Applied Thermal Engineering: Santanu Pramanik, R.V. Ravikrishna

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 36

Applied Thermal Engineering 127 (2017) 602–637

Contents lists available at ScienceDirect

Applied Thermal Engineering


journal homepage: www.elsevier.com/locate/apthermeng

A review of concentrated solar power hybrid technologies


Santanu Pramanik, R.V. Ravikrishna ⇑
Department of Mechanical Engineering, Indian Institute of Science, Bangalore, India

h i g h l i g h t s

 This paper reviews hybrids of CSP with coal, natural gas, biomass, wind, geothermal.
 Technologies have been categorized into high, medium, and low-renewable hybrids.
 CO2 emissions highest for low-renewable hybrids, followed by medium and high.
 Low-renewable hybrids offer higher efficiency and capacity factor at lower cost.

a r t i c l e i n f o a b s t r a c t

Article history: This paper reviews the hybrid power generation technologies of concentrated solar power (CSP) and
Received 5 April 2017 other renewable and non-renewable resources such as biomass, wind, geothermal, coal, and natural
Revised 6 August 2017 gas. The technologies have been categorized into high, medium, and low-renewable hybrids based on
Accepted 7 August 2017
their renewable energy component. The high-renewable hybrids report the least specific CO2 emissions
Available online 12 August 2017
(<100 kg/MW h), followed by the medium (<200 kg/MW h) and low-renewable hybrids (>200 kg/MW h).
The hybrids have been compared based on their plant characteristics and performance metrics using data
Keywords:
from the literature and of actual hybrid power plants. The low-renewable hybrids such as ISCC, solar-
Concentrated solar power
Hybrid
Brayton, and solar-aided coal Rankine power systems are technologically mature and offer superior per-
Biomass formance over the high and medium-renewable hybrids. The medium renewable hybrids such as solar
Wind plants with natural gas backup offer high solar share but suffer mostly from low efficiency and high cost
Geothermal that hinders their market penetration. The high-renewable hybrids such as CSP-wind, CSP-biomass, and
Coal CSP-geothermal have minimum negative impact on the environment. However, several parameters such
Natural gas as energy efficiency, solar-to-electricity efficiency, capacity factor, and cost effectiveness need to improve
for these systems to be competitive.
Ó 2017 Elsevier Ltd. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 603
2. Hybrid technology categories . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 604
3. Review of hybrid technologies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 605
3.1. High renewable hybrids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 605
3.1.1. CSP-biomass hybrid . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 605
3.1.2. CSP-geothermal hybrids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 608
3.1.3. CSP-wind hybrids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 611
3.2. Medium-renewable hybrids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 613
3.2.1. Size of the solar field . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 613
3.2.2. Impact on the environment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 613
3.2.3. Performance comparison . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 614
3.2.4. Other studies. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 614
3.2.5. Operational plants and challenges. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 614
3.2.6. Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 614

⇑ Corresponding author.
E-mail address: ravikris@mecheng.iisc.ernet.in (R.V. Ravikrishna).

http://dx.doi.org/10.1016/j.applthermaleng.2017.08.038
1359-4311/Ó 2017 Elsevier Ltd. All rights reserved.
S. Pramanik, R.V. Ravikrishna / Applied Thermal Engineering 127 (2017) 602–637 603

3.3. Low renewable hybrids. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 615


3.3.1. Solar-Brayton cycles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 615
3.3.2. Solar-aided coal power plants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 616
3.3.3. Integrated solar combined cycles (ISCC) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 619
4. Comparison of the hybrid technologies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 624
4.1. CO2 emissions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 624
4.2. Solar share . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 625
4.3. Capacity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 625
4.4. Solar-to-electricity efficiency (S-E g). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 626
4.5. Overall energy and exergy efficiency . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 626
4.6. Capacity factor. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 626
4.7. Investment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 627
4.8. Cost of electricity (COE) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 628
5. Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 628
Appendix A. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 633
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 633

1. Introduction level to improve efficiency, cost-effectiveness and ensure energy


autonomy. In this context, research and development of hybrid
Solar radiation is the most promising renewable energy source solar energy technologies have assumed more significance than
that can displace fossil fuels and meet our future energy demands. ever. This paper is an attempt to review the major hybrid technolo-
The massive use of fossil fuels in the past century has led to an gies involving concentrated solar power (CSP), their current state-
unprecedented rise in greenhouse gases, resulting in global warm- of-the-art, their comparative strengths and weaknesses, and rec-
ing and climate change [1]. Apart from climate change, burning of ommendations for future research.
fossil fuels has added costs associated with human and ecosystem Solar energy can be harnessed in most parts of the world due to
toxicity, acidification of land, particulate formation and freshwater its distributed nature (Fig. 1). The working of a CSP plant is similar
eutrophication [2]. The Paris Climate Agreement [3] that has to conventional power plants in terms of usage of a power cycle,
recently come into force is a major step in the direction of mitigat- working fluid, turbines, etc., where the fossil fuel is replaced by
ing these effects. Hence, countries are increasingly looking towards solar energy. A CSP system can provide power similar to a conven-
renewable sources to meet their energy requirements and to tional plant [6], although the intermittency of solar radiation
reduce their carbon footprint. Powering the world by renewable (Figs. 2 and 3) presents several challenges. This intermittency
sources such as solar, wind, and water by the year 2050 is feasible, results in low plant capacity factor (23–50%) [7] and efficiency
but requires large-scale expansion of the transmission infrastruc- (15–30%) [8] for standalone CSP plants, thereby increasing the
ture, sensible policy deployment, and major changes in the social investment and cost of electricity. With technological advance-
and political landscape [4,5]. The interconnection of these ments, such as the addition of thermal energy storage systems
resources at the grid level will reduce fluctuations in the supply. and hybridization, it is estimated that CSP can become cost-
An alternate scenario, in which the renewable energy technologies competitive with natural gas around the year 2020 and with
are hybridized at the power plant level, can overcome several out- coal-based power by around 2025 [9]. The Gemasolar solar plant
standing challenges while providing a boost to distributed power with 15 h of energy storage and 15% hybridization with natural
generation, energy autonomy and faster realization of energy sus- gas can operate at 74% capacity factor [10], close to that of conven-
tainability. The optimal path to realizing this transformation to tional power plants. Hybridization will also enable CSP plants to be
renewable energy resources may finally be a combination of both established in regions with moderate direct normal irradiation
the strategies: integrating geographically dispersed variable (DNI  1700 kW h/m2/year), thereby moving it away from dry arid
energy resources through expansion of the transmission grid while regions and closer to load centers [6]. In a future where CSP plants
hybridizing co-located renewable resources at the power plant become a key supplier of dispatchable energy, hybridization seems

Fig. 1. Map of the yearly averaged downward surface solar radiation reaching the surface (W/m2) [4].
604 S. Pramanik, R.V. Ravikrishna / Applied Thermal Engineering 127 (2017) 602–637

Fig. 2. Hourly variation of DNI (W/m2) on four different dates at representative locations of four different continents, namely, Daggett in the USA, Sevilla in Spain, Ahmedabad
in India, and Antofagasta in Chile. The profiles have been obtained from the Solar Advisor Model (SAM), NREL [11].

hybridization with Rankine cycle power plants using fuels such


as coal, natural gas, biomass, and waste materials. The authors
report that Fresnel reflectors perform best in the low-
temperature category (<400 °C) due to low land usage and cleaning
water consumption. This is followed by parabolic trough technol-
ogy. In the high-temperature category (>450 °C), solar towers offer
the best performance. In another study, Peterseim et al. [14] have
analyzed the hybridization of CSP with coal, natural gas, biomass,
geothermal and wind for Australian conditions. They have catego-
rized the systems into light, medium and strong hybrids depending
on the level of integration and synergy between the technologies.
The authors report that the cost of a hybrid plant can be 50% lower
Fig. 3. Variation of monthly averaged DNI (W/m2) at representative locations of in comparison to standalone CSP plants. However, there are no
four different continents, namely, Daggett in the USA, Sevilla in Spain, Ahmedabad
studies currently that review and compare all CSP hybrid technolo-
in India, and Antofagasta in Chile. The profiles have been obtained from the Solar
Advisor Model (SAM), NREL [11]. gies present in the literature and the present work attempts to
address this gap.
The present study performs an in-depth assessment of CSP
inevitable. Hence, the purpose of the present work is to review the hybridization with other energy sources such as coal, natural gas,
most important CSP hybrid technologies currently available. biomass, geothermal, and wind. Among the CSP technologies, para-
Hybrid plant configurations involving CSP that simultaneously bolic trough, linear Fresnel, and solar tower technologies have been
use solar energy and fossil fuel have been reviewed by Sheu considered. In the following sections, the individual hybrid tech-
et al. [8]. The authors have classified the solar-fossil fuel hybrids nologies will be reviewed and their performance characteristics
into three broad categories such as the solarized gas turbines, com- will be explored. This will be followed by a comparison of the
bined cycles, and solar reforming. The authors have also proposed a hybrid systems based on system characteristics and performance
linear combination metric for comparison of the hybrid plants with metrics such as specific CO2 emissions, capacity, capacity factor,
the currently operational power plants. Behar et al. [12] have specific investment, cost of electricity (COE), solar share, energy
reviewed integrated solar combined cycle (ISCC) systems that use efficiency, exergy efficiency, and solar-to-electricity efficiency (S-
parabolic trough (PT) technology. This work reports the current E g) [8]. Finally, the insights obtained from the comparative anal-
status of such systems in terms of performance, configuration, ysis will be used to make recommendations for future research in
and future directions of research and development. The authors these areas.
have reported that direct steam generation parabolic trough
(DSG-PT) technology offers better performance compared to the
best heat-transfer-fluid parabolic trough (HTF-PT) technology in 2. Hybrid technology categories
terms of efficiency, electricity cost and greenhouse gas (GHG)
emissions. They also state that the ISCC plants would become com- Since the primary objective for using renewable technologies
petitive with conventional combined cycle power plants with a such as CSP is sustainability and pollution reduction, the hybrids
gradual increase in the price of the fossil fuels. Jamel et al. [13] have been classified based on the renewable component of the
have reviewed the integration of solar energy with conventional generated power. While hybrids of CSP with wind, biomass and
(combined cycles, steam Rankine cycles, and gas Brayton cycles) geothermal are expected to have the least environmental impact,
and non-conventional power plants (geothermal). The authors the impact of non-renewable hybrids on the environment will
concluded that the ISCC plants offer technical and economic depend on its solar share. Hence, the hybrids have been catego-
advantages over other hybrid configurations. Peterseim et al. [6] rized into three broad divisions based on their renewable energy
have tried to determine the most promising CSP technology for component:
S. Pramanik, R.V. Ravikrishna / Applied Thermal Engineering 127 (2017) 602–637 605

(a) High – hybrids of CSP with wind, biomass, and geothermal [18]. Also, it can provide greater penetration to both CSP and
energy resources. This category has the highest potential in biomass-generated power into the market compared to standalone
decelerating global warming and climate change. CSP and biomass plants.
(b) Medium – this category includes solar plants that use sup-
plementary firing of fossil fuels such as natural gas to 3.1.1.1. Combinations of CSP and biomass technologies. Various
enhance the plant output and capacity factor. Examples of hybrid combinations of CSP and biomass technologies offer distinct
this category include SEGS II-IX in Mojave Desert, California advantages in terms of efficiency and cost. Several hybrid systems
[15] that uses natural gas backup for evaporation and super- (Table 1) have been studied by Peterseim et al. [23]. The maximum
heating. The use of backup fossil fuel in such plants is lim- energy efficiency of 33.2% is obtained for solar tower and gasifica-
ited to 12–15% in Spain, and to 25% in the US [10]. tion, whereas Fresnel and fluidized bed offer the lowest specific
(c) Low – this category represents conventional fossil fuel investment ($/MWe). The efficiency difference between the 17
plants that incorporate solar energy for auxiliary functions combinations is 13%, while the difference in the specific invest-
such as preheating and evaporation. This category includes ments is 31%.
ISCC plants, solar-aided coal power plants, and solar Brayton
cycles. The solar share in such plants is usually less than 20%. 3.1.1.2. Plant capacity. The selection of capacity for CSP-biomass
hybrids is critical as the scaling of cost with plant size follow
3. Review of hybrid technologies reverse trends for standalone CSP and biomass plants [18]. A
capacity of at least 50 MWe is recommended for CSP plants as they
This section will review the existing literature on the various incur high initial investment. Contrarily, the capacity of biomass
CSP-hybrid technologies suitable for power generation. The appli- plants is restricted below 50 MWe due to the increased logistical
cation of such hybrids for other purposes such as heating and cool- costs of biomass transportation. A small size plant receives a con-
ing and polygeneration will be reviewed only briefly. tinuous supply of biomass, although at the expense of a reduction
in efficiency (Fig. 5). A minimum capacity of 5 MWe is recom-
3.1. High renewable hybrids mended for biomass plants to obtain economy of scale benefits
[14]. Peterseim et al. [19] report that several locations in Australia
As discussed in the previous section, these hybrids are formed can support 5–25 MWe hybrid plants, while some others can sup-
by integrating CSP with other renewable sources of energy. port 50 MWe and higher capacity plants. A similar case study for a
30 MWe hybrid plant combining biomass with molten salt solar
3.1.1. CSP-biomass hybrid tower system and 3 h of storage has been reported for Griffith, Aus-
Biomass provides a promising renewable extension to CSP tralia [20]. The first operating commercial CSP-biomass hybrid
plants for continuous operation. Several regions in the world have plant, the Termosolar Borges, has a capacity of 22.5 MWe. How-
biomass and solar resources suitable for hybridization (Fig. 4). The ever, lower capacity plants in the range of 2–10 MWe are possible
potential number of locations for CSP-biomass hybridization is in the case of trigeneration systems. Nixon et al. [16] recommend
improved if multiple feedstocks are considered, although the initial trigeneration as an efficient choice for CSP-biomass hybrids under
investment is higher. Although CSP-biomass hybridization is in the Indian conditions.
nascent stages, a number of feasibility and technology assessment
studies have been conducted for countries such as India [16], Brazil 3.1.1.3. Installation cost. The installation cost for CSP-biomass
[17,18], Indonesia [18], Australia [14,19,20] and Europe [21]. The hybrids is lower than standalone CSP with similar power output.
global warming potential (GWP) of such a hybrid can be up to 10 Also, the specific investment decreases with an increase in the
times lower than that of CSP plants with non-biomass extensions plant capacity (Fig. 6). Significant cost reduction of the order of

Fig. 4. Potential regions for CSP-biomass hybrid plants worldwide [22].


606 S. Pramanik, R.V. Ravikrishna / Applied Thermal Engineering 127 (2017) 602–637

Table 1
Comparison of various CSP-biomass hybridization schemes [23]. Exchange rate: 1 AU$ = 0.74 US$.

CSP technology CSP working fluid Biomass technology Peak net efficiency (%) Specific investment (m$/MWe)
Parabolic trough Thermal oil Grate 29.3 4.96
Parabolic trough Thermal oil Fluidised bed 29.5 4.88
Parabolic trough DSG Fluidised bed 30.3 4.74
Fresnel DSG Fluidised bed 30.4 3.55
Parabolic trough DSG Fluidised bed 31.5 4.59
Fresnel DSG Fluidised bed 31.5 3.4
Solar tower DSG Fluidised bed 31.5 3.85
Parabolic trough Molten salt Fluidised bed 32.2 4.59
Fresnel DSG Fluidised bed 32.5 3.33
Solar tower DSG Fluidised bed 32.5 3.77
Solar tower Molten salt Fluidised bed 32.3 3.85
Parabolic trough Molten salt Fluidised bed 32.7 4.59
Solar tower DSG Fluidised bed 33.0 3.7
Solar tower Molten salt Fluidised bed 32.8 3.85
Parabolic trough Molten salt Gasification 32.8 4.59
Solar tower DSG Gasification 33.2 3.7
Solar tower Molten salt Gasification 32.9 3.85

50% is possible with CSP-biomass hybridization [14]. Other authors 3.1.1.4. Cost of electricity. The cost of electricity (COE) of CSP-
have reported a 43% and 69% reduction in investment compared to biomass hybrids in India is currently greater than that of a com-
a standalone CSP system [20,23]. In comparison to a biomass-only mercial CSP by a factor of two and coal-fired power plant by a fac-
plant, hybridization reduces the biomass demand and conse- tor of four [16]. The COE is also higher in comparison to standalone
quently decreases land usage by 14–29% [16]. But, the decreased biomass plants. However, the COE may be comparable if the price
demand for biomass and land could not offset the increased initial of biomass increases by 1.2–3.2 times or the cost of solar technol-
investment due to the high cost of CSP. The investment can be ogy decreases by 47.7–98.5% [16]. A scenario such as this is likely
reduced significantly by local production of CSP components and in the future owing to the rising biomass prices and advancements
through advancements in technology. Currently, the installation in CSP technologies. For Indian conditions, the cost of grid expan-
cost of CSP-biomass hybrids is greater than biomass-only plants sion is higher than the establishment of CSP-biomass hybrids. Local
but less than standalone CSP, which makes the hybrid systems outsourcing of biomass can significantly reduce the logistical costs
commercially viable. Hybridization also provides an affordable and reduce the COE in comparison to a centralized system that
way to dispatch CSP in locations where the electricity price is tra- charges higher for distant locations due to transmission and distri-
ditionally low due to large fossil fuel reserves [20]. bution losses.

3.1.1.5. Novel schemes for hybridization. Various novel schemes


have also been proposed in the literature for CSP-biomass
hybridization. Tanaka et al. [24] have analyzed the performance
of a combined cycle consisting of a bubbling fluidized-bed gasifier
and a solar-tower system. The gasifier supplies to the topping gas
turbine cycle whereas the solar tower and the gas turbine exhaust
supplies to the bottoming Rankine cycle. A peak thermal efficiency
of 36.7% has been obtained for the system. Peterseim et al. [20]
have analyzed the performance of a 30 MWe hybrid plant with a
solar tower and biomass boiler (vibrating grate), each with a capac-
ity of 15 MWe. The maximum cycle efficiency obtained at peak
load is 33.4%, while the efficiency with only the biomass boiler
was 30.2%. A hybrid power plant combining parabolic trough and
Fig. 5. Variation of net cycle thermal efficiency of solar-biomass hybrids with plant biomass boiler supplying to a steam turbine has been studied by
capacity for different biomass feedstocks [19].

Fig. 6. Variation of specific investment of solar-biomass hybrids with plant capacity Fig. 7. Exergy destroyed in individual components as a fraction of the input exergy.
for different biomass feedstocks [19]. The regions selected receive DNI > 21–24 MJ/ The exergy lost in the exhaust is 8.72% of the input exergy. The gasifier is operated
m2/day. Exchange rate: 1AU$ = 0.74 US$. at Tg = 1073 K with biomass containing 20% moisture [26].
S. Pramanik, R.V. Ravikrishna / Applied Thermal Engineering 127 (2017) 602–637 607

Table 2 mass plants. The superheating increases the overall hybrid plant
Details of the Termosolar Borges plant [32,33]. efficiency that can lead to the decrease in CSP cost by up to
Technology Parabolic trough-biomass 23.5%. Liu et al. [28] have analyzed two hybrid combined cycles
Background where the solar energy is utilized either for gasification of biomass
Status Operational or preheating of the compressed air. The solar gasification system
Country Spain offers better performance in terms of energy efficiency
Power 22.5 MWe (gI = 29.4%) and solar-to-electricity efficiency (S-E g = 18.5%) in
Land area 96 hectares
Electricity generation 98000 MW h/year (estimated)
comparison to the air preheating configuration (gI=28% and S-E
Cost (approx.) 153 million euros g = 15.1%). The scope of CSP-biomass hybrids for polygeneration
Start date December 2012 under Indian conditions has been analyzed by Sahoo et al. [29].
Technology Thermodynamic and economic analysis of a solar aided sugarcane
Receiver type Evacuated tube parabolic trough bagasse cogeneration plant has been performed by Burin et al.
Heat transfer fluid (HTF) Dowtherm A [30,31]. The hybridization led to an increase of annual electricity
HTF inlet temperature 293 C
production in the range of 1.3–19.8%.
HTF outlet temperature 393 C
Power cycle Steam Rankine
Turbine efficiency 37% at full load 3.1.1.6. Challenges. Several challenges need to be addressed before
Backup type Biomass (2  22 MWt) CSP-biomass hybrid plants are implemented extensively. Most of
Storage None
the CSP plants around the world are located in areas with
DNI > 20 MJ/m2/day. The number of locations that simultaneously
have large biomass resources suitable for hybridization is less.
Srinivas et al. [25]. The addition of solar energy increases the plant The use of cultivable land for the growth of high energy crops for
fuel efficiency (which is the ratio of the work output to the fuel utilization as biomass has given rise to the food versus fuel debate.
energy input, in this case, biomass) from 15% to 32% while decreas- Recent developments in the use of cellulosic biomass for energy
ing the thermal efficiency from 15% to 11% due to the low effi- production have alleviated some concerns. Increasing climate
ciency of the solar reflectors [25]. A couple of recent studies change concerns have also created uncertainty in the production
[26,27] have proposed a novel solar cycle incorporating a biomass and supply of biomass. For the construction of any biomass/hybrid
gasifier, a solar tower system, and a supercritical CO2 Brayton plant, a steady supply of fuel for an average plant lifespan of
cycle. A maximum cycle thermal efficiency of 40.8% is estimated 25 years needs to be ensured. Finally, with the current technolog-
in this hybridization scheme. An exergetic analysis reveals that ical capability, only 30% of the electricity is produced by solar
maximum exergy destruction occurred in the gasifier, followed energy in a CSP-biomass hybrid for a rated steam flow rate (ton-
by the combustion system (Fig. 7). Peterseim et al. [22] have inves- nes/h) [23]. A higher storage capacity (7 h considered in the anal-
tigated the concept of using biomass for superheating the steam ysis) can increase the solar share above 30%.
generated from a 50 MWe parabolic trough system with 7.5 h of
thermal energy storage. The concept is also feasible in regions with 3.1.1.7. Operational plant. Inspite of these challenges, the first and
small biomass resources and would provide higher conversion effi- the only commercial CSP-biomass plant has been operational in
ciencies compared to the installation of small-scale standalone bio- Spain since 2012 [32–34]. It has a capacity of 22.5 MWe and is sup-

Fig. 8. Termosolar Borges power plant in Spain [32,34–36].


