e i g h t P S A = 8.050 × V C o l u m n
CO2 capture, kmol/h P E C C O 2 c a p t u r e = 18.1 × 10 6 m C O 2 1227 0.65 [54]521.9 W e i g h t A B S = 8.050 × V C o l u m n
1 Average exchange rate: EUR 1 = USD 1.13 in 2003. 2 Average exchange rate: EUR 1 = USD 1.37 in 2007.
Table A2. CEPCI in different years [51].
Table A2. CEPCI in different years [51].
YearCEPCIYearCEPCIYearCEPCIYearCEPCI
1996381.72007525.42013567.32019607.5
2002395.62008575.42014576.12020603.1
2003402.02009521.92015556.82021708.8
2004444.22010550.82016541.72022816
2005468.22011585.72017567.52023800.8
2006499.62012584.62018603.1
Table A3. Total purchased equipment cost of the components.
Table A3. Total purchased equipment cost of the components.
UnitModel 1Model 2Model 3Model 4UnitModel 1Model 2Model 3Model 4
Receiver 23.3723.7812.7317.23GT0.940.690.380.93
Heliostat62.3266.7069.0370.86HRSG0.000.000.000.00
S-HE2.732.842.902.95ST0.000.000.000.00
TES2.442.434.284.08Gen0.000.000.510.46
S-pump0.590.590.690.87Cooler0.760.002.571.74
Dryer0.020.020.020.00ABS51.980.0048.3849.72
Gasifier5.212.610.004.98Cond0.000.000.000.00
ASU31.640.000.000.00Boiler0.000.000.000.00
W-HE0.380.561.351.47FR0.000.003.210.00
WGS10.170.008.9010.88SR0.000.001.570.00
HEX0.460.000.101.82AR0.000.000.850.00
W-pump0.120.110.160.42Cycl0.000.010.030.00
H2S15.6513.9212.7513.16Exp0.000.006.789.95
PSA38.7735.970.0038.49Scomp0.000.001.762.64
CC0.060.060.080.06Calciner0.001.990.000.00
AComp6.327.337.305.70Reformer0.000.000.004.65
Total cost, USD (in millions) 253.9159.59186.32243.07
Table A4. Environmental impacts of some components based on the LCA [7,27,84].
Table A4. Environmental impacts of some components based on the LCA [7,27,84].
ComponentMaterial CompositionEco-Indicator 99 (mPts/kg)Material (mPts/kg)Process (mPts/kg)Disposal (mPts/kg)Total (mPts/kg)
Air CompressorSteel 33.33%8613011.7−70.071.7
Steel low-alloy 44.5%110
Cast iron 22.22%240
Combustion ChamberSteel 33.34%8663520.0−70.0585
Steel high-alloy 66.66%910
TurbinesSteel 25%8670412.1−70.0646
Steel high-alloy 75%910
PumpsSteel 35%8618616.9−70.0132.8
Cast iron 65%240
GasifierSteel 100%868612.1−70.028.0
DeaeratorSteel 100%868612.1−70.028.0
Heat ExchangersSteel 25%8669612.1−70.028.0
Steel high-alloy 75%910
Super HeaterSteal 25%8670412.1−70.0646
Steel high-alloy 75%910
ColumnsSteel 100%868612.1−70.028.0
ReactorSteel 100%868612.1−70.028.0
Collectors Steel 98%86857.3−69.023.2
Glass 2%58
Steam turbineStainless steel 30%31029980−70309
Steel high-alloy 4.3%910
Steel low-alloy 50%110
Chromium steel 11%970
Copper 0.3%1400
Aluminum, prod mix 1.7%60
Table A5. Comparison of the performance between the presented models and the models proposed in past studies.
Table A5. Comparison of the performance between the presented models and the models proposed in past studies.
SystemUseful OutputMain OutcomesRef.
Solar-driven hydrogen and power production system based on biomass gasificationElectricity, heating, CO2, and hydrogenExergy efficiency of 41.62%; Exergoeconomic factor of 69.4%; Total cost rate of 7603.5 USD/h; LCOE of 83.4 USD/MWh; LCOH of 13.9 USD/MWh; LCOP of 10.6 USD/MWh; Unit exergy cost of hydrogen of 5.28 USD/GJ; LCC of 1.67 USD/kgH2; Specific CO2 emission of 28.9 kgCO2/MW; Total environmental impact rate of 1899.7 Pts/h; Exergoenvironmental factor of 67.6%; Social ecological factor of 1.71This study
Solar-driven hydrogen and power production system based on sorption-enhanced reforming Electricity, heating, CO2, and hydrogenExergy efficiency of 37.98%; Exergoeconomic factor of 59.2%; Total cost rate of 6076.8 USD/h; LCOE of 49.2 USD/MWh; LCOH of 10.7 USD/MWh; LCOP of 7.8 USD/MWh; Unit exergy cost of hydrogen of 3.3 USD/GJ; LCC of 1.23 USD/kgH2; Specific CO2 emission of 93.82 kgCO2/MW; Total environmental impact rate of 1942.7 Pts/h; Exergoenvironmental factor of 66.2%; Social ecological factor of 1.61This study
Solar-driven hydrogen and power production system based on iron chemical reformingElectricity, heating, CO2, and hydrogenExergy efficiency of 34.22%; Exergoeconomic factor of 67.9%; Total cost rate of 6941 USD/h; LCOE of 55.4 USD/MWh; LCOH of 19.5 USD/MWh; LCOP of 12.5 USD/MWh; Unit exergy cost of hydrogen of 4.69 USD/GJ; LCC of 2.24 USD/kgH2; Specific CO2 emission of 32.76 kgCO2/MW; Total environmental impact rate of 1983 Pts/h; Exergoenvironmental factor of 70.7%; Social ecological factor of 1.52This study
Solar-driven hydrogen and power production system based on supercritical water gasificationElectricity, heating, CO2, and hydrogenExergy efficiency of 43.99%; Exergoeconomic factor of 76.9%; Total cost rate of 7981.4 USD/h; LCOE of 88.4 USD/MWh; LCOH of 16.1 USD/MWh; LCOP of 12.1 USD/MWh; Unit exergy cost of hydrogen of 5.19 USD/GJ; LCC of 1.95 USD/kgH2; Specific CO2 emission of 26.56 kgCO2/MW; Total environmental impact rate of 2176 Pts/h; Exergoenvironmental factor of 73.9%; Social ecological factor of 1.78This study
