You are on page 1of 14

Applied Thermal Engineering 186 (2021) 116499

Contents lists available at ScienceDirect

Applied Thermal Engineering


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

Analysis of a 10 MW recompression supercritical carbon dioxide cycle for


tropical climatic conditions
Sharath Sathish a, Pramod Kumar a, *, Abdul Nassar b
a
Thermal Systems Laboratory, Department of Mechanical Engineering, Indian Institute of Science, Bangalore 560 012, India
b
Softinway Turbomachinery Solutions Pvt. Ltd., Bangalore 560 052, India

A R T I C L E I N F O A B S T R A C T

Keywords: Recompression supercritical CO2 (sCO2) power cycles offer higher efficiencies over other sCO2 cycle configu­
sCO2 power cycle rations. This paper addresses the design of a 10 MW sCO2 recompression cycle suitable for tropical climates with
Recompression a compressor inlet temperature of 45 ◦ C and a maximum turbine inlet temperature of 565 ◦ C to facilitate the use
Thermal efficiency
of conventional steel alloys. A design space analysis is carried out by incorporating individual component per­
Part load
Off-design
formance under various operating conditions. The proposed cycle uses a single two-stage axial turbine and two
single-stage centrifugal compressors. The results are presented on an efficiency-net-specific work diagram
showing the Pareto optimal curves for a range of recompression fractions. The optimum cycle parameters,
pressure ratio, and recompression fraction are found to be 2.0 and 0.25, respectively, yielding a design point
efficiency of 35%. The paper brings forth the effect of turbomachinery and heat exchangers sized for 45 ◦ C
compressor inlet temperature and 565 ◦ C turbine inlet temperature on off-design cycle performance. The analysis
is applied to develop strategies for optimally operating the recompression cycle at part load and off-design
conditions.

a regenerated Brayton cycle utilizing a pump for pressurizing the sub-


1. Introduction cooled CO2. The improved thermal efficiency, high power density,
avoidance of blade erosion, absence of pump cavitation, and single-
Supercritical carbon dioxide (sCO2) power cycles have transitioned phase heat rejection were some of the advantages highlighted by
from the kilowatt level pilot facilities [1,2] to the megawatt level Feher [11], over the recuperated air Brayton cycle and the Rankine
demonstration plants [3–5]. Theoretical studies and commercial bene­ cycle. Subsequently, Angelino [10], describes different configurations of
fits associated with sCO2 such as high efficiency, compact turboma­ condensing sCO2 cycles. The condensing cycles considered in the anal­
chinery, fuel agnostic heat source, zero or minimum water usage, ysis are only feasible where cooling water temperatures are below 15 ◦ C.
reduced footprint, cost, etc. have been comprehensively documented in The critical outcome of Angelino’s [10], investigation suggests that for
[3,6]. Several countries have dedicated programs on sCO2 development the turbine inlet temperatures in the range of 450 ◦ C to 550 ◦ C, the sCO2
and commercialization. Notable among them are the US Department of condensing cycles are not efficient as the steam cycle. They were,
Energy (DOE) [7], and European sCO2 Flex [8], programs. The DOE however, affected by conservative assumptions on turbomachinery and
emphasis is on high temperature, high efficiency solar, and fossil ap­ heat exchanger performance. Santini et al. [12], in their case study on
plications. The European program is focused on flexible fossil power. sCO2 cycles for a nuclear power plant, state that the opinion of sCO2
Currently, with renewable energy gaining traction, the baseload steam cycles not being suitable for low-temperature applications is unjustified.
and gas turbines are forced to operate as peak load plants with the need They show that for a maximum source temperature of 299 ◦ C and a heat
for increased operational flexibility. Therefore, the industrial power sink with a minimum temperature of 19 ◦ C, a recompressed reheated
production segment (lesser than 150 MW) is slowly gaining increasing regenerative CO2 cycle would yield a net efficiency of 34%, which is
attention worldwide, emphasizing the utilization of sCO2 technology. better than the current Rankine cycle.
The earliest literature on the sCO2 power cycle dates to Angelino A significant number of the sCO2 cycle configurations have focused
[9,10] and Feher [11],. Feher [11], in his investigations using closed- on low ambient conditions (32 ◦ C compressor inlet) [2,6]. Crespi et al.
cycle engines, describes a simple recuperated sCO2 cycle operating on [13], review the published results of sCO2 cycles covering 42 standalone