608 S. Pramanik, R.V. Ravikrishna / Applied Thermal Engineering 127 (2017) 602–637

plied by 2  22 MW thermal biomass units operating on waste for- resource, several modes of utilization are possible. Geothermal
est biomass and energy crops. The operational experience obtained wells that produce dry steam can be used directly in a turbine to
from this plant will be valuable in the development of newer CSP- generate power. Flash systems use hot water at high pressure to
biomass hybridization strategies. The details of the plant are generate steam in a flash chamber that can drive the turbine. Flash
shown in Table 2 and Fig. 8. geothermal systems are the most common type of geothermal
power plants. The third type of plants uses the low-temperature
3.1.1.8. Summary. In summary, we observe that several combina- geothermal resource to generate vapors of an organic fluid with a
tions of CSP and biomass technologies can be hybridized that offer low boiling point. Such systems are referred to as binary geother-
distinct cost and efficiency advantages. The capacity of these plants mal plants and run on an organic Rankine cycle (Fig. 9). Data cor-
is restricted between 5 and 50 MWe due to the considerations of responding to operational and planned geothermal plants
cost and the steady supply of biomass. Currently, the cost of these worldwide have been compiled elsewhere [37,38].
hybrids lies in between standalone biomass and CSP plants. The Geothermal plants operate at base load with high capacity fac-
cost of electricity is also higher compared to standalone systems, tors due to the continuous supply of energy. However, the low
which can change with an increase in the cost of biomass or a temperature of the resource (150–200 °C) results in reduced
decrease in the cost of CSP components. Several novel schemes power plant efficiency (12%) [37] that increases the specific
are being proposed in the literature that can overcome some of installation cost. The output from a geothermal plant also
the current challenges and make this hybridization economically decreases with increase in the ambient temperature (Fig. 9). The
feasible. A summary of the key parameters and performance met- gradual depletion of a geothermal well reduces the power output
rics of selected studies on CSP-biomass hybrids is presented in of the plant over time. Hybridization with CSP (380 °C) can over-
Table A1 in the Appendix. come some of the challenges currently faced by standalone
geothermal and CSP plants.
3.1.2. CSP-geothermal hybrids
Geothermal energy is another renewable resource that holds 3.1.2.1. Integration schemes. Several integration schemes have been
promise for hybridization with CSP. Geothermal plants use the reported for the hybridization of CSP with geothermal energy
thermal energy beneath the earth’s surface to generate electricity. (Fig. 10). The two most common integration schemes are, (a) pre-
Depending on the temperature and type of the geothermal heating of the geothermal brine using solar energy, and (b) super-

(a) (b)

Fig. 9. (a) T-s diagram of an organic Rankine cycle with recuperation showing the various state points and operating conditions [39]. (b) Reduction in power output from a
geothermal plant with an increase in ambient temperature [39].

Fig. 10. Integration schemes for CSP-geothermal hybridization [40].


S. Pramanik, R.V. Ravikrishna / Applied Thermal Engineering 127 (2017) 602–637 609

heating of the steam by solar energy before entering the turbine. curve shows the opposite trend. The addition of CSP can address
The preheating configuration is the simplest in terms of retrofitting this issue as its output increases with a rise in the ambient temper-
to an existing geothermal power plant [40]. No significant modifi- ature (Fig. 11a). The monthly power output of a hybrid system is
cations in the operation and control of the plant are necessary as also higher and smoother compared to a standalone geothermal
the CSP raises the brine temperature close to the design point of plant (Fig. 11b). The annual decline in power output due to a
operation. This also addresses the issue of the decline in decrease in the temperature of the geothermal well is suitably
geothermal-well productivity with time. However, the utilization modified by hybridization (Fig. 11c). The authors have also
of solar energy at a lower temperature in this configuration leads reported that the solar-to-electricity efficiency of the plant
to exergy destruction and reduces the solar-to-electricity efficiency increases with a decrease in both the geothermal well temperature
of the plant. The superheating configuration offers higher solar-to- and the ambient temperature for a constant solar heat input
electricity efficiency due to high-temperature solar energy input (Fig. 11d). The increase in the plant thermal efficiency is attributed
into the working fluid before the turbine. This also raises the over- to the operation of the plant closer to the design point due to
all plant energy and exergy efficiency. The retrofit cost in such a hybridization.
configuration is significantly higher due to substantial changes in
the power block operating conditions and control strategies.
3.1.2.3. Solar field size. The selection of the solar field size is critical
in determining the hybrid plant performance. The influence of the
3.1.2.2. Output characteristics. The integration of solar energy mod- solar field size on the performance of the hybrid plant has been
ifies the output characteristics of a geothermal plant. Several such reported by Zhou et al. [42–44] under the climatic conditions of
results have been reported in a study by Turchi et al. [40,41] which Australia and New Zealand. In these studies, the parabolic trough
has analyzed the integration of parabolic trough collectors into a system is used to superheat the working fluid before entering the
binary geothermal plant. The hourly output of a standalone turbine. The results show that the solar fraction in the system
geothermal plant decreases as the ambient temperature increases increases with increase in the collector area, but decreases with
with the progress of the day. However, the daily electricity demand increase in the geothermal temperature (Fig. 12a). This can be

Fig. 11. (a) Hourly comparison of power output from a standalone geothermal plant and a hybrid plant [40]. (b) Effect of geothermal resource temperature decline on the
base and hybrid plant monthly power generation [40]. (c) Effect of geothermal resource temperature decline on the base and hybrid plant annual power generation [40]. (d)
Variation of the solar to electricity efficiency of the hybrid plant as a function of the geothermal well temperature and the ambient temperature [40]. The solar energy input is
25% of the geothermal plant design input.
610 S. Pramanik, R.V. Ravikrishna / Applied Thermal Engineering 127 (2017) 602–637

(a) (b)

Net electrical power output (kW)


Solar energy fraction

Solar aperture area (m2) Solar energy fraction

Fig. 12. (a) Determination of the solar energy fraction as a function of solar aperture area at various geothermal reservoir temperatures (Tair = 31 °C; design-point solar
DNI = 1000 W/m2) [43]. (b) Net electrical power output as a function of the solar share for various operation strategies [43].

attributed to the reduction in usage of geothermal energy as the 27.9%, and 71.1% were obtained at San Diego, Imperial, and Pisa,
temperature of the well decreases, which is compensated by a respectively. The economic analysis revealed that hybridization
higher solar fraction. Again, at a constant geothermal temperature, could lead to a 50% cost reduction compared to standalone solar
a rise in the solar fraction increases the net power output of the plants. Improvement in power output and efficiency due to
plant (Fig. 12b). The type of operation strategy also has a signifi- hybridization has also been reported by Ghasemi et al. [39,51].
cant influence on the power output. Overall, a net exergy efficiency Escobar et al. [45,52] have studied two integration configurations
of 12.4% was reported by the authors. The annual output of the for CSP-geothermal hybridization under Chilean conditions and
hybrid plant was 73.6% higher than that of the standalone geother- reported that the double flash hybrid configuration offered the best
mal plant. performance. Greenhut et al. [53] also reported that the flash
hybrid configuration performed better in comparison to the super-
3.1.2.4. Cost of electricity. The cost of electricity (COE) is highly sen- heat configuration in terms of power output, energy, and exergy
sitive to the capital cost of the solar field (Fig. 13). At low prices of efficiency. Zhou [54] has compared the performance of subcritical
the solar field, a higher solar share reduces the cost of electricity. and supercritical organic Rankine cycle in a hybrid CSP-
The opposite trend is observed when the specific cost of the solar geothermal power plant. In comparison to the subcritical plant,
field is high. A low solar share is recommended in such cases to the supercritical plant produces 4–17% more electricity and
make the COE competitive with the base plant. Zhou et al. [43] reduced the solar capital cost by 6–24% and solar to electricity cost
have reported that the COE for a hybrid plant (171$/MW h) was by 4–19%. The maximum energy efficiency is reported for the sub-
23% less than that of the standalone geothermal plant (225$/ critical condition while the exergy efficiency is higher for the
MW h). Thus, a properly designed hybrid plant can become eco- supercritical condition. However, Boghossian [55] found no syn-
nomically superior to that of two standalone plants. Escobar ergy between a Kalina hybrid cycle operating on geothermal brine
et al. [45] have also reported a reduction in COE (56.89$/MW h) and a parabolic trough CSP system. The author reported a 29%
due to hybridization. However, Manente et al. [46] have reported reduction in power output compared to the standalone plants.
that the COE of a hybrid system (180–190 $/MW h) is much greater Changes such as regeneration, reheat, different working fluids,
than that of a standalone geothermal only plant. Similar observa- and incorporating the solar energy at other locations in the cycle
tions have been made by Ghasemi et al. [47]. This shows that the can improve the system thermodynamic performance.
COE is highly sensitive to the location and the availability of
resources.

3.1.2.5. Novel schemes and other studies. Several other studies have
been conducted on CSP-geothermal hybridization for power gener-
ation. A brief analysis of the feasibility of CSP-geothermal hybrids
in Australia has been reported by Peterseim et al. [14]. The analysis
showed that the power output of the standalone geothermal plant
increased from 6.3 MWe to 8.4 MWe for parabolic trough hybrid
and 9.9 MWe for solar tower hybrid. The plant thermal efficiency
also increased to 12.5% and 14.1% from 10.2%, for a parabolic
trough hybrid and solar tower hybrid, respectively. However, ther-
mal storage systems or supplementary fuel firing are required at
night for superheating. Lentz et al. [48,49] have analyzed three
hybrid configurations using parabolic troughs for the Cierro Prieto
geothermal plant located in Mexico. The authors report that the
hybrid system can enhance the power output by augmenting the
steam flow rate. Astolfi et al. [50] have studied the integration of
parabolic trough collectors with a supercritical binary plant for
four different locations. The maximum annual average solar-to-
electricity efficiency, solar share and capacity factor of 9.4%, Fig. 13. Cost of electricity as a function of the solar field capital cost [40].
S. Pramanik, R.V. Ravikrishna / Applied Thermal Engineering 127 (2017) 602–637 611

3.1.2.6. Operational plant. The Stillwater triple hybrid power plant ing, etc. A tabulated summary of selected studies on CSP-
(Fig. 14) is the first hybrid solar-geothermal plant located in geothermal hybrids is presented in Table A2 in the Appendix.
Nevada, USA [56]. It consists of a 33.1 MWe geothermal binary
plant that was commissioned in 2009. Later in 2012, a 26.4 MWe 3.1.3. CSP-wind hybrids
PV unit was added to enhance the output during the hot summers. Wind energy is currently the lowest cost renewable energy and
The system capacity was further augmented by 2 MWe due to the is widely available throughout the world (Fig. 15). Hybridization of
addition of parabolic trough (PT) collectors that preheat the incom- CSP with wind energy is an area that has not been explored widely.
ing geothermal brine. The PT adds around 17 MW of solar thermal This is predominantly because wind and CSP do not have wide syn-
energy into the system. The plant acts as an excellent case study ergy in terms of sharing of infrastructure (Fig. 16), unlike other
for understanding the hybridization of multiple renewable energy thermal energy sources such as biomass and geothermal. Hence,
sources. The combined plant is expected to produce around CSP-wind integration has been referred to as a light hybrid [14].
200 GW h/year of electricity. The capital investment for the instal- However, solar energy naturally complements wind energy in gen-
lation of the parabolic trough system was estimated at $15 million erating power more uniformly as wind speed is lower during the
in 2013. day and summer compared to nights and winter.

3.1.2.7. Summary. CSP-geothermal hybridization can overcome 3.1.3.1. Feasibility studies. Feasibility studies for CSP-wind hybrids
several challenges faced by standalone geothermal plants such as have been conducted for several locations such as Ontario [58],
cost, low efficiency, and reduction in output over time. CSP can Iberian Peninsula [59], India [60], Italy [61] and Arabian peninsula
be incorporated in the geothermal plant both in the preheating [62]. Chen et al. [63] have proposed a model for selection of a suit-
or the superheating configuration. Such integrations lead to more able CSP-wind power generation project by considering the associ-
favorable plant output characteristics compared to standalone sys- ated benefits, opportunities, costs and risks. A study by Kost et al.
tems. However, the capacity of the solar field significantly influ- [64] has suggested that a CSP-wind portfolio of energy production
ences the plant output and the cost of electricity. Several new is economically viable compared to standalone CSP in North Africa.
configurations with thermal storage or supplementary fuel firing Sioshansi et al. [65] have analyzed several configurations for co-
are being proposed to improve the performance of these systems. locating solar and wind power plants in Texas, USA. Their analysis
CSP-geothermal hybrids have also been studied for applications suggests that deployment of hybrid plants with up to 67% CSP can
such as polygeneration, heating of greenhouses, desalination, cool- yield a positive return on investments. However, the results are

Fig. 14. Aerial view of the Stillwater triple hybrid power plant [57].

Fig. 15. Map of the yearly averaged world wind speed (m/s) at 100 m above sea level [4].
612 S. Pramanik, R.V. Ravikrishna / Applied Thermal Engineering 127 (2017) 602–637

Fig. 16. Schematic of a CSP-wind hybrid plant [67].

sensitive to CSP and transmission costs. Peterseim et al. [14] have highest at night when the utility electricity load is the minimum
reported that CSP-wind hybrids have potential in Australia as sev- and vice versa (Fig. 17a). A standalone 100 MWe CSP plant with
eral wind farms and CSP plants are co-located. A number of loca- 6 h of thermal storage performed best to match the peak utility
tions in South Australia are particularly promising as they have loads in July and August. Overall, a 67 MWe wind farm and a
wind speeds >7 m/s and DNI >19.1 MJ/m2/day. The authors also 33 MWe CSP plant with 6 h of thermal storage were found to be
suggest that the excess electricity produced at night by the wind the best match for the utility electricity load (Fig. 17b). However,
farm can be utilized to charge the thermal energy storage system the COE of the hybrid plant was much higher ($108–129/MW h)
of the CSP plant. However, a 260% difference in the day and night in comparison to that of the wind farm ($64/MW h). Reichling
time electricity prices is necessary to make this integration eco- et al. [69] have analyzed the hybridization of an 800 MWe wind
nomically feasible. Santos-Alamillos et al. [66] have also analyzed farm with a 705 MWe parabolic trough solar field in Southern Min-
the hybridization of CSP with wind farms to produce stable renew- nesota. The hybridization led to a more favorable power output
able power in the region of Andalusia in Spain. Their analysis curve that matched the load demand curve. Also, the specific CO2
shows that optimal location of wind and CSP plants can overcome emissions (10.8 g/kW h) are slightly higher than those associated
the spatiotemporal variability in standalone CSP and wind plants with wind (10.2 g/kW h), but lower than those for electricity gen-
and provide stable base-load power. The addition of thermal stor- erated from a parabolic trough solar field (13.4 g/kW h). However,
age to CSP plants enhanced the performance of the hybrids. the cost of electricity (COE) of the hybrid plant (123 $/MW h) is
higher than that of standalone wind plant (67.8 $/MW h). A reduc-
3.1.3.2. Hybrid system performance. The performance of a CSP-wind tion in capital cost for solar technology by 75% could bring the COE
hybrid plant at Texas Panhandle has been analyzed by Vick et al. of the hybrid plant to within 7% of the COE of the wind plant. Pet-
[68]. The objective was to determine the suitability of CSP-wind rakopoulou et al. [70] have analyzed the hybridization of CSP with
hybrids to match the utility electrical load in comparison to a stan- wind for energy autonomy on the Greek island Skyros. The solar
dalone wind farm. The standalone wind farm generation is the plant has a capacity of 10 MWe, while each wind turbine is rated

Fig. 17. (a) Comparison of electricity load and wind farm capacity factor [68]. (b) Annual utility loading compared to different ratios of the wind farm to CSP rated generation
(2004) [68].
S. Pramanik, R.V. Ravikrishna / Applied Thermal Engineering 127 (2017) 602–637 613

at 3.3 MWe. A total of two wind turbines has been used, thereby
taking the total capacity to 16.6 MWe. The mean annual efficiency
of the plant is 19.2% that lies between the reported efficiencies of
CSP and wind plants. The cost of electricity is 400 €/MW h. The
hybrid plant was reported to have lower land use requirement
and higher exergetic efficiency, thus making it a promising option
for energy autonomy of remote locations such as islands.

3.1.3.3. Summary. Thus, CSP-wind hybrids can generate power that


matches favourably with the load demand curve as compared to
standalone plants. However, the high capital cost of the CSP com-
ponent as compared to the wind farm increases the cost of electric-
ity and makes the hybrid plant economically unattractive. Further
research is necessary on the addition of storage and/or backup
fuel-firing, site selection and cost reduction of CSP to make the Fig. 19. Variation of annual efficiency, fossil backup percentage, the size of thermal
energy storage (TES) and cost of electricity (COE) as a function of the solar multiple
hybrid scheme feasible. A tabulated summary of selected studies [73]. Exchange rate: 1 € = 1.13 $.
on CSP-wind hybrids has been presented in Table A3 in the Appen-
dix. Currently, there are no CSP-wind hybrid plants operational in required to deliver the rated thermal input to the turbine. The solar
the world. multiple for a CSP plant is usually greater than 1. Montes et al. [73]
have performed the analysis for a 50 MWe parabolic trough power
3.2. Medium-renewable hybrids plant, with thermal energy storage and auxiliary natural gas-fired
boiler. The storage system and natural gas boiler are operated to
Standalone solar plants suffer from low capacity factor that maintain a constant power output. It was observed that the annual
necessitates the use of thermal energy storage systems. The size efficiency decreases with increase in the solar multiple as more of
of the storage system required to produce power as a base load the electricity is produced from the solar collectors that have lower
plant is two orders of magnitude higher (1000 h) than the current efficiency. The contribution of the fossil fuel in the output electric-
storage systems (10 h) [71]. For current storage systems, the solar ity also decreases with increase in the solar multiple. The thermal
fraction (fraction of the output power generated by solar energy) energy storage size increases to accommodate the greater amount
ranges from 0.4 to 0.5 that reduces to 0.2 with no storage. Hence, solar energy surplus. The higher investment owing to the larger
several standalone solar plants use fossil fuel (Fig. 18), preferably solar multiple causes an increase in the cost of electricity (COE).
natural gas, as a backup to keep the plant operational during low
insolation periods. There are, however, restrictions in some coun- 3.2.2. Impact on the environment
tries on the maximum usage of fossil fuels as a backup. In Spain, A few studies report the impact of fossil fuel backup on the
fossil fuels can contribute to 12–15% of the total power output in environment using life cycle analysis of medium-renewable hybrid
a solar plant, while in the USA the value is limited to 25% [10]. This systems. Klein et al. [74] have evaluated the performance of a
type of hybrid plants has been categorized as medium-renewable 100 MWe solar plant using parabolic trough with three backup
hybrids in the current work. The renewable energy fraction in their technologies and two cooling technologies. The performance was
power output is approximately 70%, which is lesser than high- compared in terms of life cycle greenhouse gas emissions, water
renewable hybrids (100%), but higher than low renewable consumption, and onsite land usage. The life cycle CO2 emission
hybrids (<20%). of plants without backup is 35 kg/MW h, which is 1.5–3 times
lower than with storage (60–73 kg/MW h). For plants with natural
3.2.1. Size of the solar field gas backup, the emission (127–317 kg/MW h) is 4–9 times higher
The effect of the size of the solar field (solar multiple) on the than without backup. A similar analysis for a 50 MWe parabolic
performance metrics of a medium-renewable hybrid plant trough CSP plant with 7.5-h storage has been performed by Corona
(Fig. 19) has been analyzed by Montes et al. [73]. The solar multiple et al. [2]. Fossil fuels such as coal, natural gas, fuel oil, and biofuels
is the ratio of the actual solar field aperture area to the area such as wheat straw, wood pellets, and biogas have been used for

Fig. 18. Plant configurations using solar concentrators with both thermal energy storage (TES) system and supplementary fossil fuel burner (adapted from [72]).
614 S. Pramanik, R.V. Ravikrishna / Applied Thermal Engineering 127 (2017) 602–637

backup. With 12% of the output energy being supplied by the 3.2.4. Other studies
backup, the maximum CO2 emission was obtained for coal Other studies on medium hybrids include that of Larrain et al.
(187 kg/MW h) and the minimum for wheat straw (34 kg/MW h). [82], who have analyzed the performance of a 100 MWe direct
The solar-only plant reported an emission of 26.9 kg/MW h. In steam generation (DSG) parabolic trough (PT) solar power plant
another study, Corona et al. [75] have analyzed the impact of fossil in Chile with fossil fuel backup. A model was developed to estimate
backup percentage on the environment. While CO2 emission the solar fraction and fossil fuel backup requirements for the plant
increased with increase in natural gas backup, the effects on the at four different locations. All the locations required fuel backup in
environment decreased slightly (Fig. 20) due to higher electricity the range of 40–80%, and the location with the least backup
outputs. requirement was determined to be the most suitable for construc-
tion of CSP plants. In another study, Larrain et al. [83] developed a
3.2.3. Performance comparison lifecycle model for a 100 MWe hybrid natural gas-CSP plant to
Several authors have compared the performance of medium determine the most suitable locations for the construction of such
renewable hybrids with standalone solar plants. Price et al. [76] plants in the Chilean Atacama Desert. Such a model is useful in
reported an improvement in the performance metrics of parabolic determining the locations at which a CSP plant becomes a net
trough hybrid plant with auxiliary natural gas firing in comparison energy source. Boukelia et al. [84] have optimized two parabolic
to the standalone plant. The solar-to-electricity efficiency, capacity trough plants using molten salt and Therminol VP-1 as the heat
factor and cost of electricity (COE) for the 30 MWe solar-only plant transfer fluid for Algerian conditions. The results have also been
were 10.6%, 22.2%, and 170 $/MW h, respectively. For the hybrid compared with the Andasol-1 plant. The molten salt plant was
plant with 25% backup, the values were 10.7%, 30.4%, and 141 $/ shown to be the best technology in terms of capacity factor, power
MW h. Similar results were obtained by Malagueta et al. [77,78] generation, water usage, investment cost, and cost of electricity.
who reported 50–70% decrease in COE and 100% increase in capac- Poghosyan et al. [85] tried to determine the size of thermal energy
ity factor due to natural gas backup. The impact of storage systems storage system that would be required by the Shams-I PT plant at
on standalone and hybrid plants has been analyzed by Wagner Madinat-Zayed, United-Arab-Emirates to replace the natural gas
et al. [79]. Their analysis shows that for small backup capacities heater. Their analysis showed that storage cannot completely
(1–4 h), the COE for plants with storage was lower as compared replace the heat transfer fluid heater within a reasonable size
to that of natural gas. For higher backup capacities (5–12 h), the (solar multiple of two, storage for 16 h).
COE of plants with storage had slightly higher. However, Dersch
et al. [80] reported that a solar plant with natural gas backup but 3.2.5. Operational plants and challenges
without storage offers higher efficiency. The effect of the heat Several operating solar plants use auxiliary natural gas burners
transfer fluid on the standalone and hybrid plant performance for heating purposes. Price et al. [15] have reviewed the nine Solar
has been analyzed by Boukelia et al. [81]. The highest energy effi- Electric Generating Systems (SEGS) (Table 3) located in the Mojave
ciency (18.5%) and lowest COE (75.9 $/MW h) was obtained for the Desert, California. It was reported that 8 of these 9 plants have nat-
molten salt plant with both storage and natural gas backup. The ural gas backup for evaporation and superheating. The authors pre-
thermic-oil plant with both storage and natural gas backup had dict that advanced parabolic trough plants in the future would
the highest exergy efficiency (21.8%), capacity factor (38.2%) and have the cost of electricity (COE) in the range of 50–60 $/MW h
annual power generation (165.8 GW h). that would directly compete with conventional fossil fuel power
plants. However, these systems report low solar-to-electricity effi-
ciency owing to the higher solar share. This increases the capital
investment and cost of electricity. More studies should focus on
the use natural gas backup in solar tower systems that can offer
greater solar-to-electricity and thermal efficiencies due to higher
operating temperatures. Transient energy and exergy analysis
(over a day/ month/ year) of these hybrids will also be useful in
optimizing the hourly fuel consumption and plant performance
parameters while determining the sources of exergy destruction.
Fig. 21 shows parts of two SEGS plants using solar tower and para-
bolic troughs with fossil fuel backup.