Hybrid biomass–solar-driven triple combined cycleElectricity and hydrogenExergy efficiency of 30.44%; LCOE of 61.37 USD/MWh; CO2 emission of 0.4579 kg/kWh[4]
Multi-generation system with solar power towerElectricity, cooling, heating, and hydrogenExergy efficiency of 43.11%; Total cost rate of 7799 USD/h; Unit cost of multi-generation of 8.26 USD/GJ; H2 unit cost of 45 USD/GJ; Environmental cost rate of 20%[5]
Biogas-driven poly-generation plantElectricity, fresh water, cooling, and hydrogenExergy efficiency of 31.69%; Product unit cost of 20.1 USD/GJ; Exeregoeconomic factor of 38.64%[24]
Solar–biomass integrated energy systemElectricity, hot water, and liquid hydrogenExergy efficiency of 11.06%; Total cost rate of 320 USD/h; LCOE of 40.16 USD/MWh; CO2 emission of 0.39 t/MWh[22]
Multi-generation energy system driven by integrated RESElectricity and hydrogen fuelExergy efficiency of 35.9%; Product unit cost of 36.95 USD/h; LCOE of 98 USD/MWh; LCOH of 170.11 USD/MWh[26]
Poly-generation system based on the integration of biomass and solarElectricity, fresh water, and hydrogen Exergy efficiency of 32.01%; LCOE of 14 USD/MWh; LCA of 0.002 Pts/kWh[27]
Solar-driven steam gasification of biomassElectricity and hydrogenHydrogen minimum price of 2.56 EUR/kgH2[30] 1
Geothermal–solar–wind micro-multi-energy systemElectricity, hydrogen, and hot waterExergy efficiency of 88.31%; Total product unit cost of 1859 USD/GJ; LCOE of 45 USD/MWh; LCOH of 865.89USD/MWh.[27]
Solar–biomass cascade ORC systemCooling, power, and hydrogenExergy efficiency of 11%; Total cost of 42–75 USD/h; Reduced CO2 emission from 364 kg/h to 43 kg/h [31]
Solar–biomass multi-generation systemElectricity, hydrogen, and coolingExergy efficiency of 43.12%; Hydrogen unit cost of 39.49 USD/GJ; Product unit cost of 8.35 USD/GJ (30 USD/MWh)[70]
Cogeneration system based on biomass and solar Electricity, hot water, and hydrogenExergy efficiency of 30.91%; Total cost rate of 633 USD/h; Unit product cost of 16.77 USD/MWh; LCOE of 21.08 USD/MWh[35]
Decarbonized green hydrogen production using biomass gasification systemHydrogen and CO2LCOH of 71.47–78.66 USD/MWh[52] 2
Multi-generation plant powered by geothermal energyElectricity, fresh water, hydrogen, and coolingUnit cost of 124 USD/GJ; Exergy efficiency of 24.4%; Exergoeconomic factor of 38.85%; Exergy stability factor of 0.6[58]
Biomass air–steamgasification with methane co-feedingSyngasExergy efficiency of 71.8%; Unit hydrogen cost of 2.69 USD/kg. Unit exergy cost of hydrogen of 18.89 USD/GJ[71]
Decarbonized hydrogen production based on direct biogas conversion using thermo-chemical looping cyclesElectricity, CO2, and hydrogenCumulative energy efficiency of 68.56–73.35%LCOH of 33 EUR/MWh (34.7 USD/MWh); CO2 specific emissions of 0.4–159.3 kg/MWh[53]
Solar, geothermal, and biomass multi-generation Power, hydrogen, heating, cooling, and fresh waterExergy efficiency of 27.09%; Unit product cost of 21.79 USD/GJ; Unit cost of hydrogen of 15.74 USD/GJ; Social ecological factor of 1.37[20]
Renewable energy-driven poly-generation systemElectricity, hydrogen, fuel, hot water, and chilled waterExergy efficiency of 32.3%; Unit exergy cost of hydrogen fuel of 4.29 USD/GJ[57]
Solar-powered setup Natural gas, electricity, and liquid hydrogenSecond-law efficiency of 17.05%; LCOH of 107.7 USD/MWh; CO2 emission reduction of 0.0711 kg/s[60]
Solar thermal-operated power cycleElectricity and hydrogenThermal efficiency of LCOH of 6.88 USD; SI of 1.65–1.88[75]
Integrated solar-powered electricity generation systemElectricityLCOE of 0.342 USD/kWh; Specific CO2 emissions of 102 kgCO2/MWh; Ecological effect factor of 1.05; Exergoenvironental sustainability index of 9.78[85]
Solar-based sorption-enhanced gasification co-generation systemElectricity, hydrogen, and CO2Exergy efficiency of 62.09%; LCOE of 60.1 USD/MWh; CO2 emission of co-generation and electricity production of 8.5 kgCO2/MWh and 38.37 kgCO2/MWh[61]
Solar–biomass drive integrated energy systemDesalinated water, hydrogen, and electricityExergy efficiency of 33.58%; LCOE of 13.75 USD/MWh; Total cost rate of 1889 USD/h; Environmental impact of 1.96 Pts/MWh; Total environmental impact rate of 0.274 mPts/h[66]
Multi-generation systemNG, fresh water, hydrogen, and electricityExergy efficiency of 20.7%; Cost rate of 142 USD/h; CO2 emission cost rate of 7 USD/h[86]
Fresnel collector solar-driven multi-generation plantElectricity, methanol, oxygen, fresh water, and hydrogenExergy efficiency of 27.5%; Total cost rate of 3378 USD/h; LCOE of 49.28 USD/MWh; Environmental impact of 294.1 mPts/s; LCA of 4.377 Pts/GJ[69]