* Corresponding author.
E-mail address: pramod@iisc.ac.in (P. Kumar).

https://doi.org/10.1016/j.applthermaleng.2020.116499
Received 2 August 2020; Received in revised form 13 December 2020; Accepted 20 December 2020
Available online 26 December 2020
1359-4311/© 2020 Published by Elsevier Ltd.
S. Sathish et al. Applied Thermal Engineering 186 (2021) 116499

Nomenclature Subscripts
1–8 thermodynamic state points
c specific heat(kJ/kg.K) a approach end
Ċ capacitance rate (kW/K) c cold
CIP main compressor inlet pressure(bar) d discharge end
CIT main compressor inlet temperature (◦ C) h hot
h specific enthalpy(kJ/kg) in inlet
HT high temperature min minimum
LT low temperature max maximum
LMTD log mean temperature difference (◦ K) out outlet
ṁ mass flow rate(kg/s) mc main compressor
N speed(rpm) r ratio
NTU number of transfer units rc recompressor
P pressure(bar) s isentropic
PR pressure ratio t turbine
Q̇ heat transfer rate(kW) th thermal
s specific entropy(kJ/kg.K)
Greek symbols
T Temperature (◦ C)
Δ delta
TIT turbine inlet temperature (◦ C)
∊ effectiveness
UA conductance(kW/K)
( ) η efficiency
V volume m3 ∅ recompression fraction
wsp specific work(kJ/kg) ρ density
ẇ power(kW)

layouts and 38 combined cycle configurations. They state that the recuperating the high-temperature exhaust heat from the turbine, which
average thermal efficiency of standalone cycles is in the order of 40%. is vital to the thermal efficiency enhancement of the sCO2 cycle. The
The combined cycle configuration efficiencies range between 50% and effect of operating conditions is not trivial in the sCO2 plant design [6],.
60%. In the coal-fired sCO2 power plant study by Mecheri et al. [14], the Dyreby et al. [17], emphasize evaluating cycle performance as a func­
recompression cycle is considered the better alternative among other tion of total available recuperator conductance. The recompression
cycle configurations. Turchi et al. [15], describe the application of sCO2 cycle analysis by Dyreby [18], utilizes turbomachinery performance
in solar power towers. They state that the molten salt power tower maps from Sandia National Laboratory [2,19].
Concentrated Solar Power (CSP) has an upper-temperature limit of The seasonal variation in the heat sink temperature and the conse­
565 ◦ C. Cycles proposed for higher ambient conditions (more than 40 ◦ C quent effect on the off-design behavior of the sCO2 recompression cycle
at compressor inlet) have a high turbine inlet temperature greater than is studied in detail by Floyd et al. [20],. The analysis noted that without
600 ◦ C [3],. A more recent review of sCO2 cycles by Ahn et al. [16], any active control except for maintaining fixed turbomachinery speeds,
compares thermal efficiencies of steam Rankine, air Brayton and com­ the main compressor inlet density remains relatively constant even
bined cycle power plants with direct and in-direct fired sCO2 cycles at though the compressor inlet temperature changes because inlet pressure
different turbine inlet temperatures. It was found that above 600 ◦ C, in- changes in a closed system of fixed mass and volume. They note a sig­
direct fired sCO2 cycles yield higher efficiencies compared to both the nificant fall in the cycle efficiency but suggest independent control of the
Rankine & Brayton cycles. The paper highlighted the importance of compressor speed to overcome it. Duniam et al. [21], investigate the off-

Fig. 1. Schematic of the recompression cycle.