3.2.6. Summary
Medium renewable hybrids are the most common solar plants
Fig. 20. Variation of impact on the environment with natural gas backup as they can operate continuously. The trade-off between the solar
percentage [75]. field size (solar multiple) and percentage of fossil fuel backup

Table 3
Characteristics of SEGS I-IX [15]. (Acronym: HTF – heat transfer fluid).

SEGS plant First year of operation Net output (MWe) Solar field outlet temperature (°C) Solar field area (m2) Annual output (MW h) Backup
I 1985 13.8 307 82,960 30,100 Thermal storage (3 h)
II 1986 30 316 190,338 80,500 Gas-fired
superheater
III 1987 30 349 230,300 92,780 Gas-fired HTF boiler
IV 1987 30 349 230,300 92,780 Gas-fired HTF boiler
V 1988 30 349 250,500 91,820 Gas-fired HTF boiler
VI 1989 30 390 188,000 90,850 Gas-fired HTF boiler
VII 1989 30 390 194,280 92,646 Gas-fired HTF boiler
VIII 1990 80 390 464,340 252,750 Gas-fired HTF heater
IX 1991 80 390 483,960 256,125 Gas-fired HTF heater
S. Pramanik, R.V. Ravikrishna / Applied Thermal Engineering 127 (2017) 602–637 615

Brayton cycles (Fig. 22) in the power range of 1.4–4.2 MWe has
been performed by Schwarzbozl et al. [88]. The performance was
tested using three different turbines at two different locations,
namely Sevilla in Spain and Daggett in the US. The analysis showed
that the annual plant capacity factor decreased with an increase in
the solar share. A maximum solar share reported at 100% capacity
factor was 18.1%. However, the power production cost being higher
than that of conventional fossil fuel plants, the authors recom-
mended the introduction of these plants into distributed markets
(<10 MWe) along with cogeneration until the technology matures.
A similar study has been performed by Barigozzi et al. [89] in
which they have modeled the performance of a 36 MWe solar-
hybrid gas turbine system under the climatic conditions of
Palermo, Italy. The analysis revealed that such systems can achieve
a solar thermal share (Qsolar/Qin) of the order of 60–70% during the
peak sunny hours of the day in all seasons except winter. Kalatha-
kis et al. [90] have analyzed the performance of a CSP-Brayton sys-
tem for various spool arrangements of a 5 MWe gas turbine. The
single shaft recuperated turbine was the best choice for perfor-
mance, simplicity and cost considerations. The effect of storage
in a CSP-Brayton plant has been analyzed by Spelling et al. [91].
Under constant power operation, the addition of storage increased
the solar share while reducing the water consumption. Olivenza-
León et al. [92] have developed an analytical model for a CSP-
Brayton cycle considering the irreversibilities associated with a
Fig. 21. (a) 19.9 MWe Gemasolar plant [86] located in Seville, Spain using solar real plant. Their model could satisfactorily predict the performance
tower with 15% natural gas backup. (b) Part of the 354 MWe SEGS plant in northern
San Bernardino County, California using parabolic troughs with natural gas backup
of real plants.
[87].
3.3.1.2. Solarized steam injection gas turbines. Several studies on
solarized steam injection gas turbine (STIG) systems have been
determines the cost of electricity, efficiency, and CO2 emissions. performed by Kribus et al. [93–97]. In these systems (Fig. 23),
The plants with higher fossil fuel backup report higher efficiency steam generated using low-temperature solar energy is injected
and lower cost of electricity compared to standalone solar plants, into the combustor to increase the power output and efficiency.
although they have a higher negative impact on the environment. The exhaust from the turbine is recuperated by superheating the
Restrictions on the maximum usage of fossil fuel have resulted in solar-derived steam and preheating the water. The authors have
low solar-to-electricity for these systems. Also, currently they are performed energy, exergy and economic analysis for a wide range
not competitive with conventional power plants or with photo- of operating parameters under the climatic conditions of India and
voltaics. A tabulated summary of selected studies on medium Israel. The addition of 4 h storage to the STIG cycle [97] increases
renewable hybrids has been presented in Table A4 in the Appendix. the capacity factor beyond 50% while adding stability to the power
All existing, under construction and planned SEGS plants with fos- output. Other studies on solar steam injection gas turbines with
sil fuel backup have been summarized in Table A5 in the Appendix. carbon capture have been performed by Kosugi et al. [98–100]. A
novel steam injection gas turbine system with carbon capture
3.3. Low renewable hybrids and storage has been proposed by Gou et al. [101]. The saturated
steam generated from the solar collectors is superheated by burn-
Low renewable hybrids are fossil fuel power plants with the ing natural gas with pure oxygen before injection into the gas tur-
integration of solar energy for auxiliary purposes such as preheat- bine. Under design conditions, the system offers high solar-to-
ing. The solar share in low renewable hybrids is usually less than
20%.

3.3.1. Solar-Brayton cycles


Solar energy is used in Brayton cycles either to preheat the com-
pressed air before entering the combustion chamber or to generate
steam that is injected into the combustion chamber as the working
fluid. In both the cases, the increased inlet temperature to the com-
bustion chamber reduces the fuel consumption rate while increas-
ing the cycle efficiency. The higher working fluid temperature in a
Brayton cycle increases the solar-to-electricity efficiency in such
plants. This mode of integration enables continuous plant opera-
tion even in the absence of solar energy.

3.3.1.1. Solar preheating of compressed air. Solar preheating of com-


pressed air in a Brayton cycle has been investigated by several
authors. In such a configuration, the combustion chamber provides
the additional enthalpy required to reach the turbine inlet temper-
ature (950–1300 °C) from the receiver outlet temperature
(800–1000 °C). Numerical modeling of such recuperated solar Fig. 22. Solarized gas turbine plant schematic: recuperated Brayton cycle [88].
616 S. Pramanik, R.V. Ravikrishna / Applied Thermal Engineering 127 (2017) 602–637

Fig. 23. Layout of the solar hybrid steam injection gas turbine (STIG) cycle [93].

electricity efficiency (30.8%) and solar share (59.85%) at an average that was aimed at developing a 100 KWe prototype solar-hybrid
collector temperature of 272.8 °C. Zhang et al. [102] have proposed microturbine system for cogeneration. Kerosene, used in the SOL-
a 361 MWe recuperated gas turbine system in which solar energy GATE project was replaced by biodiesel to make the operation fully
is used to produce steam for the upgrading of methane to syngas. sustainable. At megawatt scales, the first pilot solarized gas turbine
The system offered superior performance in comparison to a gas plant with a capacity of 4.6 MWe was operated under the Solugas
turbine system without a solar assist. The solar-to-electricity effi- project [103,109] at Sanlucar la Mayor, Sevilla, Spain. Fig. 24 shows
ciency varies in the range of 25–30% and reduces CO2 emission that the power output of the turbine in the Solugas project can be
by around 20%. The COE of the system is around 59 $/MW h. maintained constant irrespective of the varying weather condi-
tions. A similar 1.4 MWe hybrid solar-gas turbine system was
3.3.1.3. Summary. The first prototype of a solar-powered gas tur- tested at the Thermis site, Targasonne, France under the PEGASE
bine system was tested under the SOLGATE project (Fig. 25) in project [110,111]. However, there are currently no operational
the CESA-1 tower at Plataforma Solar de Almería (PSA) in Spain CSP-Brayton hybrid plants in the world. A tabulated summary of
[104–106]. The primary objective of the project was to develop a selected studies on solar Brayton cycles has been presented in
pressurized volumetric receiver that could heat the air above Table A6 in the Appendix.
1000 °C for direct use in a gas turbine. In the hybrid operation, a
net power of 227 kWe was produced with a net efficiency of 3.3.2. Solar-aided coal power plants
18.2% and a solar share of 60%. The high-temperature receiver is Solar-aided coal-fired power plants utilize solar energy for pre-
promising and the authors expect that the system can operate with heating and boiling. Often, the solar energy replaces the bled-off
a minimum fuel usage of 5%. This was followed by the SOLHYCO steam used for feed water heating in a regenerative Rankine cycle.
(Solar Hybrid Power and Cogeneration Plants) project [107,108] Retrofitting a solar field to a fully depreciated Rankine cycle power

Fig. 24. Full-time operation characteristics of the Solugas project [103].


S. Pramanik, R.V. Ravikrishna / Applied Thermal Engineering 127 (2017) 602–637 617

Fig. 25. SOLGATE [106,112] and SOLUGAS projects [103]. (Acronyms: LT – low temperature, MT – medium temperature, HT: high temperature).

Fig. 26. Configurations of solar-aided Rankine cycle power plants in which the bled-off steam in the (a) low pressure and (b) high-pressure feedwater heaters are replaced by
solar energy [114].

plant is reported to be much more profitable compared to the inte- steam generation (DSG) and heat-transfer-fluid (HTF) technologies
gration of carbon capture technologies [113]. to a 300 MWe coal-fired power plant. The authors have reported a
peak solar-to-electricity efficiency (S-E g) of 27%. The integration
of solar energy into a steam power plant for feedwater preheating
3.3.2.1. Configurations for integration. The configuration for integra-
has been the focus of several studies. In one such study, Yan et al.
tion of solar energy into the steam power plant (Fig. 26) is critical
[118] analyzed the performance of several plants ranging from 200
in determining the performance of the hybrid plant. Odeh et al.
to 1000 MWe. The performance was observed to be sensitive to the
[115] have analyzed three such configurations in which the solar
size of the plant and the position of integration. The performance
energy is utilized for boiling, preheating, and both boiling and pre-
improved with an increase in the power plant capacity for the
heating. The study suggests that boiling is the best configuration to
same level of integrated solar power. The S-E g also improved with
reduce fuel consumption. In a similar study, Yang et al. [116,117]
an increase in the temperature of integration. For a 200 MWe coal-
has investigated several strategies for the integration of direct
618 S. Pramanik, R.V. Ravikrishna / Applied Thermal Engineering 127 (2017) 602–637

fired power plant, Yang et al. [119] reported S-E g of 25.5% and with solar energy was the most advantageous in terms of solar
36.6% at integration temperatures of 215 °C and 260 °C, respec- share, fossil fuel saving, and plant efficiency. A similar study using
tively. Similar studies on the utilization of solar energy for feedwa- linear Fresnel reflectors has been performed by Reddy et al. [130].
ter preheating has been performed by several authors [120–126]. The study reported 20% increase in instantaneous power output by
All these studies report a higher S-E g compared to standalone replacing the turbine bleed streams with solar feed water heating.
solar plants.
3.3.2.3. Effect of solar field size. The variation of plant output charac-
3.3.2.2. Effect of CSP technology. The type of CSP technology used for teristics with an increase in the solar field size (Fig. 27) has been
hybridization has a significant influence on the hybrid plant perfor- discussed by Wu et al. [131] and Hong-juan et al. [132]. The
mance. A comparison of direct steam generation (DSG) and heat- increase in solar field area enhances the annual plant output while
transfer-fluid (HTF) parabolic trough technologies applied to operating in the power boosting mode. In the fuel saving mode,
coal-fired subcritical (550 MWe) and supercritical (660 MWe) this causes a decrease in the annual coal consumption. The
steam power plants has been conducted by Suresh et al. [127]. increase in aperture area increases the solar-to-electric efficiency
The DSG technology was observed to be more cost effective as as it replaces the bled-off steam for preheating. However, beyond
compared to the HTF technology. Also, the hybrid supercritical an optimal area, the additional solar contribution is wasted as all
plant is reported to have performed better than the subcritical of the bled-off steam has already been replaced by solar energy.
plant in terms of thermal efficiency (45.5% and 41.8%, respectively). The higher solar field size incurs additional costs and hence the
The specific CO2 emission for the subcritical and supercritical cost of electricity (COE) increases beyond the optimum value. This
plants was 790 kg/MW h and 760 kg/MW h, respectively. The use can be observed in Fig. 27(c) and (d) where the solar-to-electricity
of solar energy for feed water heating was exergetically more effi- efficiency is maximized and the COE is minimized for an optimal
cient and led to a reduction in coal consumption by 14–19%. Inte- aperture area. This optimal area depends on the total capacity
gration of parabolic trough (PT) collectors into a 300 MWe lignite- (MWe) of the plant and hence poses a limit on the maximum solar
fired power plant has also been simulated by Bakos et al. [128]. The share in these systems. It is, therefore, necessary to determine uses
plant efficiency increased from 33% to 37.6% with the addition of of solar generated steam in a coal fired plant other than the
the solar field. Popov [129] studied the integration of linear Fresnel replacement of bled-off steam. Power cycles in which the solar
reflectors into a 130 MWe coal-fired power plant. The studies generated steam is directly used in the turbine will be a favorable
showed that the replacement of high-pressure feedwater heaters development and very few studies have addressed this issue, such

Fig. 27. Annual performance parameters with solar field area and storage capacity at 100% load [131].
S. Pramanik, R.V. Ravikrishna / Applied Thermal Engineering 127 (2017) 602–637 619

as [133]. Solar towers may be particularly attractive to generate for determining the solar contribution in a solar-aided coal-fired
the high-temperature steam. Critical issues such as mode of oper- power plant has been developed by Hou et al. [142]. They have
ation (power boost or fuel saving) and sizing of the components then used this method to study a 600 MWe coal power plant in
have to be addressed. Finally, we note that the cost of electricity which the steam extracted in the first stage is completely replaced
for a solar aided coal-fired power plant (98 $/MW h) is lower than by solar energy. The method proposed provides a theoretical refer-
that of a CSP-natural gas system (140 $/MW h) [116,117]. Peng ence in determining the solar contribution and thereby help poli-
et al. [134,135] have also reported the potential of hybrid plants cymakers in granting subsidies.
to reduce the COE by 20–30% in comparison to a standalone solar
plant. 3.3.2.5. Power plants. The Lidell power station at New South Wales,
Australia is the first solar aided coal-fired power plant [143,144]. It
3.3.2.4. Novel schemes and other studies. Several other studies have uses compact linear Fresnel reflectors to generate solar steam for
been performed on solar-aided coal-fired hybrid plants. Campore- feedwater heating. The solar capacity is around 3 MWe while the
ale et al. [136] performed thermodynamic analysis on the repow- plant capacity is 2000 MWe. Another hybrid plant using parabolic
ering of an existing 364 MWe steam power plant with solar trough collectors is the Colorado Integrated Solar Project located in
energy. They obtained a reduction in fuel consumption by 8% at Colorado, USA. The solar field produces around 2 MWe of electric-
the maximum solar irradiance. Wang et al. [114] have analyzed a ity [143]. However, both the above-mentioned plants are currently
300 MWe subcritical coal-fired power plant integrated with solar non-operational. Fig. 28 shows selected images of the two hybrid
collectors and post-combustion carbon capture. Among the various plants.
configurations studied, they found that the integration of medium
temperature solar energy to replace the high-pressure feedwater 3.3.2.6. Summary. Solar energy forms a small fraction of the total
heaters provided the highest solar-to-electricity efficiency of output in solar-aided coal power plants. The solar field can be used
around 24%. Gupta et al. [137] have compared the performance for boiling, preheating, or both boiling and preheating. Both direct
of a 50 kWe solar thermal power plant to that of a 220 MWe hybrid steam generation (DSG) and heat-transfer-fluid (HTF) technologies
plant using direct steam generation (DSG) technology and reported can be used, although the DSG technology is reported to be more
a higher efficiency for the hybrid system. Zhai et al. [138] have cost effective. There also exists an optimal solar field size for a
optimized the solar contribution in a 660 MWe solar-aided plant given coal power plant that uses solar energy to replace the
under different load conditions and direct normal irradiation bled-off steam. Beyond this optimum size, an increase in the solar
(DNI) levels. They proposed a model that calculated the solar con- field incurs additional costs that result in higher cost of electricity.
tribution in the solar-aided coal fired power plant at various load Finally, the cost of electricity from these hybrid plants is lower
conditions. In another study [139], the authors have analyzed the than that from CSP-natural gas systems. A tabulated summary of
performance of a 1000 MWe plant and concluded that hybridiza- selected studies on solar-aided coal-fired power plants has been
tion facilitates solar power generation at large capacities while presented in Table A7 in the Appendix.
reducing emissions from coal-fired power plants. The addition of
storage systems to such hybrid plants also improves the perfor- 3.3.3. Integrated solar combined cycles (ISCC)
mance [140]. Zhu et al. [141] have performed exergy analysis of Integrated solar combined cycle (ISCC) refer to combined cycle
a 1000 MWe coal-fired power plant fitted with solar towers. It systems with solar energy integration in the topping or the bot-
was observed that maximum exergy loss occurs in the boiler toming cycle. Integration of solar energy into a combined cycle is
(53.5%), followed by the solar field (26%). An exergy based method attractive as they offer higher efficiency in comparison to Brayton

Fig. 28. Images of (a) Lidell solar thermal station [145], (b) Fresnel reflectors at Lidell power plant [145], (c) parabolic trough solar field at the Integrated Solar Project in
Colorado [146].
620 S. Pramanik, R.V. Ravikrishna / Applied Thermal Engineering 127 (2017) 602–637

and Rankine cycle power plants. Several ISCC plants are currently Table 5
operational or under construction around the world (Table 4). Results for integration of solar energy to the topping and bottoming cycles of a
combined cycle plant [8,148].
The type of solar integration in a combined cycle plant has been
investigated by several authors. Integration can be performed with Plant Integration method gI (%) gII (%)
the topping cycle (similar to solar-Brayton plants), the bottoming A Both supplemental heat to the topping and 26.2 27.1
cycle (similar to solar-aided coal-fired plant), or both. Oda et al. bottoming cycle
[148] has analyzed all the three configurations (Fig. 29) and B Supplemental heat to the bottoming cycle only 29.8 30.9
C Supplemental heat to the topping cycle only 27.5 28.3
reported that solar integration to the bottoming cycle provides
higher overall efficiency (Table 5). Integration in the topping cycle
is also promising due to the higher efficiency of the solar compo-
nent [8]. Similar observations have been made by Barigozzi et al.
steam for the bottoming cycle in a heat recovery steam generator
[149], who have compared the performance of a combined cycle
(HRSG). Earlier studies on this configuration include SMUD
with solar integration in the topping and bottoming cycles. The
Kokhala [151], Kribus et al. [152], Segal and Epstein [153]. While
analysis suggests that higher solar-to-electricity efficiency is
solar towers are the predominant choice [154], Amelio et al.
obtained by solar integration in the topping cycle. However, inte-
[155] reported the use of air-linear parabolic trough collectors
gration in the bottoming cycle boosts power output. Buck et al.
for air-preheating in such systems. Low-temperature parabolic
[150] have compared the performance of a 30 MWe plant with
troughs (PT) can also be integrated into the topping cycle by gen-
solar heat integrated into the topping cycle with that of a
erating steam for injection into the combustion chamber (STIG,
310 MWe plant with solar heat integrated into the bottoming
[93]). One such steam injection combined cycle system
cycle. The topping cycle integration is reported to offer several
(46.6 MWe) has been proposed by Kakaras et al. [156]. The effi-
advantages over the bottoming cycle integration.
ciency of the proposed system (57.6%) was higher than conven-
tional combined cycle systems. A novel configuration of heat
3.3.3.1. Combined cycle with solar integration in the topping cycle. transfer fluid (HTF)-ISCC employing inlet air cooling of the gas tur-
Solar integration in the topping cycle (Fig. 30) is similar to the CSP- bine has been proposed by Popov [157]. The inlet air cooling is
Brayton plant, where the gas turbine exhaust is used to generate achieved either by a PV module or by an absorption chiller pow-

Table 4
Summary of existing ISCC plants [143,147].