1 Average exchange rates: EUR 1 = USD 1.14 in 2020 2 Average exchange rates: EUR 1 = USD 1.05 in 2022.

References

  1. Kang, J.N.; Wei, Y.M.; Liu, L.C.; Han, R.; Yu, B.Y.; Wang, J.W. Energy systems for climate change mitigation: A systematic review. Appl. Energy 2020, 263, 114602. [Google Scholar] [CrossRef]
  2. Liu, X.; Hong, H.; Zhang, H.; Cao, Y.; Qu, W.; Jin, H. Solar methanol by hybridizing natural gas chemical looping reforming with solar heat. Appl. Energy 2020, 277, 115521. [Google Scholar] [CrossRef]
  3. Sokrethya, S.; Aminov, Z.; Van Quan, N.; Xuan, T.D. Feasibility of 10 MW Biomass-Fired Power Plant Used Rice Straw in Cambodia. Energies 2023, 16, 651. [Google Scholar] [CrossRef]
  4. Cao, Y.; Dhahad, H.A.; Togun, H.; Anqi, A.E.; Farouk, N.; Farhang, B. A novel hybrid biomass-solar driven triple combined power cycle integrated with hydrogen production: Multi-objective optimization based on power cost and CO2 emission. Energy Convers. Manag. 2021, 234, 113910. [Google Scholar] [CrossRef]
  5. Zoghi, M.; Habibi, H.; Choubari, A.Y.; Ehyaei, M.A. Exergoeconomic and environmental analyses of a novel multi-generation system including five subsystems for efficient waste heat recovery of a regenerative gas turbine cycle with hybridization of solar power tower and biomass gasifier. Energy Convers. Manag. 2021, 228, 113702. [Google Scholar] [CrossRef]
  6. Cao, Y.; Dhahad, H.A.; Abo-Khalil, A.G.; Sharma, K.; Mohammed, A.H.; Anqi, A.E.; El-Shafay, A. Hydrogen production using solar energy and injection into a solid oxide fuel cell for CO2 emission reduction; Thermoeconomic assessment and tri-objective optimization. Sustain. Energy Technol. Assess. 2022, 50, 101767. [Google Scholar] [CrossRef]
  7. Rabeti, S.A.M.; Manesh, M.H.K.; Amidpour, M. An innovative optimal 4E solar-biomass waste polygeneration system for power, methanol, and freshwater production. J. Clean. Prod. 2023, 412, 137267. [Google Scholar] [CrossRef]
  8. AlNouss, A.; McKay, G.; Al-Ansari, T. Enhancing waste to hydrogen production through biomass feedstock blending: A techno-economic-environmental evaluation. Appl. Energy 2020, 266, 114885. [Google Scholar] [CrossRef]
  9. Wang, J.; Al-attab, K.A.; Heng, T.Y. Techno-economic and thermodynamic analysis of solid oxide fuel cell combined heat and power integrated with biomass gasification and solar assisted carbon capture and energy utilization system. Energy Convers. Manag. 2023, 280, 116762. [Google Scholar] [CrossRef]
  10. Alikulov, K.; Aminov, Z.; Anh, L.H.; Dang, X.T.; Kim, W. Effect of ammonia and hydrogen blends on the performance and emissions of an existing gas turbine unit. Int. J. Hydrogen Energy 2024, 61, 432–443. [Google Scholar] [CrossRef]
  11. Lamnatou, C.; Chemisana, D. Concentrating solar systems: Life Cycle Assessment (LCA) and environmental issues. Renew. Sustain. Energy Rev. 2017, 78, 916–932. [Google Scholar] [CrossRef]
  12. Amin, M.; Shah, H.H.; Fareed, A.G.; Khan, W.U.; Chung, E.; Zia, A.; Farooqi, Z.U.R.; Lee, C. Hydrogen production through renewable and non-renewable energy processes and their impact on climate change. Int. J. Hydrogen Energy 2022, 47, 33112–33134. [Google Scholar] [CrossRef]
  13. Aminov, Z.; Nakagoshi, N.; Xuan, T.D.; Higashi, O.; Alikulov, K. Evaluation of the energy efficiency of combined cycle gas turbine. Case study of Tashkent thermal power plant, Uzbekistan. Appl. Therm. Eng. 2016, 103, 501–509. [Google Scholar] [CrossRef]
  14. Matjanov, E.K.; Akhrorkhujaeva, Z.M. Modernization of Existing Thermal Power Plants into Hybrid Solar and Fuel-Fired Plants. Therm. Eng. 2023, 70, 831–840. [Google Scholar] [CrossRef]
  15. Aminov, Z.; Alikulov, K.; Anh, L.H.; Xuan, T.D. Thermodynamic evaluation of decarbonized power production based on solar energy integration. Appl. Therm. Eng. 2024, 255, 124020. [Google Scholar] [CrossRef]
  16. Daghsen, K.; Perez, A.P.; Lounissi, D.; Bouaziz, N. Exergy, exergoeconomic and exergoenvironmental assessments of experimental hybrid energy systems for hot water production to improve energy sustainability. Renew. Sustain. Energy Rev. 2023, 187, 113741. [Google Scholar] [CrossRef]
  17. Hanifi, K.; Javaherdeh, K.; Yari, M. Exergoeconomic and exergoenvironmental analysis and optimisation of the three configurations of CO2 transcritical cogeneration cycle using genetic algorithm. Int. J. Exergy 2016, 19, 395. [Google Scholar] [CrossRef]
  18. Qin, Z.; Tang, Y.; Zhang, Z.; Ma, X. Techno-economic-environmental analysis of coal-based methanol and power poly-generation system integrated with biomass co-gasification and solar based hydrogen addition. Energy Convers. Manag. 2021, 228, 113646. [Google Scholar] [CrossRef]
  19. Bejan, A.; Tsatsaronis, G.; Moran, M.J. Thermal Design and Optimization; Wiley: Hoboken, NJ, USA, 1996. [Google Scholar]
  20. Bozgeyik, A.; Altay, L.; Hepbasli, A. Energetic, exergetic, exergoeconomic, environmental and sustainability analyses of a solar, geothermal and biomass based novel multi-generation system for production of power, hydrogen, heating, cooling and fresh water. Process Saf. Environ. Prot. 2023, 177, 400–415. [Google Scholar] [CrossRef]
  21. Holagh, S.G.; Haghghi, M.A.; Mohammadi, Z.; Chitsaz, A. Exergoeconomic and environmental investigation of an innovative poly-generation plant driven by a solid oxide fuel cell for production of electricity, cooling, desalinated water, and hydrogen. Int. J. Energy Res. 2020, 44, 10126–10154. [Google Scholar] [CrossRef]
  22. Assareh, E.; Agarwal, N.; Paul, M.C.; Ahmadi, P.; Ghodrat, M.; Lee, M. Investigation and development of a novel solar-biomass integrated energy system for clean electricity and liquid hydrogen production. Therm. Sci. Eng. Prog. Progress. 2023, 42, 101925. [Google Scholar] [CrossRef]
  23. Moharramian, A.; Soltani, S.; Rosen, M.A.; Mahmoudi, S.M.S.; Bhattacharya, T. Modified exergy and modified exergoeconomic analyses of a solar based biomass co-fired cycle with hydrogen production. Energy 2019, 167, 715–729. [Google Scholar] [CrossRef]