2
S. Sathish et al. Applied Thermal Engineering 186 (2021) 116499

600 5
plant of Marion et al. [3],.
P at 1,8,7,6' = 76.5 bar 1050 kJ/kg
500 3. Analytical models for the cycle and its components
P at 2',3',4,5 = 153 bar 950 kJ/kg
4
6'
400
850 kJ/kg
3.1. Assumptions and operating envelope limits
300

b ar
T, °C

The assumptions and scope that constitute the subsequent analysis

70
200 3' 700 kJ/kg are stated below.
2' 7
100
550 kJ/kg
i. The low side pressure is fixed at 76.5 bar which is enabled by an
bar 8
200 1 inventory control system. The inventory control system for part
0 load and off-design operation is detailed in Section 7.
0.2 0.4 0.6 0.8 450 kJ/kg
ii. The cooler is considered to achieve 45 ◦ C at the compressor inlet.
0.5 1.0 1.5 2.0 2.5 3.0 3.5 One percent pressure drop is assumed for the cooler and primary
s, kJ/kg-K heat exchanger in line with [18],.
iii. A pinch temperature of 5 ◦ C between the recuperator hot and cold
Fig. 2. Temperature-entropy diagram of the recompression cycle at the sides is considered for facilitating heat transfer.
design point. iv. As a first guess for design space exploration, the turbine isen­
tropic efficiency is assumed to be 90%, and isentropic efficiencies
design performance of an air-cooled sCO2 recompression cycle for CSP of both the compressors is assumed to be 85%. Subsequently, the
plants. The study concludes that the nominal power decreases by 10% isentropic efficiencies are corrected using the detailed design of
for a 10 ◦ C increase in the ambient temperature. turbomachinery.
The present paper addresses the design of a sCO2 power plant v. The electrical power trains driving the compressors are assumed
working within the practical constraints of the operating environment to have an efficiency of 95%. Turbine shaft power is considered as
and capital expenditure. The plant cost-effectiveness is ensured by uti­ the system power output.
lizing existing technology and material used in an industrial steam
turbine. A 10 MW, recompression sCO2 cycle, operating with the pri­ 3.2. Governing equations
mary compressor inlet temperature of 45 ◦ C and turbine inlet temper­
ature of 565 ◦ C, is analyzed. The maximum cycle pressure is limited to The computation of the steady-state cycle state points is carried out
200 bar. The cycle variables viz, pressure ratio, and recompression using Matlab® 2019b [22], coupled to Refprop 9.1 [23], for property
fraction are optimized for the operating envelope. A suitable turboma­ evaluations. The turbine and compressor designs are performed using a
chinery configuration is selected considering the cycle operating turbomachinery design tool, AxSTREAM® R2020 [24],. The design
bounds. The design of the cycle and components is applied to the point efficiencies and off-design performance map from the turboma­
optimal part load and off-design operation. The analysis provides vital chinery design tool are incorporated into the Matlab® program. The
insights into the overall operational envelope of a recompression sCO2 governing equations used in the analysis are listed below.
cycle. Though earlier studies [6,18,20,21] have concisely addressed the
effect of ambient temperature variation on recompression cycle off- 3.2.1. Main compressor
design performance, this work distinguishes because of specific and The main compressor exit enthalpy is evaluated using the below
detailed consideration of turbomachinery, heat exchanger design for relations.
45 ◦ C cycle inlet condition.
h2’ = h1 + (h2s − h1 )/ηmc (2)
2. Description of the recompression sCO2 Brayton cycle where enthalpies h1 and h2s are as per Eqs. (3) and (4)