Project name Location Technology Total output (MWe) Solar contribution (MWe) Status
Agua Prieta II Mexico Parabolic trough 478 14 Under construction
Ain Beni Mathar Morocco Parabolic trough 470 30 Operational
Archimede Italy Parabolic trough 765 5 Operational
ISCC Duba 1 Saudi Arabia Parabolic trough 600 43 Under construction
ISCC Hassi R’mel Algeria Parabolic trough 150 20 Operational
ISCC Kuraymat Egypt Parabolic trough 140 20 Operational
Martin US Parabolic trough 1150 75 Operational
Palmdale US Parabolic trough 570 50 Under development
Victorville 2 US Parabolic trough 563 50 Under development
Ningxia China Parabolic trough 92.5 92.5 Under development

Fig. 29. Possible configurations for solar energy integration into a combined cycle (adapted from [8]). (Acronyms: C – compressor, GT – gas turbine, HE – heat exchanger,
HRSG – heat recovery steam generator, ST – steam turbine, PT – parabolic trough).
S. Pramanik, R.V. Ravikrishna / Applied Thermal Engineering 127 (2017) 602–637 621

Fig. 30. Schematic diagram of an ISCC with integration in the topping cycle (adapted from [72]). (Acronyms: HS – hot storage, CS – cold storage, HE – heat exchanger, C –
compressor, GT – gas turbine, ST – steam turbine, HRSG – heat recovery steam generator).

Fig. 31. Schematic of solar syngas fired power plant (adapted from [163]). (Acronyms: C – compressor, GT – gas turbine).

ered by a direct steam generation (DSG) linear Fresnel solar field. of the hybrid plant are higher than that of combined cycle with car-
These configurations improve the overall thermal efficiency of bon capture and storage. The authors reported a solar thermal
the plant by 1.2% in comparison to HTF-ISCC. The authors also con- share of 28.2% and a net solar-to-electricity efficiency of 36.4%. A
cluded that the inclusion of the solar energy in the Brayton cycle is model for performance evaluation and optimization of a combined
much more advantageous than its integration into the steam cycle. cycle with hybrid solar reforming of methane has been developed
Another novel ISCC with a topping steam turbine plant supplied by by Sheu et al. [161]. Pilot-scale studies of solar reforming have
a parabolic trough system and a bottoming organic Rankine cycle been performed under the SOLASYS project in which a 250 kWe
has been proposed by Al-Sulaiman [158]. liquid-fuelled turbine was modified to operate on syngas [162].
3.3.3.1.1. Solar reforming. Solar reforming has also been studied by 3.3.3.1.2. Size of solar field. The performance of a hybrid plant as a
several authors for integration of solar energy into the topping function of the solar field capacity (Fig. 32) has been analyzed by
cycle (Fig. 31). Solar reforming is used to upgrade the fuel so that Saghafifar et al. [164]. The authors have optimized the heliostat
it can be utilized during periods of low solar insolation. The process field and thermo-economic parameters for a 50 MWe ISCC plant
is used to convert high-temperature solar energy into chemical located in the UAE. A solar tower system is integrated with the top-
energy by thermochemical processes such as methane steam ping cycle for air-preheating. The hybrid plant reported an annual
reforming (CH4+H2O = CO + 3H2), the water-gas shift reaction plant thermal efficiency of 47.16% with a solar share of 8.87%. The
(CO + H2O = CO2+H2) and steam or CO2 gasification of coal. The fuel specific CO2 emission decreased while the COE increased with
produced by this thermochemical process is a mixture of H2 and increase in the heliostat field capacity. The authors have also con-
CO, also known as syngas. Low and mid-temperature solar energy cluded that establishment of a new hybrid plant is more econom-
[470–570 K] can also be used for methanol reforming to generate ical compared to hybridization of an existing combined cycle plant.
syngas [159] that is oxidized in the combustion chamber of the The operating temperature of the receiver and the low efficiency of
topping cycle. The system can reach a solar-to-electricity efficiency the solar collector were determined to be the key bottlenecks for
of up to 35%. An advanced direct steam generation (DSG)-ISCC sys- hybridization. Similar thermo-economic optimization studies have
tem with solar thermochemical fuel conversion of methane has been performed by Spelling et al. [165] in which they have devel-
been studied by Li et al. [160]. The energy and exergy efficiencies oped a dynamic model for a combined cycle system using solar
622 S. Pramanik, R.V. Ravikrishna / Applied Thermal Engineering 127 (2017) 602–637

Fig. 32. Variation of (a) annual fuel consumption, specific CO2 emission, (b) COE and payback period with heliostat field capacity [164].

towers. The authors reported a cost of electricity (COE) in the range


of 120–240 $/MW h with solar-to-electricity efficiencies ranging
from 18 to 24%. However, at larger plant capacities, the auxiliary
power consumption is expected to reduce the plant performance.
In another study, Spelling et al. [166] have performed optimization
of an advanced ISCC plant with thermal storage.
Currently, there are no operational ISCC plants with solar inte-
gration in the topping cycle as the technology for high-
temperature high-pressure solar receivers is not well developed.

3.3.3.2. Combined cycle with solar integration in the bottoming


cycle. Solar integration in the bottoming cycle (Fig. 33) is similar
to solar-aided coal-fired plants. However, instead of using coal,
the exhaust of the topping gas turbine is used to generate steam
for the bottoming cycle. All ISCC plants (Table 4) currently incorpo-
rate solar energy in the bottoming cycle. This type of integration is
technologically mature and offers high reliability and low financial
risk [167] compared to the topping cycle integration.
3.3.3.2.1. Effect of climate on plant feasibility. The feasibility of ISCC
plants in specific climatic conditions has been analyzed in several
studies. Baghernejad et al. [169,170] have performed the energy
and exergy analysis of the ISCC power plant located in Yazd, Iran
with a view to optimize the investment cost, exergy destruction
and electricity cost using genetic algorithm. The exergy analysis
reveals that maximum exergy is lost in the combustor (29.62%),
followed by the solar field (9%) and the exhaust (7.78%). The major
sources of energy losses are the condenser (35.94%) and the stack
(10.15%). A similar study under Iranian conditions has been per-
formed by Hosseinia et al. [171] who have concluded that the cost
of electricity (COE) of ISCC is lower than that of the combined cycle
and gas turbine systems by 10% and 33%, respectively if environ- Fig. 33. Schematic of ISCC plant with solar integration in the bottoming cycle [168].
S. Pramanik, R.V. Ravikrishna / Applied Thermal Engineering 127 (2017) 602–637 623

mental costs are considered. Allani et al. [172] have studied the and technical details of Kuraymat ISCC, located in Egypt. Elsaket
integration of solar energy in a combined cycle under the climatic [179] has proposed the modification of the existing 4 x 51 MWe
conditions of Tunisia. They have concluded that the maximum Brayton cycles into an ISCC with DSG technology under Libyan cli-
power strategy of operation is better at CO2 mitigation as com- matic conditions. Such a modification would increase the plant
pared to maximum efficiency operation. Several authors have per- capacity to 286.12 MWe at the design point, thereby saving
formed thermodynamic analysis of the Hassi R’Mel ISCC power 151,260 t of fossil fuel annually and avoiding 468,910 t of CO2
plant in Algeria. Khaldi et al. [173] have found that the gas turbine per year. Li et al. [180] have studied the performance of a two-
combustor and the solar field are the major sources of energy and stage solar input DSG-ISCC under the climatic conditions of Yulin
exergy losses, whereas the turbine and the compressor exhibit high city, China.
efficiencies. The solar energy share is around 14% whereas the solar 3.3.3.2.2. Type of solar collector. The type of solar collector influ-
exergy share is 12%. The energy and exergy efficiencies at the ences the performance of the hybrid cycle. However, the optimum
design point are reported to be 56% and 53%, respectively. A study choice for collectors depends on several factors such as local cli-
by Derbal-Mokrane et al. [174] has shown that the ISCC plant at matic conditions, plant operating conditions, etc. Franchini et al.
Hassi R’Mel can produce 150 MWe at 52% efficiency, with a solar [181] have compared the performance of solar tower central recei-
share of 20%. Behar et al. [175] have calculated that the design ver system with that of parabolic troughs for the location of Seville,
point efficiency of the heat transfer fluid (HTF)-ISCC power plant Spain. The hourly transient analysis performed by them reveals
is better than that of the combined cycle. The feasibility of ISCC that parabolic troughs offer better performance as compared to
plants in southern Greece has been analyzed by Bakos et al. towers in the summer months with thermal efficiencies up to
[176]. El-Sayed [177] has studied the performance of a direct steam 60%. However, the yearly performance analysis shows that the
generation (DSG)-ISCC at Kuraymat, Egypt and concluded that the tower technology provides more energy at a higher efficiency as
power boosting mode is more economical compared to the fuel compared to the parabolic troughs. Horn et al. [182] have com-
saving mode. Brakmann et al. [178] have provided the design pared parabolic trough heat transfer fluid (HTF)-ISCC with air-

Fig. 34. Comparison of electricity production from a CCGT and an ISCC plant at Almería, Spain (top) and Las Vegas, US (bottom) [185].
624 S. Pramanik, R.V. Ravikrishna / Applied Thermal Engineering 127 (2017) 602–637

tower ISCC and concluded that the solar levellized cost of electric-
ity for the trough is lower as compared to that for the tower. How-
ever, the cost of electricity (COE) is same for both the parabolic
trough (PT) and solar tower (ST) systems (31 $/MW h), which is
higher than that of the reference combined cycle plant (24 $/
MW h). Manente et al. [183] have analyzed and compared the per-
formance of several ISCC plant configurations using solar collectors
such as parabolic trough, solar tower, and linear Fresnel reflectors.
The configuration using parabolic trough offered the best perfor-
mance with solar-to-electricity efficiency reaching 30%. Also,
Rovira et al. [184] have reported that direct steam generation
(DSG) offers better performance as compared to the heat transfer
fluid (HTF) system.
3.3.3.2.3. Performance comparison. Several studies have compared
the performance of ISCC with that of solar electric generating sys-
tems (SEGS or CSP-natural gas hybrid) and combined cycle gas tur-
bine (CCGT) systems. Montes et al. [185] have compared the Fig. 35. Trade-off between solar-to-electricity efficiency and solar share [189].

performance of a CCGT power plant with that of a natural gas-


ISCC power plant using direct steam generation (DSG) technology The system provided superior performance compared to one-
at Almería and Las Vegas (Fig. 34). In terms of cost of electricity stage ISCC and CCGT systems. Bonadies et al. [193] studied the ret-
(COE), the CCGT performed better than the ISCC plant at Almería, rofit of a solar tower and thermal storage system with a combined
whereas the ISCC plant performed better in Las Vegas. The hot cycle plant to make the plant operate for 17 h a day. The study
and dry climate of Las Vegas enhanced the performance of the ISCC. showed that the integration led to a lower running time for the
The global efficiency of the ISCC is better at Las Vegas (52.2%) com- gas turbine unit compared to that of a regular ISCC plant. Manente
pared to 51.9% for Almería. A comparison of CCGT and ISCC plants [168] developed a model to predict the performance of a 390 MWe
from energy and exergy considerations have been performed by CCGT plant that was later used for hybridization with solar energy.
Reddy et al. [186]. The analysis revealed that major energy losses The analysis suggests that an addition of 50 MWe solar power
took place in the condenser and heat recovery steam generator requires significant changes to the original plant in terms of equip-
(HRSG), whereas major exergy losses took place in the combustion ment and their capacities. Kane et al. [194] have studied the opti-
chamber followed by the HRSG. The energy and exergy efficiencies mization of the HRSG and HSSG to increase the exergy efficiency of
of the combined cycle (53.9% and 54.5%) are higher than that of the an ISCC plant. The results show that the cost of electricity with 15–
ISCC (41.7% and 49.7%). A comparative study of ISCC, SEGS, and 20% solar share is higher than similar capacity combined cycle
CCGT has been performed by Dersch et al. [80]. The study suggests plants by 20–30%. However, the cost may decrease with a reduc-
that ISCC has the highest efficiency of 68.6% in comparison to 34.7% tion in the cost of solar collectors and implementation of carbon
for a SEGS. However, the SEGS offers lower GHG emissions. Neza- credits. Gülen [195] has developed a system of equations based
mmahalleh et al. [187] have compared various configurations of on Second Law to optimize the performance of ISCC plants.
DSG-ISCC, HTF-ISCC with HTF-SEGS and concluded that DSG-ISCC 3.3.3.2.5. Summary. Several ISCC plants are currently operational
is the best configuration offering superior performance. Kolb or planned around the world. Fig. 36 shows selected images of four
[133] performed an economic comparison of several hybrids and ISCC plants located in several countries. There are currently no
standalone solar plants that use solar towers both in the power plants with solar integration in the topping cycle, although they
boost and fuel saving modes. The study showed that hybrid plants promise higher efficiency due to operation at higher temperatures
were more economical compared to solar only plants. Giuliano et al. and pressures. Solar reforming is also promising as the fuel can be
[72] compare the performance of 5 different ISCC configurations stored and used in a period with low solar insolation. For bottom-
with storage to that of a combined cycle plant. They have concluded ing cycle integration, all types of solar concentrators such as solar
that higher cost of fossil fuels and carbon trading can make ISCC towers, parabolic troughs, and Fresnel reflectors can be used. How-
competitive with CCGT plants in terms of COE. Turchi et al. [188] ever, the performance depends on parameters such as local cli-
have compared the performance of HTF-ISCC hybrid with and with- matic conditions, plant operating conditions, etc. Comparative
out storage to that of SEGS and concluded that the COE is lower for studies of ISCC plants with CCGT and SEGS systems suggest that
ISCC without storage than with storage. ISCC offers the highest efficiency. Several novel schemes for ISCC
3.3.3.2.4. Novel schemes and other studies. Several novel schemes systems have also been reported in the literature. A tabulated sum-
for ISCC with solar integration in the bottoming cycle have also mary of selected studies on ISCC plants has been presented in
been proposed in the literature. Gunasekaran et al. [189] have pro- Table A8 in the Appendix.
posed an advanced zero emission power (AZEP) cycle with carbon
capture and storage whose solar-to-electricity efficiency is compa- 4. Comparison of the hybrid technologies
rable with other technologies. They have also shown a negative
correlation between solar-to-electricity efficiency and solar share The data reported in the literature along with that of actual
(Fig. 35). Similar observations have been made by Kelly et al. power plants (Tables A1–A8) have been compiled for comparison
[190]. With increasing solar contributions, the solar energy dis- of the CSP-hybrids reviewed in this paper. The mean, minimum
placed some of the sensible heat transfer in the heat recovery and maximum of all the data reported is represented in figures
steam generator, leading to a decrease in the solar-to-electricity as described in the next few sections.
efficiency. Cau et al. [191] have proposed a novel ISCC concept
using CO2 as the heat transfer fluid (HTF). Currently, the system 4.1. CO2 emissions
cost is higher compared to combined cycle gas turbine (CCGT)
plants but holds promise for the future. Li et al. [192] have pro- Fig. 37 represents the specific CO2 emissions from the various
posed a novel cascade integrated solar combined cycle that uses CSP-hybrids. As can be observed, the high-renewable hybrids have
both parabolic trough collectors and evacuated tube collectors. the lowest specific CO2 emissions (<100 kg/MW h), followed by the
S. Pramanik, R.V. Ravikrishna / Applied Thermal Engineering 127 (2017) 602–637 625

Fig. 36. Images of ISCC plants (a) Ain Beni Mathar, Morocco [196] (b) Martin, US [197] (c) Kuraymat, Egypt [198] (d) Hassi R’Mel, Algeria [199].

MW h [203] and 14 kg/MW h [200], respectively. Using this data,


an estimate (using a linear combination) of CO2 emissions from a
CSP-geothermal hybrid plant with 15% solar share is 44.6 kg/
MW h. This is of the same order as reported for the hybrid CHP
plant [202]. However, the CO2 content of the geothermal brine is
high and can lead to high CO2 emissions in flash and dry steam
geothermal plants. The emission can vary from 4–740 kg/MW h,
with a weighted average of 122 kg/MW h [204]. This weighted
average is still of the same order as medium hybrid CSP-natural
gas plants. The geothermal fluid in a binary plant circulates
through a closed loop and hence emits less CO2 compared to flash
and dry steam plants. Rule et al. [205] reports a value of 40 kg/
MW h Wairakei binary geothermal plant in New Zealand.

4.2. Solar share

Fig. 37. Mean specific CO2 emissions from the CSP-hybrid technologies (bullets). The solar share in various CSP-hybrids is presented in Fig. 38.
The lines represent the range of emissions reported in the literature. (Acronyms: NG
The highest solar share is reported for SEGS plants that use natural
– natural gas, SEGS – solar electric generating system, ISCC – integrated solar
combined cycle). gas of the order of 10–25% for backup. CSP-biomass hybrid systems
using biomass for backup also report high solar share comparable
medium (<200 kg/MW h) and low renewable hybrids (>200 kg/ to CSP-natural gas systems. The configurations in which both bio-
MW h). The lowest CO2 emission is reported for CSP-wind hybrids, mass boiler and CSP field operate in parallel to supply steam to the
while the highest is reported for solar-aided coal-fired power turbine report low solar shares. CSP-geothermal hybrids use the
plants. The values were in the range of emissions reported for stan- geothermal brine for base power output and have a low solar
dalone plants [200,201]. While both CSP-natural gas and ISCC use share. The low hybrids mostly use solar energy for preheating
natural gas, the emission of the former is lower due to the higher and hence have a low solar contribution. This also corroborates
solar share. Thus, hybridization of CSP with conventional power with the high CO2 emission reported from such systems (Fig. 37).
plants such as natural gas combined cycle and coal-fired Rankine It can also be observed from the data that the systems with ther-
cycle does not offer a significant reduction in their specific CO2 mal energy storage have a higher solar share.
emissions.
No data for CO2 emissions from CSP-geothermal hybrid power 4.3. Capacity
plant was available in the literature. However, CO2 emissions for
a hybrid geothermal combined heat and power (CHP) plant using The various CSP-hybrid systems are suitable only in some lim-
evacuated tube collectors was reported in the range of 29.2– ited range of output capacities. CSP-biomass hybrids are restricted
38 kg/MW h [202]. This value has been presented in Fig. 37. The below 50 MWe [19] to ensure a continuous supply of biomass and
global warming potential reported for standalone geothermal bin- to reduce logistical costs. The capacity of standalone geothermal
ary plant and parabolic trough CSP plants is approximately 50 kg/ plants can vary from 0.2 to 68 MWe for binary systems to 4–
626 S. Pramanik, R.V. Ravikrishna / Applied Thermal Engineering 127 (2017) 602–637

mon infrastructure. CSP-biomass systems can reach higher effi-


ciencies (27.5% peak [22]) due to the attainment of higher
temperature from biomass combustion. The low temperature of
the geothermal resource (250 °C) limits the maximum plant
operating temperature, thereby reducing the solar-to-electricity
efficiency. CSP-natural gas systems reach mean S-E g in the range
of 10–20% with a maximum of 28.7 reported for the Ivanpah plant
[143]. For ISCC and solar-aided coal Rankine plants, the mean lies
between 20 and 30% with the maximum reaching in the range of
40–50%. The CSP-Brayton plants also report a higher mean S-E g
due to the incorporation of solar energy at elevated temperatures
in the gas turbine. Thus, the S-E g of low-renewable hybrids is
higher than the other categories. CSP-biomass systems can poten-
tially attain such high S-E g due to biomass combustion.