  24. Bo, G.; Li, F.; Chauhan, B.S.; Ghaebi, H. Multi-objective optimization and 4E analyses to simultaneous production of electricity, freshwater, H2 and cooling based on biogas-steam reforming. Process Saf. Environ. Prot. 2023, 177, 1307–1320. [Google Scholar] [CrossRef]
  25. Rahbari, A.; Shirazi, A.; Venkataraman, M.B.; Pye, J. Solar fuels from supercritical water gasification of algae: Impacts of low-cost hydrogen on reformer configurations. Appl. Energy 2021, 288, 116620. [Google Scholar] [CrossRef]
  26. Ding, G.C.; Ji, P.; Geng, M.Y. Technical assessment of Multi-generation energy system driven by integrated renewable energy Sources: Energetic, exergetic and optimization approaches. Fuel 2022, 331, 125689. [Google Scholar] [CrossRef]
  27. Rabeti, S.A.M.; Manesh, M.H.K.; Amidpour, M. Techno-economic and environmental assessment of a novel polygeneration system based on integration of biomass air-steam gasification and solar parabolic trough collector. Sustain. Energy Technol. Assess. 2023, 56, 103030. [Google Scholar] [CrossRef]
  28. Onigbajumo, A.; Taghipour, A.; Ramirez, J.; Will, G.; Ong, T.-C.; Couperthwaite, S.; Steinberg, T.; Rainey, T. Techno-economic assessment of solar thermal and alternative energy integration in supercritical water gasification of microalgae. Energy Convers. Manag. 2021, 230, 113807. [Google Scholar] [CrossRef]
  29. Onigbajumo, A.; Taghipour, A.; Will, G.; Van, T.C.; Couperthwaite, S.; Steinberg, T.; Rainey, T. Effects of process-thermal configuration on energy, exergy, and thermo-economic performance of solar driven supercritical water gasification. Energy Convers. Manag. 2022, 251, 115002. [Google Scholar] [CrossRef]
  30. Boujjat, H.; Rodat, S.; Abanades, S. Techno-economic assessment of solar-driven steam gasification of biomass for large-scale hydrogen production. Processes 2021, 9, 462. [Google Scholar] [CrossRef]
  31. Karthikeyan, B.; Kumar, G.P. Thermoeconomic and optimization approaches for integrating cooling, power, and green hydrogen production in dairy plants with a novel solar-biomass cascade ORC system. Energy Convers. Manag. 2023, 295, 117645. [Google Scholar] [CrossRef]
  32. Bamisile, O.; Cai, D.; Adedeji, M.; Dagbasi, M.; Li, J.; Hu, Y.; Huang, Q. Thermo-enviro-exergoeconomic analysis and multi-objective optimization of a novel geothermal-solar-wind micro-multi-energy system for cleaner energy production. Process Saf. Environ. Prot. 2023, 170, 157–175. [Google Scholar] [CrossRef]
  33. Lin, H.; Wu, X.; Ayed, H.; Mouldi, A.; Abbas, S.Z.; Ebrahimi-Moghadam, A. A new biomass gasification driven hybrid system for power and liquid hydrogen cogeneration: Parametric study and multi-objective evolutionary optimization. Int. J. Hydrogen Energy 2022, 47, 26394–26407. [Google Scholar] [CrossRef]
  34. Osat, M.; Shojaati, F.; Osat, M. A solar-biomass system associated with CO2 capture, power generation and waste heat recovery for syngas production from rice straw and microalgae: Technological, energy, exergy, exergoeconomic and environmental assessments. Appl. Energy 2023, 340, 120999. [Google Scholar] [CrossRef]
  35. Teymouri, M.; Sadeghi, S.; Moghimi, M.; Ghandehariun, S. 3E analysis and optimization of an innovative cogeneration system based on biomass gasification and solar photovoltaic thermal plant. Energy 2021, 230, 120646. [Google Scholar] [CrossRef]
  36. Cao, Y.; Alsharif, S.; Attia, E.A.; Shamseldin, M.A.; Ibrahim, B.F. A conceptual process design towards CO2 emission reduction by integration of solar-based hydrogen production and injection into biomass-derived solid oxide fuel cell. Process Saf. Environ. Prot. 2022, 164, 164–176. [Google Scholar] [CrossRef]
  37. Rahbari, A.; Shirazi, A.; Venkataraman, M.B.; Pye, J. A solar fuel plant via supercritical water gasification integrated with Fischer–Tropsch synthesis: Steady-state modelling and techno-economic assessment. Energy Convers. Manag. 2019, 184, 636–648. [Google Scholar] [CrossRef]
  38. Shirazi, A.; Rahbari, A.; Asselineau, C.A.; Pye, J. A solar fuel plant via supercritical water gasification integrated with Fischer–Tropsch synthesis: System-level dynamic simulation and optimisation. Energy Convers. Manag. 2019, 192, 71–87. [Google Scholar] [CrossRef]
  39. Hashemian, N.; Noorpoor, A. Assessment and multi-criteria optimization of a solar and biomass-based multi-generation system: Thermodynamic, exergoeconomic and exergoenvironmental aspects. Energy Convers. Manag. 2019, 195, 788–797. [Google Scholar] [CrossRef]
  40. Patel, H.; Müller, F.; Maiti, P.; Maiti, S. Economic evaluation of solar-driven thermochemical conversion of empty cotton boll biomass to syngas and potassic fertilizer. Energy Convers. Manag. 2020, 209, 112631. [Google Scholar] [CrossRef]
  41. Sohn, G.; Ryu, J.Y.; Park, H.; Park, S. Techno-economic analysis of oxy-fuel power plant for coal and biomass combined with a power-to-gas plant. Energy Sustain. Dev. 2021, 64, 47–58. [Google Scholar] [CrossRef]
  42. Aspen Technology Inc. Aspen Plus® Process Simulation Software. 2024. Available online: https://www.aspentech.com/en/products/engineering/aspen-plus (accessed on 6 August 2024).
  43. Xiao, Z.; Zhang, C.; Li, P.; Wang, D.; Zhang, X.; Wang, L.; Zou, J.-J.; Li, G. Engineering oxygen vacancies on Tb-doped ceria supported Pt catalyst for hydrogen production through steam reforming of long-chain hydrocarbon fuels. Chin. J. Chem. Eng. 2024, 68, 181–192. [Google Scholar] [CrossRef]
  44. Xiao, Z.; Li, P.; Zhang, H.; Zhang, S.; Tan, X.; Ye, F.; Gu, J.; Zou, J.-J.; Wang, D. A comprehensive review on photo-thermal co-catalytic reduction of CO2 to value-added chemicals. Fuel 2024, 362, 130906. [Google Scholar] [CrossRef]
  45. Systems Engineering Study Group. System Technologies. Ebsilon® Professional Process Simulation Software. 2024. Available online: www.ebsilon.com (accessed on 6 August 2024).