The recompression sCO2 cycle (Fig. 1) consists of main compressor, h1 , s1 = f (P1 , T1 ) (3)
recompressor, turbine, high temperature recuperator, low temperature
recuperator, gas cooler, and a primary heat exchanger. The temperature- h2s = f (P2 , s1 ) (4)
entropy (T-s) diagram in Fig. 2 highlights the accompanying thermo­ The assumption (iv) provides efficiency for the initial cycle design
dynamic cycle. The main compressor (1-2′ ) and the recompressor (8-3′ ) point analysis. Final design and off-design evaluations utilize the main
compress low pressure CO2, which is above its critical point to the compressor performance maps generated from the design tool. The ef­
desired high side pressure by the operating pressure ratio. However, the ficiency and mass flow rate for a given pressure ratio, recompression
main compressor and recompressor operate at different inlet tempera­ fraction and operating speed are interpolated from the subsequent sec­
tures and mass flow rates. The recompression fraction (ϕ) is the ratio of tion’s performance maps.
the mass flow rate of the recompressor to the total mass flow rate (Eq.
(1)). 3.2.2. Recompressor
ϕ = ṁrc /ṁtotal (1) The recompressor exit enthalpy is evaluated using the following
relation.
As shown in Fig. 2, sensible heat from the turbine exhaust is recov­
ered in the high temperature recuperator (6′ -7) and the low temperature h3’ = h8 + (h3s − h8 )/ηrc (5)
recuperator (7–8). The recuperated heat increases the temperature of The enthalpies h8 and h3s are as per Eqs. (6) and (7).
CO2 from 2′ to 4 before entering the primary heat exchanger. External
heat addition occurs in the primary heat exchanger (4–5) before h8 , s8 = f (P8 , T8 ) (6)
expansion in the turbine (5-6′ ). A gas cooler cools the CO2 before it
enters the main compressor (8–1), prior to which flow splits to the h3s = f (P3’ , s8 ) (7)
recompressor. The recompression cycle configuration uses independent Based on the assumption (ii) we have,
electric motor driven compressors to offer the best operational flexi­
bility. This configuration has also been proposed in the 10 MW pilot P8 = P1 /(1 − 0.01) (8)

3
S. Sathish et al. Applied Thermal Engineering 186 (2021) 116499

P3′ = P4 / (1 − ΔPLT c ) (9) ΔTd = (T3’ − T7 ) (20)


The temperature T8 is evaluated from the LT recuperator model The LT recuperator conductance is calculated using the Eq. (21)
described subsequently in Section 3.2.4. Like the main compressor, below,
assumption (iv) provides the initial cycle design point efficiency. The
UALT = Q̇LT /LMTDLT (21)
recompressor performance maps provide the efficiencies and mass flow
rates for the final design and off-design conditions.
3.2.4.2. LT recuperator off-design performance. The effectiveness of the
3.2.3. Turbine counterflow heat exchanger is found using the following relations in Eqs.
The turbine exit enthalpy relation is given below. (22) and (23),

h6’ = h5 − (h5 − h6s )*ηt (10) εLT =


1 − e− NTULT (1− Cr )
ForCr < 1 (22)
1 − Cr e− NTULT (1− Cr )
The enthalpies h5 and h6s are obtained using the following equations.
NTU LT
h5 , s5 = f (P5 , T5 ) (11) εLT = ForCr = 1 (23)
1 + NTU LT
h6s = f (P6 , s5 ) (12) The number of transfer units (NTU) is defined in Eq. (24). The
The pressures are given by Eqs. (13) and (14). capacitance ratio (Cr) which is the ratio of the minimum and maximum
heat capacitance rates is given by Eq. (25).
P5 = P6 *PR (13)
UALT
NTU LT = (24)
P6 = P7 /(1 − ΔPHT h ) (14) Ċmin