Fig. 38. Solar share in the power produced from various CSP-hybrids. The lines
represent the range of data reported in the literature. (Acronyms: NG – natural gas, 4.5. Overall energy and exergy efficiency
SEGS – solar electric generating system, ISCC – integrated solar combined cycle).
For wind, no data was available on the overall thermal effi-
330 MWe for flash systems [37]. The average capacity of 8.3 MW ciency of the hybrid plant. The exergy efficiency of such a plant
(Fig. 39) for CSP-geothermal hybrids is obtained as more hybridiza- was reported to be 19.2% in [70]. However, if we assume the CSP
tion studies have been performed for binary plants. Hybridization plant efficiency to be 15%, wind turbine efficiency of 32% [70],
studies for flash hybrids of capacity 20 MWe have also been and solar share to be 48% [70], the overall thermal efficiency
h i
reported [45]. Depending on the solar resource, larger CSP-
obtained is 20.73% 0:48þ1 0:52 ¼ 0:2073 . This has been shown in
geothermal hybrid plants can also be constructed. The capacity of 0:15 0:32

wind farms can be up to several hundred megawatts (Roscoe wind Fig. 41. The thermal efficiency of CSP-biomass hybrids is also high
farm, 781.5 MWe) with each turbine up to 10 MWe [206]. Thus, (>30%) as the plants are able to operate near the design conditions
CSP-wind hybrids can vary in capacity from 10 MWe to several by burning biomass. The CSP-geothermal hybrids report the lowest
100 MWe as observed in Fig. 39. Also, the wind farm does not con- efficiency (13%) due to the low temperature of the geothermal
strain the capacity of the CSP system as they share the power block source. The CSP-natural gas systems have a high solar share and
only. The CSP-natural gas plants can vary in capacity from 10 to hence report a low overall efficiency. The increase in the backup
500 MWe [143]. However, the average is lower (68.5 MWe) as sev- percentage from natural gas can increase the overall efficiency,
eral CSP-natural gas plants of 50 MWe are currently operational in although at the expense of higher CO2 emissions. The ISCC plants
Spain and other locations. The solar-aided coal-fired power plants report the highest overall efficiency, followed by the solar-
are typically of capacities ranging from 100 to 1000 MWe. The Brayton and solar-aided coal fired plants. These plants also have
solar-Brayton plants are limited in capacity due to the requirement the lowest solar share (Fig. 38) and are able to operate at the design
of high-temperature and high-pressure solar receivers that are cur- conditions continuously. The solar-Brayton reports slightly higher
rently unavailable. ISCC plants with solar integration in the topping efficiency than the coal-fired plants due to the higher temperature
cycle are also of smaller capacities due to the same reason. ISCC of operation. The ISCC plants report the highest efficiency due to
plants with solar integration in the bottoming Rankine cycle can the dual cycle configuration.
vary in capacity from 50 to 1000 MWe. The exergy efficiency reports similar trend as that of the energy
efficiency. The ISCC system reports the highest exergy efficiency of
60.9%, although the mean is higher for the solar-Brayton system as
4.4. Solar-to-electricity efficiency (S-E g)
only one data point is available in [114].
For CSP-wind hybrids, the solar-to-electricity efficiency is same
as that for standalone CSP plants (14%) [69] as they share only the 4.6. Capacity factor
power block (Fig. 40). This value will change as other modes of
integration are proposed in which CSP and wind share some com- The capacity factor of CSP-wind hybrids is the lowest due to the
intermittent nature of both the resources (Fig. 42). CSP-biomass

Fig. 39. Capacity of various CSP-hybrid systems. The mean capacities of some of the
CSP-hybrid systems have been shown in the plot. The lines represent the range of Fig. 40. Solar-to-electricity efficiency of various CSP-hybrid systems. The lines
data reported in the literature. (Acronyms: NG – natural gas, SEGS – solar electric represent the range of data reported in the literature. (Acronyms: NG – natural gas,
generating system, ISCC – integrated solar combined cycle). SEGS – solar electric generating system, ISCC – integrated solar combined cycle).
S. Pramanik, R.V. Ravikrishna / Applied Thermal Engineering 127 (2017) 602–637 627

bines (100%) while the capacity factors for steam injection gas
turbines (STIG) are in the range of 30–50%. For the STIG cycle ana-
lyzed in [94], the capacity factor is limited due to the operation of
the system only during peak sunshine hours. It can be increased by
the addition of storage [97] (capacity factor of 51.5%) or extended
to 100% by the continuous burning of fuel. This also implies that
systems with higher capacity factors have a low solar share.

4.7. Investment

The cost of CSP technologies is significantly higher (6000 $/


kWe) as compared to other conventional power production tech-
nologies. The CSP systems with storage such as Andasol I in Spain
cost 8700 $/kWe [207]. Thus, a CSP-hybrid system with large solar
share, such as the medium hybrids, is expected to incur high initial
investments (Fig. 43). The highest investment (12,000–18,000 $/
kWe) has been reported for CSP-geothermal hybrids [14,43]. These
reported values are significantly higher than those of standalone
CSP (6000 $/kWe) and standalone geothermal plants (1711 $/
kWe in [5]). One possible reason for such high specific investment
is the assumptions made in the calculation. For example, the
authors in [43] assume a specific investment of 20,000 $/kWe for
an enhanced geothermal system. This is clearly higher than the
values reported in [5,208], which is less than 4000 $/kWe. The
other renewable hybrids such as wind and biomass cost around
4000 $/kWe. The solar-aided plants require the lowest investment
(<1000 $/kWe) due to the maturity of the technology and wide-
spread usage.

Fig. 41. Overall energy (top) and exergy (bottom) efficiency of various CSP-hybrid
systems. The lines represent the range of data reported in the literature. (Acronyms:
NG – natural gas, SEGS – solar electric generating system, ISCC – integrated solar
combined cycle).

Fig. 43. Specific investment for various CSP-hybrid systems. The lines represent the
range of data reported in the literature. (Acronyms: NG – natural gas, SEGS – solar
electric generating system, ISCC – integrated solar combined cycle).

Fig. 42. Capacity factor of various CSP-hybrid systems. The lines represent the
range of data reported in the literature. (Acronyms: NG – natural gas, SEGS – solar
electric generating system, ISCC – integrated solar combined cycle).

hybrids have reported a maximum capacity factor of 51.4% [17] but


can reach higher values due to the stability offered by the combus-
tion of biomass. Geothermal plants can operate as a base load plant
due to the steady supply of geothermal brine and thus offer high
capacity factor (60–70%). The CSP-natural gas plants with storage
can reach capacity factors up to 74% (Gemasolar [10]), although
at a higher investment. The mean capacity factor of such systems
is in the range of 40–50%. The mean capacity factor of ISCC and
solar-aided coal plants is high (>80%) due to the low solar share
in such systems. It is interesting to note that solar-Brayton plants Fig. 44. Cost of electricity for various CSP-hybrid systems. The lines represent the
report a wide range of capacity factors, from 30 to 100%. The high- range of data reported in the literature. (Acronyms: NG – natural gas, SEGS – solar
est capacity factors have been reported for the recuperated gas tur- electric generating system, ISCC – integrated solar combined cycle).
628 S. Pramanik, R.V. Ravikrishna / Applied Thermal Engineering 127 (2017) 602–637

4.8. Cost of electricity (COE) advancement of technology, the COE of the low hybrids is expected
to rise due to increase in the price of fossil fuels driven by scarcity.
The cost of electricity is higher for the high and medium hybrids
as compared to the low hybrids (Fig. 44). The low-renewable 5. Summary
hybrids such as ISCC, solar-Brayton and solar-aided coal plants
have a low solar share (Fig. 38) and require significantly less In this work, literature pertaining to various CSP-hybrids has
investment (Fig. 43). For high and medium hybrids the mean val- been reviewed in detail. The hybrids have been categorized into
ues are close to 20 c/kW h, which is almost twice as that of the high, medium and low based on their renewable energy compo-
low hybrids. However, there is significant variability in the nent. Hybrids of CSP with wind, biomass, and geothermal energy
reported COE as it depends on the local conditions, their tariff have been categorized as high-renewable hybrids. The medium-
structure, subsidies, etc. For example, the COE for a CSP-biomass renewable hybrids refer to solar plants with natural gas backup.
hybrid at an archipelago in Indonesia has been reported to be The low-renewable hybrids encompass ISCC, solar-Brayton, and
60 c/kW h [18]. Similarly, the COE for a CSP-wind hybrid in the solar-aided coal Rankine plants that use solar energy for auxiliary
Skyros island of Greece has been reported to be 45.2 c/kW h [70]. purposes such as preheating and evaporation. The high-
Such locations are away from the national grid and have poor renewable hybrids report the least specific CO2 emissions
accessibility, leading to increase in the COE. However, the lowest (<100 kg/MW h), followed by the medium (<200 kg/MW h) and
COE values reported for the high and medium hybrids are close low hybrids (>200 kg/MW h). A comparative analysis of the
to that of the mean of the low hybrids. While the COE of the high hybrids has also been performed based on their plant characteris-
and medium hybrids are expected to decrease with time due to the tics and performance metrics such as capacity, solar share, solar-

Table A1
Summary of selected CSP-biomass hybrid plants reviewed in the present study. (Acronyms: PT – parabolic trough, ST – solar tower, gI – net thermal efficiency, gII – net exergy
efficiency, S-E g – solar-to-electricity efficiency, COE – cost of electricity).

Technology gI gII S-E g CO2 COE Investment Capacity (MWe) Capacity Solar Comments
(kg/MW h) (c/kW h) ($/kWe) factor (%) share (%)
PT + TES + wheat straw[2] 34.2 50 88 7.5 h TES
PT + TES + wood pellets[2] 37.5 50 88 7.5 h TES
Biomass + PT [17] 11.3 30 51.4 53.6
SolComBio [18] 29.6 23 4371 20 Brazil
24.8 60 8193.8 20 Indonesia
Biomass + PT [19] 22.6–34.1 3108–5624* 5–60 *
1 AU$ = 0.74 $
Biomass + ST [20] 33.4 11.5 4144* 30 23.9 *
1 AU$ = 0.74 $
PT + biomass [22] 25.6–27.5* 4514–5476** 50 *
Peak
**
1 AU$ = 0.74 $
Various hybrids [23] 29.3–33.2 3330–4958* 17.3–19.5 *
1 AU$ = 0.74 $
b-IGCC + PT in 36.7 70 MWth (gasifier)
bottoming [24] + 100 MWth CSP
IGSCC [26] 40.8 36
SGCC [28] 32.4 34.9 21.25 19 51.7 On-design
condition
SHCC [28] 30.9 33.3 17.28 20 49.3 On-design
condition
PT + bagasse [77] 19.7 4692 100 37.1 43 Table 13
Borges Termosolar [143] 30.5* 7684* 22.5 *
1 € = 1.13 $

Table A2
Summary of selected studies on CSP-geothermal hybrids reviewed in the present study. (Acronyms: PT – parabolic trough, ST – solar tower, ETC – evacuated tube collector, gI –
net thermal efficiency, gII – net exergy efficiency, S-E g – solar-to-electricity efficiency, COE – cost of electricity).

Technology gI gII S-E g CO2 COE Investment Capacity Capacity Solar Comments
(kg/kW h) (c/kW h) ($/kWe) (MWe) factor (%) share (%)
Flash + PT [14] 12.5 16,354* 8.4 For Australia
Flash + ST [14] 14.1 12,432* 9.9 1 AU$ = 0.74 $
Binary + PT [50] 8.6 16.5 61.8 27.9 Annual average quantities
9.4 19.8 62.2 22.8
9.1 26 68.3 18.3
8.8 31.3 71.1 14.8
Binary + PT [43] 12.4 17.2 16,950
17.1 16,750
17.7 18,480
Binary + PT [54] 8–11.8 11.6–14.1 6.6–13.7 1.4–5 Subcritical
9.2–10.8 8.9–14.7 -3.8–14.4 1–5.4 Supercritical
Binary + PT [53] 10.8–1.4 12.7–13.1 11–11.4 For superheat
17.4–18 12.2–12.7 11.6–12.1 For flash hybrid
Binary + PT [55] 14.5 9.6 Uses Kalina cycle
Flash + PT [45] 6.5 17.5 Single flash
5.7 20.2 Double flash
Binary + PT [39] 17.9
Binary + ETC [202] 29.2–38 Evacuated tube collector
S. Pramanik, R.V. Ravikrishna / Applied Thermal Engineering 127 (2017) 602–637 629

to-electricity efficiency, capacity factor, specific investment and emission due to high solar share. The capacity of such plants can
cost of electricity. range from 20 to 500 MWe and is suitable both for distributed
An extensive pool of literature exists for the low-renewable and centralized power systems. The capacity factor of these sys-
hybrids such as ISCC, solar-Brayton, and solar-aided coal Rankine tems can be greater than 70% with natural gas backup and thermal
power systems. For solar-Brayton with compressed air preheating, storage. However, these systems report low solar-to-electricity
the key challenge is the development and operation of high- efficiency that increases the investment and cost of electricity. Cur-
temperature high-pressure receivers. A similar challenge is faced rently, efficiency and cost are the primary constraints that prevent
by ISCC systems with topping cycle integration. For solar-aided further penetration of these systems into the market and should be
coal Rankine plants and ISCC with bottoming cycle integration, the focus of future research. Partial replacement of the backup nat-
the future studies should focus on methods of increasing the solar ural gas by renewable fuels such as syngas obtained from gasifica-
share to reduce CO2 emissions. In terms of other thermo-economic tion of local biomass can also reduce the CO2 emissions of these
parameters such as capacity factor, solar-to-electricity efficiency, systems.
investment, and cost of electricity, these systems offer superior The high-renewable hybrids such as CSP-wind, CSP-biomass,
performance over medium and high-renewable hybrids. and CSP-geothermal have minimum negative impact on the envi-
The medium renewable hybrids (CSP-natural gas) are also well ronment. This category is the least explored and presents tremen-
understood because of the operational experience gained from sev- dous potential for future research. However, such plants will be
eral such power plants. Such hybrids report moderate specific CO2 restricted to selected locations due to the requirement of two

Table A3
Summary of selected studies on CSP-wind hybrids reviewed in the present study. (Acronyms: WF – wind farm, PT – parabolic trough, TES – thermal energy storage, gI – net
thermal efficiency, gII – net exergy efficiency, S-E g – solar-to-electricity efficiency, COE – cost of electricity).

Technology gI gII S-E g CO2 COE Investment Capacity Capacity Solar Comments
(g/kW h) (c/kW h) ($/kWe) (MWe) factor (%) share (%)
WF + PT [68] 10.8–12.9 3733 100 38 33
WF + PT [69] 10.8 12.3 1508 20 800 MW wind + 708 MW PT
(equivalent to 200 MW wind)
* ** *
WF + PT+TES [70] 19.2 45.2 16.6 48 1 € = 1.13 $
**
1 CSP + 2 wind turbines

Table A4
Summary of selected studies on medium hybrids reviewed in the present study. (Acronyms: DSG – direct steam generation, PT – parabolic trough, NG – natural gas, ST – solar
tower, TES – thermal energy storage, gI – net thermal efficiency, gII – net exergy efficiency, S-E g – solar-to-electricity efficiency, COE – cost of electricity).

Technology gI gII S-E g CO2 COE Investment Capacity (MWe) Capacity Solar Comments
(kg/MW h) (c/kW h) ($/kWe) factor (%) share (%)
DSG PT + NG [82] 100 40–80% backup of natural gas
SEGS VI [76] 11 19.4 30 33 70
SEGS VIII [76] 10 16.4 80 28 70
SEGS [15] 10.7 14.1 30 30.4 75
SEGS [15] 14.1 9.6 50 39.6 75
DSG PT + NG [73] 27 8.9* 50 70.8 *
1 € = 1.13 $
25.6 9.6* 50 80.6
24.8 10.4* 50 84.8
23.9 11.3* 50 88.2
22.9 12.5* 50 91.4
PT + NG [74] 127–317 100 50 6 h NG
PT + TES + NG [2] 125 50 88 7.5 h TES
PT + TES + NG [81] 17.6 21.7 169.4 7.9 4703 50 32.4 Configuration 3
17.7 21.8 184 9.1 6904 50 38.2 Configuration 4
18.3 21.1 169.4 7.7 4003 50 29 Configuration 7
18.5 21.7 184 7.6 4752 50 34 Configuration 8
PT + TES + NG [84] 14.4 19.4 184 12.7 50 36.1 Batna
14.3 19.3 184 12.9 50 35.6 Setif
15.2 20.1 184 10.8 50 42.6 Djelfa
14.4 20.7 184 10.4 50 44.3 Ghardaia
14.7 20.8 184 9.6 50 48 Hassi R’mel
16.2 20.9 183.9 9.1 50 51.6 Adrar
16.4 21.4 183.9 8.5 50 54.6 Bechar
17 21.4 184 7.6 50 61.5 Tamanrasset
16.9 21.1 184 8.5 50 54.3 Amenas
16.6 20.8 184 8.1 50 57.4 Salah
PT + NG [77] 21.6 4638.5 100 41.3 47
PT + NG [80] 34.7 50 Without TES
32.6 50 TES
PT + NG + TES [75] 67.2 50 95
108 50 90
148 50 85
189 50 80
230 50 75
270 50 70
311 50 65
630 S. Pramanik, R.V. Ravikrishna / Applied Thermal Engineering 127 (2017) 602–637

renewable energy resources. While CSP-biomass hybrids are suit- The COE is highly variable depending on the resource, location, tar-
able for capacities less than 50 MWe, CSP-wind, and CSP- iff structure, etc. Further research is necessary to address several
geothermal hybrids can reach capacities greater than 100 MWe. issues pertaining to this class of hybrids. For CSP-biomass systems,
CSP-wind systems suffer from low capacity factor due to the inher- the primary challenge lies in the steady supply of biomass to loca-
ent fluctuation is wind and solar resource. The investment for high- tions with high DNI. Location-specific optimization of logistics,
renewable hybrids is moderately high due to medium solar share. development of burners for use of multiple feedstocks, use of gasi-

Table A5
Summary of all SEGS plants (operational/under construction/planned) with fossil fuel backup [143]. (Acronyms – gI – net thermal efficiency, gII – net exergy efficiency, S-E g –
solar-to-electricity efficiency, COE – cost of electricity).

Technology gI gII S-E g CO2 COE Investment Capacity Capacity Solar Comments
(kg/MW h) (c/kW h) ($/kWe) (MWe) factor (%) share (%)
Andasol-1 16 30.51 49.9 88
Andasol-2 16 30.51 49.9 88
Andasol-3 7119 50 85
Andasol-4 16 30.51 49.9 88
Arcosol 50 6114.2 49.9
Arenales 30.51 50 88
Aste 1A 15 30.51 50 88
Aste 1B 15 30.51 49.9 88
Astexol II 15 30.51 50
Bokpoort 11,300 50
Caceres 30.51 50 88
Casablanca 30.51 50 88
Dahan 13.7
EL REBOSO II 17 36.386 50
EL REBOSO III 17 50
Enerstar 30.51 50 88
Extresol-1 16 30.51 88
Extresol-2 16 30.51 49.9 88
Extresol-3 16 30.51 50 88
Gemasolar [10] 13060.3 19.9 74 85
Genesis 250
Guzmán 30.51 50 88
Helioenergy 1 30.51 50
Helioenergy 2 30.51 50
Helios I 30.51 50 85
Helios II 30.51 50 85
Ibersol 4520 50
Ivanpah 28.7 5835.5 377
La Africana 30.51 8746.2 50 85
La Dehesa 13.8 49.9 88
La Florida 13.8 50 88
La Risca 30.51 50 88
Lebrija 1 30.51 50 88
Manchasol-1 16 30.51 49.9 88
Manchasol-2 16 30.51 50 88
Mojave Solar 6400 250
Morón 30.51 6667 50
Nevada Solar One 3694.4 72 98
NOOR I 43.74 7359.1 160 1 Dirhams = 0.27 $
Olivenza 1 30.51 6418.4 50 88
Palen SEGS 500 98
Pedro de Valdivia 7250 360
Planta Solar 10 30.64424 11
Planta Solar 20 30.64424 20
Shams 1 6000 100
Solaben 1 30.51 50
Solaben 2 30.51 50
Solaben 3 30.51 50
Solaben 6 30.51 50
Solacor 1 30.51 50
Solacor 2 30.51 50
Solana 8000 250
SEGS II 30
SEGS III 30
SEGS IV 30
SEGS V 30
SEGS VII 30
SEGS IX 80
Solnova 1 50
Solnova 3 50
Solnova 4 50
Termesol 50 6114.2 49.9
Termosol 1 50
Termosol 2 50
S. Pramanik, R.V. Ravikrishna / Applied Thermal Engineering 127 (2017) 602–637 631

Table A6
Summary of selected studies on solar-Brayton cycles reviewed in the present study. (Acronyms: GT – gas turbine, gI – net thermal efficiency, gII – net exergy efficiency, S-E g –
solar-to-electricity efficiency, COE – cost of electricity).

Technology gI gII S-E g CO2 COE Investment Capacity Capacity Solar Comments
(kg/MW h) (c/kW h) ($/kWe) (MWe) factor (%) share (%)
Recuperated GT [88] 40.4 15.4 21.6* 1.4 100 15 *
1 € = 1.13 $
38.4 14.5 22.5* 1.4 100 18.1
35.9 14.2 11.3* 4.2 100 7.5
35.9 14.6 11.2* 4.2 100 8.8
Brayton + storage [91] 456 10.5 1897 37 17.8
451 13.7 3011 38.2 37.6
STIG [94] 39.1 36.9 17.6 37.9 Constant power
39.1 35.9 17.6 41.7
39 37.9 17.6 36.7
39 37.7 17.6 38
39.5 38.4 17.6 34.3 Variable power
39.9 39.2 17.6 35.7
39.8 40.9 17.6 30.9
40.2 42.1 17.6 30.7
STIG + storage [97] 39.2 37 17.6 51.5 Table 6, ref: RGT, real storage
STIG [100] 75.2 3.82 22.4
SOLRGT [102] 45.9 52.3 29.1 343 5.9 736 361 20.3

Table A7
Summary of selected studies on solar-aided coal fired power plants reviewed in the present study. (Acronyms: PT – parabolic trough, CR – Coal Rankine, LFR – linear Fresnel
reflector, ST – solar tower, gI – net thermal efficiency, gII – net exergy efficiency, S-E g – solar-to-electricity efficiency, COE – cost of electricity).