  46. Nakyai, T.; Saebea, D. Exergoeconomic comparison of syngas production from biomass, coal, and natural gas for dimethyl ether synthesis in single-step and two-step processes. J. Clean. Prod. 2019, 241, 118334. [Google Scholar] [CrossRef]
  47. Remund, J.; Müller, S.; Schmutz, M.; Barsotti, D.; Graf, P.; Cattin, R. Meteorological Database. Available online: http://meteonorm.com/en/meteonorm-documents (accessed on 2 February 2024).
  48. Joint Stock Company “Thermal Power Plants”. Report: Industrial Indicators. Available online: https://tpp.uz/en/page/industrial-indicators (accessed on 6 August 2024).
  49. Gobio-Thomas, L.B.; Darwish, M.; Stojceska, V. Environmental impacts of solar thermal power plants used in industrial supply chains. Therm. Sci. Eng. Prog. 2023, 38, 101670. [Google Scholar] [CrossRef]
  50. Rahman, A.; Farrok, O.; Haque, M.M. Environmental impact of renewable energy source based electrical power plants: Solar, wind, hydroelectric, biomass, geothermal, tidal, ocean, and osmotic. Renew. Sustain. Energy Rev. 2022, 161, 112279. [Google Scholar] [CrossRef]
  51. Osat, M.; Shojaati, F.; Osat, M. Techno-economic assessment of butanol and pentanol productions from sorption enhanced chemical looping gasification of a lignocellulosic biomass. Renew. Energy 2023, 217, 112279. [Google Scholar] [CrossRef]
  52. Cormos, C.C. Green hydrogen production from decarbonized biomass gasification: An integrated techno-economic and environmental analysis. Energy 2023, 270, 126926. [Google Scholar] [CrossRef]
  53. Dumbrava, I.D.; Cormos, C.C. Techno-economical evaluations of decarbonized hydrogen production based on direct biogas conversion using thermo-chemical looping cycles. Int. J. Hydrogen Energy 2021, 46, 23149–23163. [Google Scholar] [CrossRef]
  54. Zhang, H.; Wang, L.; Pérez-Fortes, M.; Van Herle, J.; Maréchal, F.; Desideri, U. Techno-economic optimization of biomass-to-methanol with solid-oxide electrolyzer. Appl. Energy 2020, 258, 114071. [Google Scholar] [CrossRef]
  55. Alikulov, K.; Aminov, Z.; Anh, L.H.; Xuan, T.D.; Kim, W. Comparative Technical and Economic Analyses of Hydrogen-Based Steel and Power Sectors. Energies 2024, 17, 1242. [Google Scholar] [CrossRef]
  56. Mohan, G.; Venkataraman, M.B.; Coventry, J. Sensible energy storage options for concentrating solar power plants operating above 600 °C. Renew. Sustain. Energy Rev. 2019, 107, 319–337. [Google Scholar] [CrossRef]
  57. Zhang, S.; Jian, W.; Zhou, J.; Li, J.; Yan, G. A new solar, natural gas, and biomass-driven polygeneration cycle to produce electrical power and hydrogen fuel; thermoeconomic and prediction approaches. Fuel 2023, 334, 126825. [Google Scholar] [CrossRef]
  58. Hai, T.; Radman, S.; Abed, A.M.; Shawabkeh, A.; Abbas, S.Z.; Deifalla, A.; Ghaebi, H. Exergo-economic and exergo-environmental evaluations and multi-objective optimization of a novel multi-generation plant powered by geothermal energy. Process Saf. Environ. Prot. 2023, 172, 57–68. [Google Scholar] [CrossRef]
  59. Yang, Y.; Liu, J.; Shen, W.; Li, J.; Chien, I.L. High-efficiency utilization of CO2 in the methanol production by a novel parallel-series system combining steam and dry methane reforming. Energy 2018, 158, 820–829. [Google Scholar] [CrossRef]
  60. Bouzgarrou, S.M.; Abed, A.M.; Chauhan, B.S.; Alsenani, T.R.; Alharbi, F.S.; Alkhalaf, S.; Albaijan, I.; Abdullaeva, B.S.; Deifalla, A. Thermo-economic-environmental evaluation of an innovative solar-powered system integrated with LNG regasification process for large-scale hydrogen production and liquefaction. Case Stud. Therm. Eng. 2024, 53, 103904. [Google Scholar] [CrossRef]
  61. Khosravi, S.; Saray, R.K.; Neshat, E.; Arabkoohsar, A. Towards an environmentally friendly power and hydrogen co-generation system: Integration of solar-based sorption enhanced gasification with in-situ CO2 capture and liquefaction process. Chemosphere 2023, 343, 140226. [Google Scholar] [CrossRef]
  62. Hamedani, S.R.; Villarini, M.; Colantoni, A.; Moretti, M.; Bocci, E. Life cycle performance of hydrogen production via agro-industrial residue gasification-a small scale power plant study. Energies 2018, 11, 675. [Google Scholar] [CrossRef]
  63. GreenDelta GmbH Consulting and Software Company. OpenLCA Source Software. 2024. Available online: https://www.openlca.org/onlinelca/ (accessed on 6 August 2024).