The turbine inlet temperature T5 is fixed based on the cycle re­ Ċmin
quirements. Assumption (iv) provides efficiency for the initial cycle Cr = (25)
Ċmax
design point analysis. The turbine performance maps provide the effi­
ciencies for the final design and off-design conditions. The turbine mass The capacitance rate is stream mass flow rate times the average of the
flow rate is determined from the cycle based on the desired net power. stream specific heat capacity. The capacitance rates of the hot and cold
streams are defined below.
3.2.4. Low temperature (LT) recuperator
Ċh = ṁh ch (26)
The LT recuperator is modeled as a counterflow heat exchanger. The
hot and cold stream inlet exit temperatures are known for the design
Ċc = ṁc cc (27)
condition. The log mean temperature difference (LMTD) method is used
to determine the heat exchanger conductance (UA). The ε-NTU method The minimum heat capacity rate is the minimum of hot and cold
is used in the evaluation of off-design conditions of the heat exchanger. stream capacitance rate, and the maximum heat capacity rate is the
It leads to the solution of unknown temperatures (T8 and T4) using the maximum of the two. The maximum heat transfer rate is defined using
design conductance value. Eq. (28), and the actual heat transfer rate because of the heat exchanger
effectiveness is determined using Eq (29).
3.2.4.1. Design condition. The temperatures at 2′ and 3′ are known from
Q̇max = Ċmin (Th,in − Tc,in ) (28)
the main and recompressor models, respectively. The temperature at 7 is
obtained from the high temperature (HT) recuperator model output. The
Q̇actual = εLT Q̇max (29)
hot stream exit temperature T8 is estimated iteratively, as described in
Flowchart A.1 in Appendix A. The detailed temperature profiles and The hot and cold stream temperature differences are obtained using
pressure drops are evaluated as per the method described in [25],. The Eqs. (30) and (31), respectively. Subsequently, unknown temperatures
design yields a recuperator length of 0.5 m with 1.5 mm channel are evaluated using Eqs. (32) and (33).
diameter. The pressure drop across the hot stream is 0.11 bar and 0.02
bar across the cold stream. Appendix B summarizes the pinch analysis ΔTh = Q̇actual /Ċh (30)
for the recuperator.
ΔTc = Q̇actual /Ċc (31)
h8 , s8 = f (P8 , T8 ) (15)
T8 = T7 − ΔTh (32)
Again, using assumption (ii) we obtain,
P8 = P1 /(1 − 0.01) (16) T3’ = T2’ + ΔTc (33)

The heat transfer in the LT recuperator is provided below, The pressure drop at the off-design condition is evaluated as in [18],.

Q̇LT = ṁtotal (h7 − h8 )= (1− ϕ)ṁtotal (h3’ − h2’ ) (17) 3.2.5. High temperature (HT) recuperator
The log mean temperature difference across the LT recuperator is Like the LT recuperator, the HT recuperator is modeled as a coun­
found using the expression, terflow heat exchanger.

ΔTa − ΔTd 3.2.5.1. Design condition. The temperatures at 6′ and 3′ are known from
LMTDLT = ( ) (18)
ΔTa
ln ΔT the turbine and recompressor models, respectively. The hot stream
discharge temperature T7 follows from assumption (iii).
d

The following equations provide the approach and discharge end T7 = T3 + Pinch (34)
temperature difference of the LT recuperator.
The hot stream exit enthalpy is obtained using the following
ΔTa = (T2’ − T8 ) (19) equations.

4
S. Sathish et al. Applied Thermal Engineering 186 (2021) 116499

P7 = P8 /(1 − ΔPLT h ) (35)

h7 , s7 = f (P7 , T7 ) (36)
The evaluation of the state points at 4 is obtained as follows.
h4 = h3 + h6 − h7 (37)

T4 = f (P4 , h4 ) (38)

P4 = P5 /(1 − ΔPHT c ) (39)


The heat transfer in the HT recuperator is,

Q̇HT = ṁtotal (h6’ − h7 ) = ṁtotal (h4 − h3’ ) (40)


The log mean temperature difference (LMTD) across the HT recu­
perator is given by,
Fig. 3. Cycle efficiency variations with PR & ϕ for TIT 565 ◦ C, CIT 45 ◦ C and
ΔTa − ΔTd CIP 76.5 bar.
LMTDHT = ( ) (41)
ΔTa
ln ΔT d

The Eqs. (42) and (43) provide the HT recuperator approach and
discharge end temperature difference.
ΔTa = (T6’ − T4 ) (42)

ΔTd = (T7 − T3’ ) (43)


The HT recuperator conductance is calculated using,
PR
UAHT = Q̇HT /LMTDHT (44)
The recuperator length achieved is 2 m with a 1.5 mm channel
diameter. The pressure drop across the hot stream is 0.75 bar and 0.35
bar across the cold stream. Appendix B summarizes the pinch analysis
for the recuperator.