Technology gI gII S-E g CO2 COE Investment Capacity Capacity Solar Comments
(kg/MW h) (c/kW h) ($/kWe) (MWe) factor (%) share (%)
PT + CR [116,117] 27 9.8 300
PT + CR [119] 36.6 200
PT + CR [131] 21* 6.3 330 *
Annual, Fig. 8a
PT + CR [134] 21 12* 349 *
1 Yuan = 0.15 $
PT + SubcR [127] 41.8 38.2 790 3.62* 789* 500 85 14.4 *
INR 1 = 0.015 $
PT + SupcR [127] 45.5 41.6 760 3.67* 843* 660 85 18.9
CR [120] 30.12 500
PT + CR [121] 45.9 200
LFR + CR [129] 33.1 17.3 104 6.5 For case B
35.4 34 116 9.9 For case C
35.7 39.2 117 22.8 For case D
LFR + CR [130] 38.4 41.3 9.4 636 3.7 LPH
35.8 41.7 20.6 738 17 LPH and HPH
PT + CR [118] 36.6 219 10.3 Subcritical
36.24 334 8.53 Subcritical
40.26 653 8.23 Subcritical
41.22 655 7.91 Supercritical
35.61 646 8.31 Ultra-supercritical
PT + CR [142] 24.1 600 4.3
PT + CR [122] 24.2 225 10.6
PT + CR [124] 23.6 100
25 125
25.6 200
24.6 330
25.3 600
27.9 1000
PT + CR [125] 36 300 11.5
PT + CR [126] 3.4–18 600
*
PT + CR [128] 8.6 275 Fuel saving
*
1 € = 1.13 $
34.9 285 2.2 Power boost
*
35.2 290 4 1 € = 1.13 $
36 296 6.1
36.7 8.5* 302 7.9
PT + CR [139] 42.5 13.6 4.2 637.5 1000 5.8 Fuel saving
42.9 13.5 4 601.6 1058 5.4 Power boost
PT + CR [132] 26.2 7.1 358 12.5 Configuration 1
1 Yuan = 0.15 $
ST + CR [133] 8.7 450 Power boost low capacity factor
7.8 450 Power boost high capacity factor
2.9 350 Fuel saver base loaded
PT + CR [209] 45.3–48 5.1–28.5 600
632 S. Pramanik, R.V. Ravikrishna / Applied Thermal Engineering 127 (2017) 602–637

Table A8
Summary of selected studies on ISCC power plants reviewed in the present study. (Acronym: CC – combined cycle, ST – solar tower, GT – gas turbine, PT – parabolic trough, R –
Rankine, gI – net thermal efficiency, gII – net exergy efficiency, S-E g – solar-to-electricity efficiency, COE – cost of electricity).

Technology gI gII S-E g CO2 COE Investment Capacity Capacity Solar


(kg/MW h) (c/kW h) ($/kWe) (MWe) factor (%) share (%)
CC + ST [148] 26.2 27.1 86.5 Both topping and bottoming
29.8 30.9 86.5 Bottoming
27.5 28.3 86.5 Topping
CC + ST GT [149] 56.3 29.3 66.5 9.8 Topping cycle integration
CC + PT R [149] 28.7 10.8 66.5 4.7 Bottoming cycle integration
CC + ST GT [164] 47.2 371.9 7.8 50 100 8.9
CC + ST GT [151] 19.6 4.8 30 21 Daggett
19.3 4.9 30 15 Sacramento
CC + ST GT [152] 21.3 6–7 2588 34
CC + GT [156] 57.6 46.6
CC + PT R [157] 49.9 469 ISCC
CC + PT abs. ch [157] 51.7 469 Turbine air cooling using solar
CC + ST GT [150] 43.1 32.3 5.1 30 11.3 REFOS
CC + PT R [150] 53.5 30.9 3.7 310 4.1 ISCC
CC + PT GT [155] 47.4 367 Ta = 25 °C, gg2
CC + ST GT [88] 44.9 18.3 7.1* 16.1 100 11.6 *
1 € = 1.13 $
43.4 19.3 7.2* 16.1 100 16.2
43.9 19 7.8* 16.1 100 27.8
CC + ST reforming [161] 47.6 26.1 386 5.4
CC + PT R [167] 15.7 17.6 140
16.1 17.5 140
ISCC [183] 29.7 317.4 1270 440 PT
29.2 317.7 1275 440 PT
28.2 317.4 1291 440 PT
27.5 317.4 1411 440 ST
26.6 317.8 1206 440 LFR
26.8 317.6 1204 440 LFR
CC + PT R [172] 60.2 350.6 7 18
67.1 314.4 9 2303 26
60.3 350 7.8 20.3
68.9 306.1 10.5 31.2
57 367 6.6 13.4
62.4 338.3 8.6 2310 20
57.3 368.6 7.3 15.1
63 334.7 10 24
CC + PT R [194] 362.1 4.8 1900 100 66 14
CC + PT R [171] 50.9 2.2 506 407 79.3
51.6 2 564 444 73.7
50.9 2.3 573 444 76
CC + PT R [169] 46.2 45.6 467
CC + PT R [170] 46.8 5.3 400
CC + PT R [210] 56 53 160 12
CC + PT R [174] 52 150 20
CC + PT R [175] 67 157 15
CC + PT R [176] 5.2 883.2 54.5 Power boost mode
5.4 1127.8 59 1 € = 1.13 $
5.7 1346.7 63.8
6 1542.8 68.5
6.2 1717.9 73.1
6.5 1881.1 78
5.1 924.5 50 Fuel saving mode
5.4 1249.2 50 1 € = 1.13 $
5.6 1582.7 50
5.9 1931.6 50
6.1 2287.5 50
6.4 2640.3 50
CC + PT R [181] 48.8 18 89.1 Annual average parameters
CC + ST R [181] 50.5 21.8 89.3 Annual average parameters
* *
CC + PT R [191] 23.6 7.8 553.6 278.1 9.2 Configuration 1
*
1 € = 1.13 $
24.8 7.8* 554.5* 279.4 9.6 Configuration 2
*
1 € = 1.13 $
CC + PT R [185] 52.2 21.5 9.2* 220 1.2 Almería
*
1 € = 1.13 $
*
51.9 27.3 9.1 220 2.4 Las Vegas
*
1 € = 1.13 $
CC + PT R [177] 117 8.4 Power boosting mode
87 9.1 Fuel saver mode
CC + PT R [180] 50.9 590 73.6 Annual performance
CC + PT R [182] 410 3.1 894 127 79.8 9 PT
CC + ST R [182] 409 3.1 902 127 80 8.2 ST
CC + PT R [188] 46.4 16.1 13.6 3330.9 138 Hybrid without TES
S. Pramanik, R.V. Ravikrishna / Applied Thermal Engineering 127 (2017) 602–637 633

Table A8 (continued)

Technology gI gII S-E g CO2 COE Investment Capacity Capacity Solar


(kg/MW h) (c/kW h) ($/kWe) (MWe) factor (%) share (%)
48.3 16.2 15.4 4049.3 138 Hybrid with 3 h TES
47.7 16 12.9 3553.5 88 Hybrid with 4 h TES
CC + LFR R [186] 41.7 49.7 13.1 325 9.7
CC + PT R [80] 68.6 310 ISCC
68.1 310 ISCC with TES
CC + PT R [187] 474 7.5 451.5 73.1 14.8 DSG
486 7.6 444.8 73.5 15.1 HTF
CC + PT R [184] 50.1 44.6 130.2 HTFev
47.7 32.2 124 HTFph-ev
50.1 44.6 130.2 HTFev-sh
48 33.6 124.7 HTFph-ev-sh
50.1 44.6 130.2 DSGev
47.2 29.8 122.8 DSGph-ev
50.1 44.6 130.2 DSGev-sh
48.3 35.1 125.5 DSGph-ev-sh
CC + ST R [133] 8.7 450 Power boost low capacity factor
7.8 450 Power boost high capacity factor
3.2 350 Fuel saver CC
CC + PT + ET R [192] 53.4 60.9 30.8 6 816 594 77.4 Cascade ISCC
54.1 58.4 28.6 569 One-stage ISCC
CC + PT R [168] 54.6 26.1 440 1 Power boost 1
49.6 24.2 440 Power boost 2
51.4 28.9 440 Power boost 3
27.1–27.6 440 Fuel saving 1
24.9 440 Fuel saving 2
Archimede [143,147] 15.6 765 0.65
Agua Prieta II 478 2.9
Ain Beni Mathar 470 6.4
ISCC Duba 1 600 7.2
ISCC Hassi R’mel 150 13.3
ISCC Kuraymat 140 14.3
Martin 1150 6.5
Palmdale 570 8.8
Victorville 2 563 8.9

fied biomass (syngas) to reduce logistical costs, and hybridization [6] J.H. Peterseim, S. White, A. Tadros, U. Hellwig, Concentrated solar power
hybrid plants, which technologies are best suited for hybridisation?, Renew
with other novel solar cycles such as the supercritical CO2 Brayton
Energy 57 (2013) 520–532, http://dx.doi.org/10.1016/j.renene.2013.02.014.
cycle can be areas of focus for future research. For CSP-geothermal [7] H.L. Zhang, J. Baeyens, J. Degrève, G. Cacères, Concentrated solar power
systems, the idea of replenishing the geothermal well using solar plants: review and design methodology, Renew. Sustain. Energy Rev. 22
energy can be explored. These systems along with medium- (2013) 466–481, http://dx.doi.org/10.1016/j.rser.2013.01.032.
[8] E.J. Sheu, A. Mitsos, A.A. Eter, E.M.A. Mokheimer, M.A. Habib, A. Al-Qutub, A
renewable hybrids require favorable policies to compete with the review of hybrid solar-fossil fuel power generation systems and performance
low-renewable hybrids and attain technological maturity. metrics, J. Sol. Energy Eng. 134 (2012), http://dx.doi.org/10.1115/1.4006973,
041006-041006.
[9] F. Trieb, T. Fichter, M. Moser, Concentrating solar power in a sustainable
Appendix A future electricity mix, Sustain. Sci. 9 (2014) 47–60, http://dx.doi.org/10.1007/
s11625-013-0229-1.
[10] D.A. Baharoon, H.A. Rahman, W.Z.W. Omar, S.O. Fadhl, Historical
See Tables A1–A8. development of concentrating solar power technologies to generate clean
electricity efficiently – a review, Renew. Sustain. Energy Rev. 41 (2015) 996–
1027, http://dx.doi.org/10.1016/j.rser.2014.09.008.
References [11] P. Gilman, N. Blair, M. Mehos, C. Christensen, S. Janzou, C. Cameron, Solar
Advisor Model: User Guide for Version 2.0, 2008. <http://www.
[1] M. Van Aalst, N. Adger, D. Arent, J. Barnett, R. Betts, E. Bilir, J. Birkmann, J. nrel.gov/docs/fy08osti/43704.pdf>.
Carmin, D. Chadee, A. Challinor, M. Chatterjee, W. Cramer, D. Davidson, Y. [12] O. Behar, A. Khellaf, K. Mohammedi, S. Ait-Kaci, A review of integrated solar
Estrada, J.-P. Gattuso, Y. Hijioka, O.H. Guldberg, H.-Q. Huang, G. Insarov, R. combined cycle system (ISCCS) with a parabolic trough technology, Renew.
Jones, S. Kovats, P.R. Lankao, J.N. Larsen, I. Losada, J. Marengo, R. McLean, L. Sustain. Energy Rev. 39 (2014) 223–250, http://dx.doi.org/10.1016/j.
Mearns, R. Mechler, J. Morton, I. Niang, T. Oki, J.M. Olwoch, M. Opondo, E. rser.2014.07.066.
Poloczanska, H.-O. Pörtner, M.H. Redsteer, A. Reisinger, A. Revi, D. Schmidt, R. [13] M.S. Jamel, A. Abd Rahman, A.H. Shamsuddin, Advances in the integration of
Shaw, W. Solecki, D. Stone, J. Stone, K. Strzepek, A. Suarez, P. Tschakert, R. solar thermal energy with conventional and non-conventional power plants,
Valentini, S. Vicuna, A. Villamizar, K. Vincent, R. Warren, L. White, T. Renew. Sustain. Energy Rev. 20 (2013) 71–81, http://dx.doi.org/10.1016/j.
Wilbanks, P.P. Wong, G. Yoh, Climate change 2014: impacts, adaptation, rser.2012.10.027.
and vulnerability, Assess. Rep. 5 (2014) 1–76. [14] J.H. Peterseim, S. White, A. Tadros, U. Hellwig, Concentrating solar power
[2] B. Corona, G. San Miguel, Environmental analysis of a concentrated solar hybrid plants – enabling cost effective synergies, Renew. Energy 67 (2014)
power (CSP) plant hybridised with different fossil and renewable fuels, Fuel 178–185, http://dx.doi.org/10.1016/j.renene.2013.11.037.
145 (2015) 63–69, http://dx.doi.org/10.1016/j. fuel. 2014.12.068. [15] H. Price, D. Kearney, Reducing the cost of energy from parabolic trough solar
[3] The Paris Agreement, n.d.. <http://unfccc.int/paris_agreement/items/9485. power plants, in: Proc. ISEC 2003 2003 Int. Sol. Energy Conf., Hawaii, USA, n.d.
php> (accessed November 24, 2016). <http://citeseerx.ist.psu.edu/viewdoc/download?doi=10.1.1.114.7601&rep=
[4] M.Z. Jacobson, M.A. Delucchi, Providing all global energy with wind, water, rep1&type=pdf>.
and solar power, Part I: Technologies, energy resources, quantities and areas [16] J.D. Nixon, P.K. Dey, P.A. Davies, The feasibility of hybrid solar-biomass power
of infrastructure, and materials, Energy Policy 39 (2011) 1154–1169, http:// plants in India, Energy 46 (2012) 541–554, http://dx.doi.org/10.1016/j.
dx.doi.org/10.1016/j.enpol.2010.11.040. energy.2012.07.058.
[5] M.A. Delucchi, M.Z. Jacobson, Providing all global energy with wind, water, [17] R. Soria, J. Portugal-Pereira, A. Szklo, R. Milani, R. Schaeffer, Hybrid
and solar power, Part II: Reliability, system and transmission costs, and concentrated solar power (CSP)-biomass plants in a semiarid region: a
policies, Energy Policy 39 (2011) 1170–1190, http://dx.doi.org/10.1016/j. strategy for CSP deployment in Brazil, Energy Policy 86 (2015) 57–72, http://
enpol.2010.11.045. dx.doi.org/10.1016/j.enpol.2015.06.028.
634 S. Pramanik, R.V. Ravikrishna / Applied Thermal Engineering 127 (2017) 602–637

[18] M.B. da Fonseca, W.-R. Poganietz, H.-J. Gehrmann, Environmental and [44] C. Zhou, Figure of merit analysis of a hybrid solar-geothermal power plant,
economic analysis of SolComBio concept for sustainable energy supply in Engineering 5 (2013) 26–31, http://dx.doi.org/10.4236/eng.2013.51B005.
remote regions, Appl. Energy 135 (2014) 666–674, http://dx.doi.org/10.1016/ [45] J.M. Cardemil, F. Cortés, A. Díaz, R. Escobar, Thermodynamic evaluation of
j.apenergy.2014.07.057. solar-geothermal hybrid power plants in northern Chile, Energy Convers.
[19] J.H. Peterseim, A. Herr, S. Miller, S. White, D.A. O’Connell, Concentrating solar Manage. 123 (2016) 348–361, http://dx.doi.org/10.1016/j.
power/alternative fuel hybrid plants: annual electricity potential and ideal enconman.2016.06.032.
areas in Australia, Energy 68 (2014) 698–711, http://dx.doi.org/10.1016/j. [46] G. Manente, R. Field, R. DiPippo, J.W. Tester, M. Paci, N. Rossi, Hybrid solar-
energy.2014.02.068. geothermal power generation to increase the energy production from a
[20] J.H. Peterseim, A. Tadros, S. White, U. Hellwig, J. Landler, K. Galang, Solar binary geothermal plant, in: Vol. 4 Energy Syst. Anal. Thermodyn. Sustain.
tower-biomass hybrid plants – maximizing plant performance, Energy Proc. Combust. Sci. Eng. Nanoeng. Energy, Parts A B, 2011: pp. 109–119.
(2013) 1197–1206, http://dx.doi.org/10.1016/j.egypro.2014.03.129. doi:10.1115/IMECE2011-63665.
[21] C.M.I. Hussain, B. Norton, A. Duffy, Technological assessment of different [47] M. Ayub, A. Mitsos, H. Ghasemi, Thermo-economic analysis of a hybrid solar-
solar-biomass systems for hybrid power generation in Europe, Renew. binary geothermal powerplant, Energy 87 (2015) 326–335, http://dx.doi.org/
Sustain. Energy Rev. (2016), http://dx.doi.org/10.1016/j.rser.2016.08.016. 10.1016/j.energy.2015.04.106.
[22] J.H. Peterseim, A. Tadros, U. Hellwig, S. White, Increasing the efficiency of [48] Á. Lentz, R. Almanza, Solar-geothermal hybrid system, Appl. Therm. Eng. 26
parabolic trough plants using thermal oil through external superheating with (2006) 1537–1544, http://dx.doi.org/10.1016/j.applthermaleng.2005.12.008.
biomass, Energy Convers. Manage. 77 (2014) 784–793, http://dx.doi.org/ [49] Á. Lentz, R. Almanza, Parabolic troughs to increase the geothermal wells flow
10.1016/j.enconman.2013.10.022. enthalpy, Sol. Energy 80 (2006) 1290–1295, http://dx.doi.org/10.1016/
[23] J.H. Peterseim, U. Hellwig, A. Tadros, S. White, Hybridisation optimization of j.solener.2006.04.010.
concentrating solar thermal and biomass power generation facilities, Sol. [50] M. Astolfi, L. Xodo, M.C. Romano, E. Macchi, Technical and economical analysis
Energy 99 (2014) 203–214, http://dx.doi.org/10.1016/j.solener.2013.10.041. of a solar-geothermal hybrid plant based on an organic Rankine cycle,
[24] Y. Tanaka, S. Mesfun, K. Umeki, A. Toffolo, Y. Tamaura, K. Yoshikawa, Geothermics 40 (2011), http://dx.doi.org/10.1016/j.geothermics.2010.09.009.
Thermodynamic performance of a hybrid power generation system using [51] H. Ghasemi, A. Mitsos, A hybrid geothermal-solar power system: optimal
biomass gasification and concentrated solar thermal processes, Appl. Energy design and operation, in: Vol. 6A Energy, ASME, 2013, p. V06AT07A043.
160 (2015) 664–672, http://dx.doi.org/10.1016/j.apenergy.2015.05.084. doi:10.1115/IMECE2013-63576.
[25] T. Srinivas, B.V. Reddy, Hybrid solar-biomass power plant without energy [52] I. Mir, R. Escobar, J. Vergara, J. Bertrand, Performance analysis of a hybrid
storage, Case Stud. Therm. Eng. 2 (2014) 75–81, http://dx.doi.org/10.1016/j. solar-geothermal power plant in Northern Chile, in: World Renew. Energy
csite.2013.12.004. Congr., Linköping, Sweden, 2011, pp. 1281–1288. doi:10.3384/
[26] S. Pramanik, R.V. Ravikrishna, A novel syngas-fired hybrid heating source for ecp110571281.
solar-thermal applications: energy and exergy analysis, Appl. Therm. Eng. [53] A.D. Greenhut, J.W. Tester, R. Dipippo, R. Field, C. Love, K. Nichols, F. Batini, B.
109 (2016) 1011–1022. Price, G. Gigliucci, I. Fastelli, Solar-geothermal hybrid cycle analysis for low
[27] M. Gupta, S. Pramanik, R.V. Ravikrishna, Development of a syngas-fired enthalpy solar and geothermal resources, in: World Geotherm. Congr. 2010,
catalytic combustion system for hybrid solar-thermal applications, Appl. 2010, pp. 25–29.
Therm. Eng. 109 (2016) 1023–1030. [54] C. Zhou, Hybridisation of solar and geothermal energy in both subcritical and
[28] Q. Liu, Z. Bai, X. Wang, J. Lei, H. Jin, Investigation of thermodynamic supercritical organic Rankine cycles, Energy Convers. Manage. 81 (2014) 72–
performances for two solar-biomass hybrid combined cycle power 82, http://dx.doi.org/10.1016/j.enconman.2014.02.007.
generation systems, Energy Convers. Manage. 122 (2016) 252–262, http:// [55] J.G. Boghossian, Dual-temperature Kalina Cycle for Geothermal-Solar Hybrid
dx.doi.org/10.1016/j.enconman.2016.05.080. Power Systems, Massachusetts Institute of Technology, 2011. <http://dspace.
[29] U. Sahoo, R. Kumar, P.C. Pant, R. Chaudhury, Scope and sustainability of mit.edu/handle/1721.1/68995>.
hybrid solar-biomass power plant with cooling, desalination in [56] G. Dimarzio, L. Angelini, W. Price, C. Chin, S. Harris, The stillwater triple
polygeneration process in India, Renew. Sustain. Energy Rev. 51 (2015) hybrid power plant: integrating geothermal, solar photovoltaic and solar
304–316, http://dx.doi.org/10.1016/j.rser.2015.06.004. thermal power generation, in: World Geotherm. Congr., 2015, pp. 1–5.
[30] E.K. Burin, L. Buranello, P. Lo Giudice, T. Vogel, K. Görner, E. Bazzo, Boosting <https://pangea.stanford.edu/ERE/db/WGC/papers/WGC/2015/38001.pdf>.
power output of a sugarcane bagasse cogeneration plant using parabolic [57] Officials Celebrate Hybrid Geothermal-Solar Power Plant, n.d..
trough collectors in a feedwater heating scheme, Appl. Energy. 154 (2015) <http://solarindustrymag.com/officials-celebrate-hybrid-geothermal-solar-
232–241, http://dx.doi.org/10.1016/j.apenergy.2015.04.100. power-plant/> (accessed November 3, 2016).
[31] E.K. Burin, T. Vogel, S. Multhaupt, A. Thelen, G. Oeljeklaus, K. Görner, E. Bazzo, [58] C.E. Hoicka, I.H. Rowlands, Solar and wind resource complementarity:
Thermodynamic and economic evaluation of a solar aided sugarcane bagasse advancing options for renewable electricity integration in Ontario, Canada,
cogeneration power plant, Energy (2016), http://dx.doi.org/10.1016/j. Renew. Energy 36 (2011) 97–107, http://dx.doi.org/10.1016/j.
energy.2016.06.071. renene.2010.06.004.
[32] Termosolar Borges, n.d.. <http://www.cspworld.org/cspworldmap/ [59] S. Jerez, R.M. Trigo, A. Sarsa, R. Lorente-Plazas, D. Pozo-Vázquez, J.P.
termosolar-borges> (accessed September 13, 2016). Montávez, Spatio-temporal complementarity between solar and wind
[33] Borges Termosolar, n.d.. <http://www.nrel.gov/csp/solarpaces/project_detail. power in the Iberian Peninsula, Energy Proc. (2013) 48–57, http://dx.doi.
cfm/projectID=242> (accessed September 13, 2016). org/10.1016/j.egypro.2013.08.007.
[34] A. Cot, A. Amettler, J. Vall-Llovera, J. Aguilo, J.M. Arque, Termosolar borges: a [60] V. Khare, S. Nema, P. Baredar, Status of solar wind renewable energy in India,
thermosolar hybrid plant with biomass, in: Third Int. Symp. Energy from Renew. Sustain. Energy Rev. 27 (2013) 1–10, http://dx.doi.org/10.1016/j.
Biomass Waste, 2010. rser.2013.06.018.
[35] A biomass-solar hybrid plant begins operations in Spain, n.d.. <http:// [61] F. Monforti, T. Huld, K. Bódis, L. Vitali, M. D’Isidoro, R. Lacal-Arántegui,
bioenergycrops.com/blog/2013/01/23/a-biomass-solar-hybrid-plant-begins- Assessing complementarity of wind and solar resources for energy
operations-in-spain/> (accessed September 16, 2016). production in Italy. A Monte Carlo approach, Renew. Energy 63 (2014)
[36] Supply of 50,000 tons of forest biomass to Termosolar Borges, n.d.. <http:// 576–586, http://dx.doi.org/10.1016/j.renene.2013.10.028.
prensa.comsa.com/suministro-de-50-000-toneladas-de-biomasa-forestal-a- [62] A.Z. Sahin, Applicability of wind–solar thermal hybrid power systems in the
termosolar-borges/> (accessed September 13, 2016). Northeastern Part of the Arabian Peninsula, Energy Sources 22 (2000) 845–
[37] S.J. Zarrouk, H. Moon, Efficiency of geothermal power plants: a worldwide 850, http://dx.doi.org/10.1080/009083100300001645.
review, Geothermics 51 (2014) 142–153, http://dx.doi.org/10.1016/ [63] H.H. Chen, H.Y. Kang, A.H.I. Lee, Strategic selection of suitable projects for
j.geothermics.2013.11.001. hybrid solar-wind power generation systems, Renew. Sustain. Energy Rev. 14
[38] A. Franco, M. Villani, Optimal design of binary cycle power plants for water- (2010) 413–421, http://dx.doi.org/10.1016/j.rser.2009.08.004.
dominated, medium-temperature geothermal fields, Geothermics 38 (2009) [64] C. Kost, B. Pfluger, W. Eichhammer, M. Ragwitz, Fruitful symbiosis: why an
379–391, http://dx.doi.org/10.1016/j.geothermics.2009.08.001. export bundled with wind energy is the most feasible option for North
[39] H. Ghasemi, E. Sheu, A. Tizzanini, M. Paci, A. Mitsos, Hybrid solar-geothermal African concentrated solar power, Energy Policy 39 (2011) 7136–7145, http://
power generation: optimal retrofitting, Appl. Energy 131 (2014) 158–170, dx.doi.org/10.1016/j.enpol.2011.08.032.
http://dx.doi.org/10.1016/j.apenergy.2014.06.010. [65] R. Sioshansi, P. Denholm, Benefits of colocating concentrating solar power
[40] D. Wendt, G. Mines, C. Turchi, G. Zhu, Geothermal Risk Reduction via and wind, Sustain. Energy IEEE Trans. 4 (2013) 877–885, http://dx.doi.org/
Geothermal/Solar Hybrid Power Plants. Final Report, Idaho Falls, ID (United 10.1109/TSTE.2013.2253619.
States), 2015. doi:10.2172/1245529. [66] F.J. Santos-Alamillos, D. Pozo-Vázquez, J.A. Ruiz-Arias, L. Von Bremen, J.
[41] C. Turchi, G. Zhu, M. Wagner, T. Williams, Geothermal/solar hybrid designs: Tovar-Pescador, Combining wind farms with concentrating solar plants to
use of geothermal energy for csp feedwater heating, in: Geotherm. Resour. provide stable renewable power, Renew. Energy 76 (2015) 539–550, http://
Counc. 38th Annu. Meet., 2014, pp. 1–16. dx.doi.org/10.1016/j.renene.2014.11.055.
[42] C. Zhou, E. Doroodchi, I. Munro, B. Moghtaderi, A feasibility study on hybrid [67] H.M.I. Pousinho, J. Esteves, V.M.F. Mendes, M. Collares-Pereira, C. Pereira
solar–geothermal power generation, in: New Zeal. Geotherm. Work., 2011. Cabrita, Bilevel approach to wind-CSP day-ahead scheduling with spinning
<https://www.geothermal-energy.org/pdf/IGAstandard/NZGW/2011/64. reserve under controllable degree of trust, Renew. Energy 85 (2016) 917–927,
pdf>. http://dx.doi.org/10.1016/j.renene.2015.07.022.
[43] C. Zhou, E. Doroodchi, B. Moghtaderi, An in-depth assessment of hybrid solar- [68] B.D. Vick, T.A. Moss, Adding concentrated solar power plants to wind farms to
geothermal power generation, Energy Convers. Manage. 74 (2013) 88–101, achieve a good utility electrical load match, Sol. Energy 92 (2013) 298–312,
http://dx.doi.org/10.1016/j.enconman.2013.05.014. http://dx.doi.org/10.1016/j.solener.2013.03.007.
S. Pramanik, R.V. Ravikrishna / Applied Thermal Engineering 127 (2017) 602–637 635