  64. Jovijari, F.; Kosarineia, A.; Mehrpooya, M.; Nabhani, N. Exergy, Exergoeconomic, and Exergoenvironmental Analysis in Natural Gas Liquid Recovery Process. Iran. J. Chem. Chem. Eng. 2023, 42, 237–268. [Google Scholar]
  65. Huang, W.; Liu, Y.; Zheng, D.; Chen, X.; Yang, M.; Li, Y.; Li, X. Exergy-environment assessment for energy system: Distinguish the internal and total exergy loss, and modify the contribution of utility. Energy Convers. Manag. 2022, 251, 114975. [Google Scholar] [CrossRef]
  66. Ghasemi, A.; Rad, H.N.; Akrami, M.; Marefati, M. Exergoeconomic and exergoenvironmental analyzes of a new biomass/solar-driven multigeneration energy system: An effort to maximum utilization of the waste heat of gasification process. Therm. Sci. Eng. Prog. Progress. 2024, 48, 102407. [Google Scholar] [CrossRef]
  67. Mofrad, K.G.; Zandi, S.; Salehi, G. Exergoeconomic and exergoenvironmental assessment of a geothermal-driven cogeneration system utilizing dual-pressure organic Rankine cycle and zeotropic mixtures. Energy Rep. 2023, 9, 5206–5223. [Google Scholar] [CrossRef]
  68. Arfan, M.; Eriksson, O.; Wang, Z.; Soam, S. Life cycle assessment and life cycle costing of hydrogen production from biowaste and biomass in Sweden. Energy Convers. Manag. 2023, 291, 117262. [Google Scholar] [CrossRef]
  69. Wang, H.; Song, M.; Taghavi, M. Comprehensive analysis and optimization of a low-carbon multi-generation system driven by municipal solid waste and solar thermal energy integrated with a microbial fuel cell. Int. J. Low-Carbon. Technol. 2024, 19, 455–467. [Google Scholar] [CrossRef]
  70. Cao, Y.; Habibi, H.; Zoghi, M.; Raise, A. Waste heat recovery of a combined regenerative gas turbine-recompression supercritical CO2 Brayton cycle driven by a hybrid solar-biomass heat source for multi-generation purpose: 4E analysis and parametric study. Energy 2021, 236, 121432. [Google Scholar] [CrossRef]
  71. Nakyai, T.; Authayanun, S.; Patcharavorachot, Y.; Arpornwichanop, A.; Assabumrungrat, S.; Saebea, D. Exergoeconomics of hydrogen production from biomass air-steam gasification with methane co-feeding. Energy Convers. Manag. 2017, 140, 228–239. [Google Scholar] [CrossRef]
  72. Baykara, S.Z. Hydrogen: A brief overview on its sources, production and environmental impact. Int. J. Hydrogen Energy 2018, 43, 10605–10614. [Google Scholar] [CrossRef]
  73. Hren, R.; Vujanović, A.; Van Fan, Y.; Klemeš, J.J.; Krajnc, D.; Čuček, L. Hydrogen production, storage and transport for renewable energy and chemicals: An environmental footprint assessment. Renew. Sustain. Energy Rev. 2023, 173, 113113. [Google Scholar] [CrossRef]
  74. Chisalita, D.A.; Petrescu, L.; Galusnyak, S.C.; Cormos, C.C. Environmental evaluation of hydrogen production employing innovative chemical looping technologies—A Romanian case study. Int. J. Hydrogen Energy 2023, 48, 12112–12128. [Google Scholar] [CrossRef]
  75. Aravindan, M.; Kumar, G.P. Optimizing novel green hydrogen production from solar and municipal solid waste: A thermo-economic investigation with environmental comparison between integrated low temperature power cycles. Process Saf. Environ. Prot. 2024, 186, 421–447. [Google Scholar] [CrossRef]
  76. Zhang, Y.; Ma, S.; Yue, W.; Tian, Z.; Yang, C.; Gao, W. Energy, exergy, economic and environmental (4E) evaluation of a solar-integrated energy system at medium–high temperature using CO2 as the parabolic trough collector (PTC) working medium. Energy Convers Manag. 2023, 296, 117683. [Google Scholar] [CrossRef]
  77. Olatayo, K.I.; Wichers, J.H.; Stoker, P.W. Energy and economic performance of small wind energy systems under different climatic conditions of South Africa. Renew. Sustain. Energy Rev. 2018, 98, 376–392. [Google Scholar] [CrossRef]
  78. Qi, X.; Yao, X.; Guo, P.; Han, Y.; Liu, L. Applying life cycle assessment to investigate the environmental impacts of a PV–CSP hybrid system. Renew. Energy 2024, 227, 120575. [Google Scholar] [CrossRef]
  79. Gasa, G.; Prieto, C.; Lopez-Roman, A.; Cabeza, L.F. Life cycle assessment (LCA) of a concentrating solar power (CSP) plant in tower configuration with different storage capacity in molten salts. J. Energy Storage 2022, 53, 105219. [Google Scholar] [CrossRef]
  80. Cavalcanti, E.J.C. Exergoeconomic and exergoenvironmental analyses of an integrated solar combined cycle system. Renew. Sustain. Energy Rev. 2017, 67, 507–519. [Google Scholar] [CrossRef]
  81. Nabat, M.H.; Soltani, M.; Razmi, A.R.; Nathwani, J.; Dusseault, M.B. Investigation of a green energy storage system based on liquid air energy storage (LAES) and high-temperature concentrated solar power (CSP): Energy, exergy, economic, and environmental (4E) assessments, along with a case study for San Diego, US. Sustain. Cities Soc. 2021, 75, 103305. [Google Scholar] [CrossRef]
  82. Cen, S.; Li, K.; Liu, Q.; Jiang, Y. Solar energy-based hydrogen production and post-firing in a biomass fueled gas turbine for power generation enhancement and carbon dioxide emission reduction. Energy Convers. Manag. 2021, 233, 113941. [Google Scholar] [CrossRef]
  83. Li, M.; Zhuang, Y.; Song, M.; Li, W.; Du, J. Techno-economic and carbon footprint feasibility assessment for polygeneration process of carbon-capture coal-to-methanol/power and molten carbonate fuel cell. Energy Convers. Manag. 2021, 235, 114015. [Google Scholar] [CrossRef]
  84. Moosazadeh, M.; Tariq, S.; Safder, U.; Yoo, C.K. Techno-economic feasibility and environmental impact evaluation of a hybrid solar thermal membrane-based power desalination system. Energy 2023, 278, 127923. [Google Scholar] [CrossRef]
  85. Tiktaş, A.; Gunerhan, H.; Hepbasli, A.; Açıkkalp, E. Exergy-based techno-economic and environmental assessments of a proposed integrated solar powered electricity generation system along with novel prioritization method and performance indices. Process Saf. Environ. Prot. 2023, 178, 396–413. [Google Scholar] [CrossRef]
  86. Dong, R.E.; AlZubaidi, A.A.H.; Nutakki, T.U.K.; Alsenani, T.R.; Bouzgarrou, S.M.; Albani, A.; Alharbi, F.S.; Abdullaev, S.; Deifalla, A. A heat recovery-based ecofriendly solar thermal-driven electricity/hydrogen/freshwater multigeneration scheme using LNG regasification: Energy, exergy, economic, and environmental (4E) analysis. Case Stud. Therm. Eng. 2024, 53, 103853. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of the hybrid biomass–solar driven CHP with steam- and oxygen-based biomass gasification.
Figure 1. Schematic diagram of the hybrid biomass–solar driven CHP with steam- and oxygen-based biomass gasification.
Energies 17 04264 g001
Figure 2. Schematic diagram of the hybrid biomass–solar driven CHP with sorption-enhanced reforming.
Figure 2. Schematic diagram of the hybrid biomass–solar driven CHP with sorption-enhanced reforming.
Energies 17 04264 g002
Figure 3. Schematic of the hybrid biomass–solar driven CHP with iron chemical looping reforming.
Figure 3. Schematic of the hybrid biomass–solar driven CHP with iron chemical looping reforming.