[69] J.P. Reichling, F.A. Kulacki, Utility scale hybrid wind-solar thermal electrical [93] M. Livshits, A. Kribus, Solar hybrid steam injection gas turbine (STIG) cycle,
generation: a case study for Minnesota, Energy 33 (2008) 626–638, http://dx. Sol. Energy 86 (2012) 190–199, http://dx.doi.org/10.1016/
doi.org/10.1016/j.energy.2007.11.001. j.solener.2011.09.020.
[70] F. Petrakopoulou, A. Robinson, M. Loizidou, Simulation and evaluation of a [94] G. Polonsky, M. Livshits, A. Immanuel Selwynraj, S. Iniyan, L. Suganthi, A.
hybrid concentrating-solar and wind power plant for energy autonomy on Kribus, Annual performance of the solar hybrid STIG cycle, Sol. Energy 107
islands, Renew. Energy 96 (2016) 863–871, http://dx.doi.org/10.1016/j. (2014) 278–291, http://dx.doi.org/10.1016/j.solener.2014.05.022.
renene.2016.05.030. [95] A.I. Selwynraj, S. Iniyan, G. Polonsky, L. Suganthi, A. Kribus, Exergy analysis
[71] R. Adinberg, Simulation analysis of thermal storage for concentrating solar and annual exergetic performance evaluation of solar hybrid STIG (steam
power, Appl. Therm. Eng. 31 (2011) 3588–3594, http://dx.doi.org/10.1016/j. injected gas turbine) cycle for Indian conditions, Energy 80 (2015) 414–427,
applthermaleng.2011.07.025. http://dx.doi.org/10.1016/j.energy.2014.12.001.
[72] S. Giuliano, R. Buck, S. Eguiguren, Analysis of solar-thermal power plants with [96] A. Immanuel Selwynraj, S. Iniyan, G. Polonsky, L. Suganthi, A. Kribus, An
thermal energy storage and solar-hybrid operation strategy, J. Sol. Energy economic analysis of solar hybrid steam injected gas turbine (STIG) plant for
Eng. 133 (2011) 31007, http://dx.doi.org/10.1115/1.4004246. Indian conditions, Appl. Therm. Eng. 75 (2015) 1055–1064, http://dx.doi.org/
[73] M.J. Montes, A. Abánades, J.M. Martínez-Val, Performance of a direct steam 10.1016/j.applthermaleng.2014.10.055.
generation solar thermal power plant for electricity production as a function [97] G. Polonsky, A. Kribus, Performance of the solar hybrid STIG cycle with latent
of the solar multiple, Sol. Energy 83 (2009) 679–689, http://dx.doi.org/ heat storage, Appl. Energy 155 (2015) 791–803, http://dx.doi.org/10.1016/j.
10.1016/j.solener.2008.10.015. apenergy.2015.06.067.
[74] S.J.W. Klein, E.S. Rubin, Life cycle assessment of greenhouse gas emissions, [98] P.S. Pak, T. Hatikawa, Y. Suzuki, A hybrid power generation system utilizing
water and land use for concentrated solar power plants with different energy solar thermal energy with CO2 recovery based on oxygen combustion
backup systems, Energy Policy 63 (2013) 935–950, http://dx.doi.org/10.1016/ method, Energy Convers. Manage. 36 (1995) 823–826, http://dx.doi.org/
j.enpol.2013.08.057. 10.1016/0196-8904(95)00130-6.
[75] B. Corona, G. San Miguel, E. Cerrajero, Life cycle assessment of concentrated [99] P.S. Pak, Y. Suzuki, T. Kosugi, A CO2-capturing hybrid power-generation
solar power (CSP) and the influence of hybridising with natural gas, Int. J. Life system with highly efficient use of solar thermal energy, Energy (1997) 295–
Cycle Assess. 19 (2014) 1264–1275, http://dx.doi.org/10.1007/s11367-014- 299, http://dx.doi.org/10.1016/S0360-5442(96)00115-6.
0728-z. [100] T. Kosugi, P. Sik, Pak, Economic evaluation of solar thermal hybrid H2O
[76] H. Price, E. Lüpfert, D. Kearney, E. Zarza, G. Cohen, R. Gee, R. Mahoney, turbine power generation systems, Energy 28 (2003) 185–198, http://dx.doi.
Advances in parabolic trough solar power technology, J. Sol. Energy Eng. 124 org/10.1016/S0360-5442(02)00092-0.
(2002) 109, http://dx.doi.org/10.1115/1.1467922. [101] C. Gou, R. Cai, H. Hong, A novel hybrid oxy-fuel power cycle utilizing solar
[77] D. Malagueta, A. Szklo, B.S.M.C. Borba, R. Soria, R. Aragão, R. Schaeffer, R. thermal energy, Energy 32 (2007) 1707–1714, http://dx.doi.org/10.1016/j.
Dutra, Assessing incentive policies for integrating centralized solar power energy.2006.12.001.
generation in the Brazilian electric power system, Energy Policy 59 (2013) [102] N. Zhang, N. Lior, N.L. Na Zhang, Use of low/mid-temperature solar heat for
198–212, http://dx.doi.org/10.1016/j.enpol.2013.03.029. thermochemical upgrading of energy, Part I: Application to a novel chemically-
[78] D. Malagueta, A. Szklo, R. Soria, R. Dutra, R. Schaeffer, B.S. Moreira Cesar recuperated gas-turbine power generation (SOLRGT) system, J. Eng. Gas Turb.
Borba, Potential and impacts of concentrated solar power (CSP) integration in Power. 134 (2012) 72301, http://dx.doi.org/10.1115/1.4006083.
the Brazilian electric power system, Renew. Energy 68 (2014) 223–235, [103] R. Korzynietz, J.A. Brioso, A. Del Río, M. Quero, M. Gallas, R. Uhlig, M. Ebert, R.
http://dx.doi.org/10.1016/j.renene.2014.01.050. Buck, D. Teraji, Solugas – comprehensive analysis of the solar hybrid Brayton
[79] S.J. Wagner, E.S. Rubin, Economic implications of thermal energy storage for plant, Sol. Energy 135 (2016) 578–589, http://dx.doi.org/10.1016/
concentrated solar thermal power, Renew. Energy 61 (2014) 81–95, http://dx. j.solener.2016.06.020.
doi.org/10.1016/j.renene.2012.08.013. [104] U. Fisher, C. Sugarmen, A. Ring, J. Sinai, Gas turbine ‘‘solarization”-
[80] J. Dersch, M. Geyer, U. Herrmann, S.A. Jones, B. Kelly, R. Kistner, W. Ortmanns, modifications for solar/fuel hybrid operation, J. Sol. Energy Eng. 126 (2004)
R. Pitz-Paal, H. Price, Trough integration into power plants-a study on the 872, http://dx.doi.org/10.1115/1.1763602.
performance and economy of integrated solar combined cycle systems, [105] J. Sinai, C. Sugarmen, U. Fisher, Adaptation and modification of gas turbines
Energy 29 (2004) 947–959, http://dx.doi.org/10.1016/S0360-5442(03)00199- for solar energy applications, in: Vol. 5 Turbo Expo 2005, ASME, 2005, pp. 87–
3. 94. doi:10.1115/GT2005-68122.
[81] T.E. Boukelia, M.S. Mecibah, B.N. Kumar, K.S. Reddy, Investigation of solar [106] P. Heller, M. Pfänder, T. Denk, F. Tellez, A. Valverde, J. Fernandez, A. Ring, Test
parabolic trough power plants with and without integrated TES (thermal and evaluation of a solar powered gas turbine system, Sol. Energy 80 (2006)
energy storage) and FBS (fuel backup system) using thermic oil and solar salt, 1225–1230, http://dx.doi.org/10.1016/j.solener.2005.04.020.
Energy 88 (2015) 292–303, http://dx.doi.org/10.1016/j.energy.2015.05.038. [107] L. Amsbeck, R. Buck, P. Heller, J. Jedamski, R. Uhlig, Development of a tube
[82] T. Larraín, R. Escobar, J. Vergara, Performance model to assist solar thermal receiver for a solar-hybrid microturbine system, in: Proc. 14th Bienn. CSP
power plant siting in northern Chile based on backup fuel consumption, Solarpaces Symp., 2008. <http://elib.dlr.de/54223/1/Lars_Amsbeck_
Renew. Energy 35 (2010) 1632–1643, http://dx.doi.org/10.1016/j. Development_of_a_tube_receiver_for_a_solar-hybrid_microturbine_system_
renene.2010.01.008. Solarpaces2008_proceedings.doc.pdf>.
[83] T. Larraín, R. Escobar, Net energy analysis for concentrated solar power plants [108] SolHyCo (SOLar HYbrid power and COgeneration plants), n.d.. <http://www.
in northern Chile, Renew. Energy 41 (2012) 123–133, http://dx.doi.org/ psa.es/en/areas/ussc/grupoalta/projects/solhyco.php> (accessed November
10.1016/j.renene.2011.10.015. 21, 2016).
[84] T.E. Boukelia, M.S. Mecibah, B.N. Kumar, K.S. Reddy, Optimization, selection [109] M. Quero, R. Korzynietz, M. Ebert, A.A. Jiménez, A. del Río, J.A. Brioso, Solugas
and feasibility study of solar parabolic trough power plants for Algerian – operation experience of the first solar hybrid gas turbine system at MW
conditions, Energy Convers. Manage. 101 (2015) 450–459, http://dx.doi.org/ scale, Energy Proc. 49 (2014) 1820–1830, http://dx.doi.org/10.1016/
10.1016/j.enconman.2015.05.067. j.egypro.2014.03.193.
[85] V. Poghosyan, M.I. Hassan, Techno-economic assessment of substituting [110] P. Garcia, A. Ferriere, G. Flamant, P. Costerg, R. Soler, B. Gagnepain, Solar field
natural gas based heater with thermal energy storage system in parabolic efficiency and electricity generation estimations for a hybrid solar gas turbine
trough concentrated solar power plant, Renew. Energy 75 (2015) 152–164, project in France, J. Sol. Energy Eng. 130 (2008) 14502, http://dx.doi.org/
http://dx.doi.org/10.1016/j.renene.2014.09.025. 10.1115/1.2807211.
[86] Gemasolar: When the Sun Lights the Night, n.d.. <http://www. [111] B. Grange, C. Dalet, Q. Falcoz, F. Siros, A. Ferrière, Simulation of a hybrid solar
directindustry.com/emag/issue-6/industry-avant-garde-gemasolar- gas-turbine cycle with storage integration, Energy Proc. (2013) 1147–1156,
thermosolar-plant-1-0606.html> (accessed November 9, 2016). http://dx.doi.org/10.1016/j.egypro.2014.03.124.
[87] Solar Energy Generating Systems, n.d.. <https://en.wikipedia.org/wiki/Solar_ [112] Plataforma Solar de Almería: Annual Report, 2014. <http://www.psa.es/en/
Energy_Generating_Systems> (accessed November 9, 2016). techrep/2014/ANNUAL> (REPORT 2014.pdf).
[88] P. Schwarzbözl, R. Buck, C. Sugarmen, A. Ring, M.J. Marcos Crespo, P. Altwegg, [113] V.H. Dalvi, S.V. Panse, J.B. Joshi, Solar thermal technologies as a bridge from
J. Enrile, Solar gas turbine systems: design, cost and perspectives, Sol. Energy fossil fuels to renewables, Nat. Clim. Change. 5 (2015) 1007–1013, http://dx.
80 (2006) 1231–1240, http://dx.doi.org/10.1016/j.solener.2005.09.007. doi.org/10.1038/nclimate2717.
[89] G. Barigozzi, G. Bonetti, G. Franchini, A. Perdichizzi, S. Ravelli, Thermal [114] F. Wang, H. Li, J. Zhao, S. Deng, J. Yan, Technical and economic analysis of
performance prediction of a solar hybrid gas turbine, Sol. Energy 86 (2012) integrating low-medium temperature solar energy into power plant, Energy
2116–2127, http://dx.doi.org/10.1016/j.solener.2012.04.014. Convers. Manage. 112 (2016) 459–469, http://dx.doi.org/10.1016/j.
[90] C. Kalathakis, N. Aretakis, I. Roumeliotis, A. Alexiou, K. Mathioudakis, enconman.2016.01.037.
Investigation of different solar hybrid gas turbines and exploitation of [115] S.D. Odeh, M. Behnia, G.L. Morrison, Performance evaluation of solar thermal
rejected sun power, in: Vol. 3 Coal, Biomass Altern. Fuels, Cycle Innov. Electr. electric generation systems, Energy Convers. Manage. 44 (2003) 2425–2443,
Power; Ind. Cogener. Org. Rank. Cycle Power Syst., ASME, 2016: p. http://dx.doi.org/10.1016/S0196-8904(03)00011-6.
V003T06A017. doi:10.1115/GT2016-57700. [116] Y. Cui, Y. Yang, J. Chen, Utilization of solar energy in a coal-fired plant, in:
[91] J. Spelling, R. Guédez, B. Laumert, A thermo-economic study of storage Challenges Power Eng. Environ., Springer Berlin Heidelberg, Berlin,
integration in hybrid solar gas-turbine power plants, J. Sol. Energy Eng. 137 Heidelberg, 2007, pp. 1207–1212. doi:10.1007/978-3-540-76694-0_225.
(2015) 1–10, http://dx.doi.org/10.1115/1.4028142. [117] Y. Yang, Y. Cui, H. Hou, X. Guo, Z. Yang, N. Wang, Research on solar aided coal-
[92] D. Olivenza-León, A. Medina, A. Calvo Hernández, Thermodynamic modeling fired power generation system and performance analysis, Sci. China Ser. E
of a hybrid solar gas-turbine power plant, Energy Convers. Manage. 93 (2015) Technol. Sci. 51 (2008) 1211–1221, http://dx.doi.org/10.1007/s11431-008-
435–447, http://dx.doi.org/10.1016/j.enconman.2015.01.027. 0172-z.
636 S. Pramanik, R.V. Ravikrishna / Applied Thermal Engineering 127 (2017) 602–637