Energies 17 04264 g003
Figure 4. Schematic diagram of the hybrid biomass–solar driven CHP with supercritical water gasification.
Figure 4. Schematic diagram of the hybrid biomass–solar driven CHP with supercritical water gasification.
Energies 17 04264 g004
Figure 5. Hourly variation in the DNI for Tashkent and the efficiency of the CSP plant.
Figure 5. Hourly variation in the DNI for Tashkent and the efficiency of the CSP plant.
Energies 17 04264 g005
Figure 6. Average monthly exergy efficiency of the CSP plants.
Figure 6. Average monthly exergy efficiency of the CSP plants.
Energies 17 04264 g006
Figure 7. Total purchased equipment cost of the system.
Figure 7. Total purchased equipment cost of the system.
Energies 17 04264 g007
Figure 8. Specific CAPEX (a) and OPEX (b) costs for the examined renewable hydrogen and power production.
Figure 8. Specific CAPEX (a) and OPEX (b) costs for the examined renewable hydrogen and power production.
Energies 17 04264 g008
Figure 9. Exergy efficiency, unit cost of hydrogen, and cost rate of the system.
Figure 9. Exergy efficiency, unit cost of hydrogen, and cost rate of the system.
Energies 17 04264 g009
Figure 10. Comparison of CO2 emissions (a) and the ecological footprint (b) of different hydrogen and power production systems.
Figure 10. Comparison of CO2 emissions (a) and the ecological footprint (b) of different hydrogen and power production systems.
Energies 17 04264 g010
Figure 11. Exergoenvironmental parameters of the proposed models: impact factor, damage effectiveness index, impact improvement, and sustainability index (a); ecological effect factor and social ecological factor (b).
Figure 11. Exergoenvironmental parameters of the proposed models: impact factor, damage effectiveness index, impact improvement, and sustainability index (a); ecological effect factor and social ecological factor (b).
Energies 17 04264 g011
Figure 12. The effects of annual variation in the CSP system on the following environmental indicators: ecological effect factor (a); social ecological factor (b).
Figure 12. The effects of annual variation in the CSP system on the following environmental indicators: ecological effect factor (a); social ecological factor (b).
Energies 17 04264 g012
Figure 13. The effects of variations in operation time on the following costs: exergetic performance and exergy destruction (a); specific CAPEX and investment cost rate (b).
Figure 13. The effects of variations in operation time on the following costs: exergetic performance and exergy destruction (a); specific CAPEX and investment cost rate (b).
Energies 17 04264 g013
Figure 14. The effects of operation time on the LCOE, LCOH, and LCOP of the proposed systems.
Figure 14. The effects of operation time on the LCOE, LCOH, and LCOP of the proposed systems.
Energies 17 04264 g014
Figure 15. The effects of operation time on the specific exergy hydrogen cost, relative difference, and exergoeconomic factor.
Figure 15. The effects of operation time on the specific exergy hydrogen cost, relative difference, and exergoeconomic factor.
Energies 17 04264 g015
Figure 16. The effects of operation time on the environmental indexes.
Figure 16. The effects of operation time on the environmental indexes.
Energies 17 04264 g016
Figure 17. The effects of different modes on the environmental impacts of the system.
Figure 17. The effects of different modes on the environmental impacts of the system.
Energies 17 04264 g017
Table 1. The main input parameters used for the modeling [15,22,47,48].
Table 1. The main input parameters used for the modeling [15,22,47,48].
ParameterValue
C, H, N, S, O (%)44.54, 6.74, 0.8, 0.23, 40.68
LHV of cotton stalk (MJ/kg)16.82
Moisture (%)10.32
LocationTashkent, Uzbekistan
Direct normal irradiation (W/m2)845
Optical efficiency of the heliostats (%)71
Optical efficiency of the receiver (%)92
Operating temperature of the working fluid (°C)600–870
Operating hours (hours)10, 12, 14, 16
Gas turbine inlet pressure (bar)16
Ambient temperature (°C)25
Turbine and compressor isentropic efficiency (%)85
Steam turbine inlet pressure (bar)30
Steam turbine inlet temperature (°C)430
Interest fate (%)10
Economic life (years)25
Table 2. Primary premise of the CAPEX and OPEX estimates and economic assumptions.
Table 2. Primary premise of the CAPEX and OPEX estimates and economic assumptions.
CAPEXMethodRef.
Purchased Equipment Cost (PEC)Purchased equipment cost of all equipment as per Table A1 of Appendix B
Direct Cost (DC) P E C s y s t o t P E C C S P   p l a n t × ϕ + P E C C S P   p l a n t
Installation   factor ,   ϕ :
Purchased equipment Installation 30 %   of   P E C [51]
Piping 5 %   of   P E C [9]
Electrical system 10 %   of   P E C [40]
Yard improvement 10 %   of   P E C [40]
Building and services 4 %   of   P E C [9]
Indirect cost (IDC)Engineering and construction 10 %   of   P E C [51]
Total Direct and Indirect Cost (TDIC) D C + I C
Contingency (CCont) 15 %   of   T D I C [51]
CAPEX T D I C + C C o n t + C L a n d
OPEXValueCEPCIRef.
Variable cost
Cooling water price0.01 USD/t800.8[48]
Boiler water treatment cost95,000 USD/month800.8[48]
Cost of biomass disposal16 USD/t816[52]
Biomass cost2 USD/GJ800.8[40]
Fe-based OC solvent cost91.2 USD/t596.2 1[53]
Limestone CaO solvent cost22.8 USD/t596.2 1[53]
MEA solvent1250 USD/t607.5[54]
MEA make-up1.5 kgMEA/tCO2607.5[54]
Distilled water0.08 USD/kg800.8[55]
Electricity cost0.08 USD/kWh800.8[48]
Hot water cost2.02 USD/GJ800.8[48]
Reformer catalyst57,000 USD/m3596.2 1[53]
WGS catalyst15,960 USD/ta596.2 1[53]
ZnO catalyst1570 USD/t596.2[51]
NaCl49 USD/t541.7[56]
KCl455 USD/t541.7[56]
ZnCl21870 USD/t541.7[56]
Fixed cost
Capital annuity cost (CAnn) C R F × C A P E X [40]
O&M 9.33   U S D m 2 y 1 × A h e l i o + 3 % o f P E C I n s [40]
Property, taxes, and insurance costs (CPTI) 2 %   of   C A P E X [9]
Interest rate (ir)3% [9]
Annual operation hours, τCalculated based on daily operation time (10 h, 12 h, 14 and 16 h)
Plant life (n)25 years [52]
Discount rate8% [40]
1 Average exchange rate: EUR 1 = USD 1.14 in 2020.
Table 3. Main thermodynamic performance indexes.