[118] Q. Yan, E. Hu, Y. Yang, R. Zhai, Evaluation of solar aided thermal power [143] NREL, n.d.. <http://www.nrel.gov/csp/solarpaces/> (accessed September 22,
generation with various power plants, Int. J. Energy Res. 35 (2011) 909–922, 2016).
http://dx.doi.org/10.1002/er.1748. [144] Liddell Solar Thermal Station, n.d.. <http://www.areva.com/arevasolar/
[119] Y. Yang, Q. Yan, R. Zhai, A. Kouzani, E. Hu, An efficient way to use medium-or- liblocal/docs/Areva_Liddell_flyer_HR.pdf>.
low temperature solar heat for power generation – integration into [145] Liddell Solar Thermal Station, n.d.. <http://www.cspworld.org/cspworldmap/
conventional power plant, Appl. Therm. Eng. 31 (2011) 157–162, http://dx. liddell-solar-thermal-station> (accessed November 29, 2016).
doi.org/10.1016/j.applthermaleng.2010.08.024. [146] First solar/coal hybrid power station up and running, n.d.. <http://
[120] E. Hu, Y. Yang, A. Nishimura, F. Yilmaz, A. Kouzani, Solar thermal aided power newatlas.com/first-solarcoal-hybrid-power-station/15728/> (accessed
generation, Appl. Energy 87 (2010) 2881–2885, http://dx.doi.org/10.1016/j. November 29, 2016).
apenergy.2009.10.025. [147] E. Okoroigwe, A. Madhlopa, An integrated combined cycle system driven by a
[121] Q. Yan, Y. Yang, A. Nishimura, A. Kouzani, E. Hu, Multi-point and multi-level solar tower: a review, Renew. Sustain. Energy Rev. 57 (2016) 337–350, http://
solar integration into a conventional coal-fired power plant, Energy Fuels 24 dx.doi.org/10.1016/j.rser.2015.12.092.
(2010) 3733–3738, http://dx.doi.org/10.1021/ef9012906. [148] S.D. Oda, H.H. Hashem, A case study for three combined cycles of a solar-
[122] Y. Zhao, H. Hong, H. Jin, Proposal of a solar-coal power plant on off-design conventional power generation unit, Sol. Wind Technol. 5 (1988) 263–270,
operation, J. Sol. Energy Eng. 135 (2013) 31005, http://dx.doi.org/10.1115/ http://dx.doi.org/10.1016/0741-983X(88)90023-9.
1.4023359. [149] G. Barigozzi, G. Franchini, a. Perdichizzi, S. Ravelli, Simulation of solarized
[123] Y. Zhao, H. Hong, H. Jin, Mid and low-temperature solar-coal hybridization combined cycles: comparison between hybrid GT and ISCC plants, Ceram.
mechanism and validation, Energy 74 (2014) 78–87, http://dx.doi.org/ Conc. Sol. Power Plants Control. Diagnostics Instrumentation; Educ. Electr.
10.1016/j.energy.2014.03.092. Power Fans Blowers. 4 (2013) V004T05A008. doi:10.1115/GT2013-95483.
[124] Y. Zhao, H. Hong, H. Jin, Evaluation criteria for enhanced solar-coal hybrid [150] R. Buck, T. Bräuning, T. Denk, M. Pfänder, P. Schwarzbözl, F. Tellez, Solar-
power plant performance, Appl. Therm. Eng. 73 (2014) 575–585, http://dx. hybrid gas turbine-based power tower systems (REFOS), J. Sol. Energy Eng.
doi.org/10.1016/j.applthermaleng.2014.08.003. 124 (2002) 2, http://dx.doi.org/10.1115/1.1445444.
[125] J. Qin, E. Hu, G.J. Nathan, The performance of a Solar Aided Power Generation [151] H.I. Price, H.W. Whitney, D.D. Beebe, SMUD Kokhala power tower study, in:
plant with diverse ‘‘configuration-operation” combinations, Energy Convers. Am. Soc. Mech. Eng. – International Sol. Energy Conf., 1996.
Manage. 124 (2016) 155–167, http://dx.doi.org/10.1016/j. [152] A. Kribus, R. Zaibel, D. Carey, A. Segal, J. Karni, A solar-driven combined cycle
enconman.2016.07.015. power plant, Sol. Energy 62 (1998) 121–129, http://dx.doi.org/10.1016/
[126] J. Li, X. Yu, J. Wang, S. Huang, Coupling performance analysis of a solar aided S0038-092X(97)00107-2.
coal-fired power plant, Appl. Therm. Eng. 106 (2016) 613–624. [153] A. Segal, M. Epstein, Optimized working temperatures of a solar central
[127] M.V.J.J. Suresh, K.S. Reddy, A.K. Kolar, 4-E (Energy, exergy, environment, and receiver, Sol. Energy 75 (2003) 503–510, http://dx.doi.org/10.1016/
economic) analysis of solar thermal aided coal-fired power plants, Energy j.solener.2003.08.036.
Sustain. Dev. 14 (2010) 267–279, http://dx.doi.org/10.1016/j. [154] S. Heide, U. Gampe, U. Orth, M. Beukenberg, B. Gericke, M. Freimark, U.
esd.2010.09.002. Langnickel, R. Pitz-Paal, R. Buck, S. Giuliano, Design and operational aspects
[128] G.C. Bakos, C. Tsechelidou, Solar aided power generation of a 300 MW lignite of gas and steam turbines for the novel solar hybrid combined cycle SHCC, in:
fired power plant combined with line-focus parabolic trough collectors field, Proc. ASME Turbo Expo 2010 Power Land, Sea, Air, 2010, pp. 465–474.
Renew. Energy 60 (2013) 1–8, http://dx.doi.org/10.1016/j. doi:10.1115/GT2010-22124.
renene.2013.05.024 Technical note. [155] M. Amelio, V. Ferraro, V. Marinelli, A. Summaria, An evaluation of the
[129] D. Popov, An option for solar thermal repowering of fossil fuel fired power performance of an integrated solar combined cycle plant provided with air-
plants, Sol. Energy 85 (2011) 344–349, http://dx.doi.org/10.1016/ linear parabolic collectors, Energy 69 (2014) 742–748, http://dx.doi.org/
j.solener.2010.11.017. 10.1016/j.energy.2014.03.068.
[130] V.S. Reddy, S.C. Kaushik, S.K. Tyagi, Exergetic analysis of solar concentrator [156] E. Kakaras, A. Doukelis, R. Leithner, N. Aronis, Combined cycle power plant
aided coal fired super critical thermal power plant (SACSCTPT), Clean with integrated low temperature heat (LOTHECO), Appl. Therm. Eng. (2004)
Technol. Environ. Policy 15 (2013) 133–145, http://dx.doi.org/10.1007/ 1677–1686, http://dx.doi.org/10.1016/j.applthermaleng.2003.10.030.
s10098-012-0492-3. [157] D. Popov, Innovative solar augmentation of gas turbine combined cycle
[131] J. Wu, H. Hou, Y. Yang, E. Hu, Annual performance of a solar aided coal-fired plants, Appl. Therm. Eng. 64 (2014) 40–50, http://dx.doi.org/10.1016/j.
power generation system (SACPG) with various solar field areas and thermal applthermaleng.2013.12.002.
energy storage capacity, Appl. Energy 157 (2015) 123–133, http://dx.doi.org/ [158] F.A. Al-Sulaiman, Exergy analysis of parabolic trough solar collectors
10.1016/j.apenergy.2015.08.022. integrated with combined steam and organic Rankine cycles, Energy
[132] H. Hong-juan, Y. Zhen-yue, Y. Yong-ping, C. Si, L. Na, W. Junjie, Performance Convers. Manage. 77 (2014) 441–449, http://dx.doi.org/10.1016/j.
evaluation of solar aided feedwater heating of coal-fired power generation enconman.2013.10.013.
(SAFHCPG) system under different operating conditions, Appl. Energy 112 [159] H. Hong, H. Jin, J. Ji, Z. Wang, R. Cai, Solar thermal power cycle with
(2013) 710–718, http://dx.doi.org/10.1016/j.apenergy.2013.05.062. integration of methanol decomposition and middle-temperature solar
[133] G.J. Kolb, Economic evaluation of solar-only and hybrid power towers using thermal energy, Sol. Energy 78 (2005) 49–58, http://dx.doi.org/10.1016/
molten-salt technology, Sol. Energy 62 (1998) 51–61, http://dx.doi.org/ j.solener.2004.06.019.
10.1016/S0038-092X(97)00075-3. [160] Y. Li, N. Zhang, R. Cai, Low CO2-emissions hybrid solar combined-cycle power
[134] S. Peng, Z. Wang, H. Hong, D. Xu, H. Jin, Exergy evaluation of a typical 330 system with methane membrane reforming, Energy 58 (2013) 36–44, http://
MW solar-hybrid coal-fired power plant in China, Energy Convers. Manage. dx.doi.org/10.1016/j.energy.2013.02.005.
85 (2014) 848–855, http://dx.doi.org/10.1016/j.enconman.2013.12.073. [161] E.J. Sheu, A. Mitsos, Optimization of a hybrid solar-fossil fuel plant: solar
[135] S. Peng, H. Hong, Y. Wang, Z. Wang, H. Jin, Off-design thermodynamic steam reforming of methane in a combined cycle, Energy 51 (2013) 193–202,
performances on typical days of a 330 MW solar aided coal-fired power plant http://dx.doi.org/10.1016/j.energy.2013.01.027.
in China, Appl. Energy 130 (2014) 500–509, http://dx.doi.org/10.1016/j. [162] C. Sugarmen, A. Rotstein, U. Fisher, J. Sinai, Modification of gas turbines and
apenergy.2014.01.096. operation with solar produced syngas, J. Sol. Energy Eng. 126 (2004) 867,
[136] S.M. Camporeale, A. Saponaro, Repowering of a Rankine cycle power plant by http://dx.doi.org/10.1115/1.1758725.
means of concentrating solar collectors, in: Proc. ASME Turbo Expo, 2011, pp. [163] R. Tamme, R. Buck, M. Epstein, U. Fisher, C. Sugarmen, Solar upgrading of
1–8. fuels for generation of electricity, J. Sol. Energy Eng. 123 (2001) 160–163,
[137] M.K. Gupta, S.C. Kaushik, Exergetic utilization of solar energy for feed water http://dx.doi.org/10.1115/1.1353177.
preheating in a conventional thermal power plant, Int. J. Energy Res. 33 [164] M. Saghafifar, M. Gadalla, Thermo-economic analysis of conventional
(2009) 593–604, http://dx.doi.org/10.1002/er.1500. combined cycle hybridization: United Arab Emirates case study, Energy
[138] R. Zhai, M. Zhao, K. Tan, Y. Yang, Optimizing operation of a solar-aided coal- Convers. Manage. 111 (2016) 358–374, http://dx.doi.org/10.1016/j.
fired power system based on the solar contribution evaluation method, Appl. enconman.2015.12.016.
Energy 146 (2015) 328–334, http://dx.doi.org/10.1016/j. [165] J. Spelling, D. Favrat, A. Martin, G. Augsburger, Thermoeconomic optimization
apenergy.2015.01.140. of a combined-cycle solar tower power plant, Energy 41 (2012) 113–120,
[139] R. Zhai, H. Liu, C. Li, M. Zhao, Y. Yang, Analysis of a solar-aided coal-fired http://dx.doi.org/10.1016/j.energy.2011.03.073.
power generation system based on thermo-economic structural theory, [166] J. Spelling, B. Laumert, T. Fransson, Advanced hybrid solar tower combined-
Energy 102 (2016) 375–387, http://dx.doi.org/10.1016/j.energy.2016.02.086. cycle power plants, Energy Proc. (2013) 1207–1217, http://dx.doi.org/
[140] R. Zhai, C. Li, Y. Chen, Y. Yang, K. Patchigolla, J.E. Oakey, Life cycle assessment 10.1016/j.egypro.2014.03.130.
of solar aided coal-fired power system with and without heat storage, Energy [167] C.S. Turchi, Z. Ma, M. Erbes, Gas turbine/solar parabolic trough hybrid
Convers. Manage. 111 (2016) 453–465, http://dx.doi.org/10.1016/j. designs, in: Proc. ASME Turbo Expo, ASME, 2011, pp. 989–996. doi:10.1115/
enconman.2015.12.053. GT2011-45184.
[141] Y. Zhu, R. Zhai, H. Peng, Y. Yang, Exergy destruction analysis of solar tower [168] G. Manente, High performance integrated solar combined cycles with
aided coal-fired power generation system using exergy and advanced minimum modifications to the combined cycle power plant design, Energy
exergetic methods, Appl. Therm. Eng. 108 (2016) 339–346, http://dx.doi. Convers. Manage. 111 (2016) 186–197, http://dx.doi.org/10.1016/j.
org/10.1016/j.applthermaleng.2016.07.116. enconman.2015.12.079.
[142] H. Hou, Z. Xu, Y. Yang, An evaluation method of solar contribution in a solar [169] A. Baghernejad, M. Yaghoubi, Exergy analysis of an integrated solar combined
aided power generation (SAPG) system based on exergy analysis, Appl. cycle system, Renew. Energy 35 (2010) 2157–2164, http://dx.doi.org/
Energy 182 (2016) 1–8, http://dx.doi.org/10.1016/j.apenergy.2016.08.109. 10.1016/j.renene.2010.02.021.
S. Pramanik, R.V. Ravikrishna / Applied Thermal Engineering 127 (2017) 602–637 637

[170] A. Baghernejad, M. Yaghoubi, Exergoeconomic analysis and optimization of Washington (DC), 2001: pp. 393–398. <http://www.researchgate.net/
an integrated solar combined cycle system (ISCCS) using genetic algorithm, publication/229000659_Optimization_studies_for_
Energy Convers. Manage. 52 (2011) 2193–2203, http://dx.doi.org/10.1016/j. integrated_solar_combined_cycle_systems/file/e0b4951badbf2a9af4.pdf%
enconman.2010.12.019. 5Cnpapers2://publication/uuid/07203F08-07BE-4B8E-96AB-
[171] R. Hosseini, M. Soltani, G. Valizadeh, Technical and economic assessment of A6A7554B42DC>.
the integrated solar combined cycle power plants in Iran, Renew. Energy 30 [191] G. Cau, D. Cocco, V. Tola, Performance and cost assessment of Integrated Solar
(2005) 1541–1555, http://dx.doi.org/10.1016/j.renene.2004.11.005. Combined Cycle Systems (ISCCSs) using CO2 as heat transfer fluid, Sol. Energy
[172] Y. Allani, D. Favrat, M. von Spakovsky, CO2 mitigation through the use of 86 (2012) 2975–2985, http://dx.doi.org/10.1016/j.solener.2012.07.004.
hybrid solar-combined cycles, Energy Convers. Manage. 38 (1997) S661– [192] Y. Li, J. Yuan, Y. Yang, Performance analysis of a novel cascade integrated solar
S667, http://dx.doi.org/10.1016/S0196-8904(97)00012-5. combined cycle system, Energy Proc. (2015) 540–546, http://dx.doi.org/
[173] F. Khaldi, Energy and exergy analysis of the first hybrid solar-gas power plant 10.1016/j.egypro.2015.07.449.
in Algeria, in: Proc. ECOS, 2012. <https://www.researchgate.net/profile/ [193] M.F. Bonadies, M. Mohagheghi, M. Ricklick, J.S. Kapat, Solar retrofit to
Fouad_Khaldi/publication/ combined cycle power plant with thermal energy storage, in: Proc. ASME
257235961_Energy_and_exergy_analysis_of_the_first_hybrid_solar- TURBO EXPO 2010, vOL 3, 2010, pp. 921–931.
gas_power_plant_in_Algeria/links/54da22840cf2970e4e7dc9c6.pdf>. [194] M. Kane, D. Favrat, K. Ziegler, Y. Allani, Thermoeconomic analysis of advanced
[174] H. Derbal-Mokrane, S. Bouaichaoui, N. El Gharbi, M. Belhamel, A. Benzaoui, solar-fossil combined power plants, Int. J. Appl. Thermodyn. 3 (2000) 191–
Modeling and numerical simulation of an integrated solar combined cycle 198.
system in Algeria, in: Procedia Eng., 2012, pp. 199–208. doi:10.1016/j. [195] S.C. Gülen, Second law analysis of integrated solar combined cycle power
proeng.2012.01.1194. plants, J. Eng. Gas Turb. Power. 137 (2015) 51701, http://dx.doi.org/10.1115/
[175] O. Behar, A. Kellaf, K. Mohamedi, M. Belhamel, Instantaneous performance of 1.4028741.
the first integrated solar combined cycle system in Algeria, Energy Proc. [196] Abengoa, n.d.. <http://www.abengoa.com/web/en/noticias_y_publicaciones/
(2011) 185–193, http://dx.doi.org/10.1016/j.egypro.2011.05.022. noticias/historico/2014/12_diciembre/abg_20141215.html> (accessed
[176] G.C. Bakos, D. Parsa, Technoeconomic assessment of an integrated solar November 30, 2016).
combined cycle power plant in Greece using line-focus parabolic trough [197] Martin Next Generation Solar Energy Center, n.d.. <http://www.cspworld.
collectors, Renew. Energy 60 (2013) 598–603, http://dx.doi.org/10.1016/j. org/cspworldmap/martin-next-generation-solar-energy-center> (accessed
renene.2013.05.025. November 30, 2016).
[177] M.A.H. El-Sayed, Solar supported steam production for power generation in [198] ISCC’s outlook brightens with wave of projects, n.d.. <http://www.
Egypt, Energy Policy 33 (2005) 1251–1259, http://dx.doi.org/10.1016/j. powerengineeringint.com/articles/print/volume-19/issue-11/features/isccs-
enpol.2003.11.021. outlook-brightens-with-wave-of-projects.html> (accessed November 30,
[178] G. Brakmann, F.A. Mohammad, M. Dolejsi, M. Wiemann, Construction of the 2016).
ISCC Kuraymat, in: Int. SolarPACES Conf., Berlin, 2009. [199] Abengoa. Innovative technology solutions for sustainability, n.d.. <http://
[179] G.M. Elsaket, Simulating the Integrated Solar Combined Cycle for Power www.abengoa.com/htmlsites/boletines/en/septiembre2011/ingenieria/>
Plants Application in Libya, Cranfield University, 2007. <http://data.tour- (accessed November 30, 2016).
solaire.fr/thesis/2007_Elsaket_SIMULATING-THE-INTEGRATED-SOLAR- [200] M. Pehnt, Dynamic life cycle assessment (LCA) of renewable energy
COMBINED-CYCLE-FOR-POWER-PLANTS-APPLICATION-IN-LIBYA_116p.pdf>. technologies, Renew. Energy 31 (2006) 55–71, http://dx.doi.org/10.1016/j.
[180] Y. Li, Y. Yang, Thermodynamic analysis of a novel integrated solar combined renene.2005.03.002.
cycle, Appl. Energy 122 (2014) 133–142, http://dx.doi.org/10.1016/j. [201] N.A. Odeh, T.T. Cockerill, Life cycle GHG assessment of fossil fuel power plants
apenergy.2014.02.017. with carbon capture and storage, Energy Policy 36 (2008) 367–380, http://dx.
[181] G. Franchini, A. Perdichizzi, S. Ravelli, G. Barigozzi, A comparative study doi.org/10.1016/j.enpol.2007.09.026.
between parabolic trough and solar tower technologies in solar Rankine cycle [202] F. Ruzzenenti, M. Bravi, D. Tempesti, E. Salvatici, G. Manfrida, R. Basosi,
and integrated solar combined cycle plants, Sol. Energy 98 (2013) 302–314, Evaluation of the environmental sustainability of a micro CHP system fueled
http://dx.doi.org/10.1016/j.solener.2013.09.033. by low-temperature geothermal and solar energy, Energy Convers. Manage.
[182] M. Horn, H. Führing, J. Rheinländer, Economic analysis of integrated solar 78 (2014) 611–616, http://dx.doi.org/10.1016/j.enconman.2013.11.025.
combined cycle power plants, Energy 29 (2004) 935–945, http://dx.doi.org/ [203] S. Frick, M. Kaltschmitt, G. Schröder, Life cycle assessment of geothermal
10.1016/S0360-5442(03)00198-1. binary power plants using enhanced low-temperature reservoirs, Energy 35
[183] G. Manente, S. Rech, A. Lazzaretto, Optimum choice and placement of (2010) 2281–2294, http://dx.doi.org/10.1016/j.energy.2010.02.016.
concentrating solar power technologies in integrated solar combined cycle [204] P. Bayer, L. Rybach, P. Blum, R. Brauchler, Review on life cycle environmental
systems, Renew. Energy 96 (2016) 172–189, http://dx.doi.org/10.1016/j. effects of geothermal power generation, Renew. Sustain. Energy Rev. 26
renene.2016.04.066. (2013) 446–463, http://dx.doi.org/10.1016/j.rser.2013.05.039.
[184] A. Rovira, M.J. Montes, F. Varela, M. Gil, Comparison of heat transfer fluid and [205] B.M. Rule, Z.J. Worth, C. a Boyle, Comparison of life cycle carbon dioxide
direct steam generation technologies for integrated solar combined cycles, emissions and embodied energy in four renewable electricity generation
Appl. Therm. Eng. 52 (2013) 264–274, http://dx.doi.org/10.1016/j. technologies in New Zealand, Environ. Sci. Technol. 43 (2009) 6406–6413,
applthermaleng.2012.12.008. http://dx.doi.org/10.1021/es900125e.
[185] M.J. Montes, A. Rovira, M. Muñoz, J.M. Martínez-Val, Performance analysis of [206] D.Y.C. Leung, Y. Yang, Wind energy development and its environmental
an integrated solar combined cycle using direct steam generation in impact: a review, Renew. Sustain. Energy Rev. 16 (2012) 1031–1039, http://
parabolic trough collectors, Appl. Energy 88 (2011) 3228–3238, http://dx. dx.doi.org/10.1016/j.rser.2011.09.024.
doi.org/10.1016/j.apenergy.2011.03.038. [207] J. Hernández-Moro, J.M. Martínez-Duart, CSP electricity cost evolution and
[186] V. Siva, Reddy., S.C. Kaushik, S.K. Tyagi, Exergetic analysis of solar grid parities based on the IEA roadmaps, Energy Policy 41 (2012) 184–192,
concentrator aided natural gas fired combined cycle power plant, Renew. http://dx.doi.org/10.1016/j.enpol.2011.10.032.
Energy 39 (2012) 114–125, http://dx.doi.org/10.1016/j.renene.2011.07.031. [208] C.R. Chamorro, M.E. Mondéjar, R. Ramos, J.J. Segovia, M.C. Martín, M.A.
[187] H. Nezammahalleh, F. Farhadi, M. Tanhaemami, Conceptual design and Villamañán, World geothermal power production status: Energy,
techno-economic assessment of integrated solar combined cycle system with environmental and economic study of high enthalpy technologies, Energy
DSG technology, Sol. Energy 84 (2010) 1696–1705, http://dx.doi.org/ 42 (2012) 10–18, http://dx.doi.org/10.1016/j.energy.2011.06.005.
10.1016/j.solener.2010.05.007. [209] L. Feng, H. Chen, Y. Zhou, S. Zhang, T. Yang, L. An, The development of a
[188] C.S. Turchi, Z. Ma, Co-located gas turbine/solar thermal hybrid designs for thermo-economic evaluation method for solar aided power generation,
power production, Renew. Energy 64 (2014) 172–179, http://dx.doi.org/ Energy Convers. Manage. 116 (2016) 112–119, http://dx.doi.org/10.1016/j.
10.1016/j.renene.2013.11.005. enconman.2016.01.072.
[189] S. Gunasekaran, N.D. Mancini, R. El-Khaja, E.J. Sheu, A. Mitsos, Solar-thermal [210] F. Khaldi, Energy and exergy analysis of the first hybrid solar–gas power plant
hybridization of advanced zero emissions power cycle, Energy 65 (2014) in Algeria, in: Proc. 25th Int. Conf. Effic. Cost, Optim. Simul. Environ. Impact
152–165, http://dx.doi.org/10.1016/j.energy.2013.12.021. Energy Syst., Perugia, Italy, n.d.
[190] B. Kelly, U. Hermann, M. Hale, Optimization studies for integrated solar
combined cycle systems, in: Proc. Sol. Energy Forum Power to Choose,

You might also like