Table 3. Main thermodynamic performance indexes.
ParameterUnitModel 1Model 2Model 3Model 4
Biomass inputt/h104.07104.7593.2487.16
Biomass exergy rateMW559.39563.02501.15468.52
Average hourly exergy rate of solar energyMW171.35183.45188.88194.57
Exergy rate of airMW123.98136.5135.94106.24
Exergy rate of water feedMW0.741.291.070.84
Net power outputMWe46.8355.3766.3651.48
Exergy rate of heatingMWth23.323.323.2623.3
Hydrogen exergy rateMW281.07255.13188.88283.46
Exergy efficiency of H2 production% 32.8628.8522.8436.8
Energy penalty % point−2.34+1.67+7.65−5.98
Exergy rate of CO2MW17.7815.1315.216.6
Exergy efficiency of the overall system%43.1339.4635.6148.67
Total exergy destruction rateMW463.84507.77415.44336.89
Total exergy lossMW22.6327.55116.9558.43
Exergy destruction yield%54.2257.4250.2543.74
Table 4. Estimated costs for the solar-based biomass gasification power plants (in USD millions).
Table 4. Estimated costs for the solar-based biomass gasification power plants (in USD millions).
Model 1Model 2Model 3Model 4
Total purchased equipment cost (TPEC)253.91159.59186.32243.07
Direct cost (DC)350.11197.25243.77330.36
Indirect cost (IDC)25.3915.9618.6324.31
Total direct and indirect cost375.51213.21262.4354.67
Contingency cost (Ccont)56.3331.9839.3653.2
The cost of land (Cland)1.261.351.391.43
Total CAPEX433.09246.54303.16409.3
Capital annuity cost (CAnn)24.8714.1617.4123.51
Property, taxes, and insurance cost (CPTI)8.664.936.068.19
Material cost (Material)16.415.2514.7213.93
Variable cost (O&M)9.016.327.489.12
Total OPEX52.8435.7539.2451.68
Total investment cost of the system485.93282.29342.4460.98
Table 5. Levelized cost of the products for renewable hydrogen and power production processes.
Table 5. Levelized cost of the products for renewable hydrogen and power production processes.
Model 1Model 2Model 3Model 4
LCOE, USD/MWh117.658.959.0101
LCOH, USD/MWh19.612.820.718.3
LCOP, USD/MWh14.99.413.313.9
LCC, USD/kg2.381.472.392.24
Table 6. Exergoeconomic data of the solar- and biomass-based CHP plant.
Table 6. Exergoeconomic data of the solar- and biomass-based CHP plant.
E ˙ D , MW C ˙ t o t , USD/h c F , USD/GJ c P , USD/GJ C ˙ D , USD/h C ˙ L , USD/h Z T , USD/h r k f k , %
Model 1463.8514,124.231.318.802191.07106.917649.655.70776.9
Model 2507.7811,626.091.286.782341.16127.014445.54.29764.3
Model 3415.4411,053.281.218.511822.73513.135390.815.98269.77
Model 4336.8913,247.481.228.551482.19257.068152.995.99882.42
Table 7. Environmental impact points of the components.
Table 7. Environmental impact points of the components.
UnitModel 1Model 2Model 3Model 4UnitModel 1Model 2Model 3Model 4
Receiver3090.53246.83374.73461.6GT67,257.277,417.673,574.368,150.4
Heliostat 173,509371,385384,354394,572HRSG185,538185,538185,538185,538
S-HE 1,548,039195,541.2182,109.5215,786.5ST35,085.135,085.135,085.135,085.1
TES1,005,3751,002,745176,53701,684,474Gen95.295.298.698.3
S-pump767.3766.2963.81346.6Cooler79.841.9387.4212.8
Dryer23,458.123,610.421,015.80.0ABS31,929.30.028,584.329,816.1
Gasifier41,924.439,049.90.034,857.2Cond106.5106.5106.5106.5
ASU2338.10.00.00.0Boiler96.996.996.996.9
W-HE89,257.0116,523.0216,709.8229,842.7FR0.00.036,848.60.0
WGS5521.60.0855.63892.0SR0.00.02804.70.0
HEX51,103.80.012,077.1185,605.5AR0.00.019,278.20.0
Wpump147.8136.4208.0233.5Cycl0.0206.1827.30.0
H2S 33,949.119,566.540,687.031,882.2Exp0.00.035,380.632,834.5
PSA33,898.119,515.10.031,834.6Scomp0.00.02330.43638.1
CC75,491.699,416.948,668.778,319.6Calciner0.028,851.20.00.0
AComp4978.35093.64867.94241.9Refor0.00.00.018,401.7
Total EI, Pts/h 37.4224.3934.0235.86
Table 8. Exergoenvironmental parameters of the proposed power plants.
Table 8. Exergoenvironmental parameters of the proposed power plants.
System b F B ˙ F B ˙ P B ˙ D Y U B ˙ t o t r b f b Y M Y ´
mPts/GJPts/hPts/hPts/hPts/hPts/h-%Pts/hPts/MWh
Model 1500.001006.902863.89834.921856.992691.913.5368.9846.610.995
Model 2500.001013.432928.06914.001914.622828.623.8767.6943.820.792
Model 3500.00902.072849.62747.791947.552695.344.6872.2654.120.816
Model 4500.00843.332979.28606.402135.952742.353.6277.8956.471.097
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

Aminov, Z.; Alikulov, K.; Xuan, T.-D. Economic and Environmental Analyses of an Integrated Power and Hydrogen Production Systems Based on Solar Thermal Energy. Energies 2024, 17, 4264. https://doi.org/10.3390/en17174264

AMA Style

Aminov Z, Alikulov K, Xuan T-D. Economic and Environmental Analyses of an Integrated Power and Hydrogen Production Systems Based on Solar Thermal Energy. Energies. 2024; 17(17):4264. https://doi.org/10.3390/en17174264

Chicago/Turabian Style

Aminov, Zarif, Khusniddin Alikulov, and Tran-Dang Xuan. 2024. "Economic and Environmental Analyses of an Integrated Power and Hydrogen Production Systems Based on Solar Thermal Energy" Energies 17, no. 17: 4264. https://doi.org/10.3390/en17174264

APA Style

Aminov, Z., Alikulov, K., & Xuan, T.-D. (2024). Economic and Environmental Analyses of an Integrated Power and Hydrogen Production Systems Based on Solar Thermal Energy. Energies, 17(17), 4264. https://doi.org/10.3390/en17174264

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