Fig. 4. Cycle efficiency variations with net specific work.


3.2.5.2. Off-design condition. For brevity, the ε-NTU relations for off-
design performance are skipped since they are identical to that
described earlier in Section 3.2.4.2. The unknown temperatures are
evaluated using the following equations. This section analyzes the above-stated sensitivity and finds the optimal
T7 = T6’ − ΔTh (45) operating parameters for the selected boundary conditions of 565 ◦ C
turbine inlet and 45 ◦ C compressor inlet temperatures. The design point
T4 = T3’ + ΔTc (46) studies at other conditions are elaborated in [26],.
The efficiency contour is plotted in Fig. 3 as a function of recom­
pression fraction and pressure ratio. It is to be noted that the turboma­
3.3. Cycle computation
chinery efficiency is based on the assumption (iv) for the initial analysis.
The maximum efficiency island is observed for a recompression fraction
The governing equations described above are iteratively solved, as
range 0.2 to 0.3 and pressure ratio range 1.7 to 2.5 with the center
illustrated in Flowchart A.1. The Matlab® program’s prescribed inputs
approximately at a recompression fraction of 0.25 and pressure ratio 2.0
are total mass flow rate, main compressor inlet temperature, main
(point A in Fig. 3). To better capture, the multi-objective Pareto front of
compressor inlet pressure, turbine inlet temperature, pressure ratio, and
the recompression cycle, efficiency-net specific work chart is plotted for
recompression fraction. The parameters of the cycle viz., state points,
the same variation in the pressure ratio and recompression fraction. The
component efficiencies, net thermal efficiency (47), specific work (48),
work reported in [27], on the simple recuperated cycle highlighted the
power (49), etc., are outputs from the program.
significance of such a representation to depict the cycle performance.
( ) ( )
h5 − h’6 − (1 − ϕ) h’2 − h1 − ϕ(h’3 − h8 ) This work further extends the paradigm by increasing variable dimen­
ηth = (47)
(h5 − h4 ) sion to account for the recompression fraction.
Fig. 4 shows the variation in the cycle efficiency and net specific
( ) ( )
wsp = h5 − h’6 − (1 − ϕ) h’2 − h1 − ϕ(h’3 − h8 ) (48) work for different recompression fractions from 0.05 to 0.55. The
pressure ratio increases from left to right on each curve. Following the
ẇnet = ṁtotal wsp (49) efficiency contour in Fig. 3, the maximum cycle efficiency is observed
for the recompression fraction varying between 0.2 and 0.3. The critical
4. Cycle design space exploration insight from Fig. 4 is that the net specific work is not maximized at the
maximum cycle efficiency condition but occurs at a higher pressure
Design space exploration assesses the cycle’s key performance pa­ ratio. Fig. 5 shows the locus of the maximum cycle efficiency and net
rameters’ sensitivity to the system’s primary design variables. For the specific work.
recompression cycle, the pressure ratio and recompression fraction are Fig. 5a shows that the net specific work penalty at the maximum
the primary design variables. The non-dimensional parameters: cycle cycle efficiency point is 40 kJ/kg. The cycle efficiency penalty at
efficiency, and net specific work are the key performance parameters. maximum net specific work is 2% points. This result may indicate that

5
S. Sathish et al. Applied Thermal Engineering 186 (2021) 116499

Table 2
Turbine flow-path geometry parameters.
Hub Diameter (mm) Tip Diameter (mm) No of blades

Stage 1 Nozzle 141.8 233.3 16

Blade 141.8 233.3 17


Stage 2 Nozzle 141.8 260.6 18
Blade 141.8 260.6 19

5. Turbomachinery design and performance

5.1. Boundary conditions for turbomachinery design


a
The compressors and the turbine are designed using the design tool’s
flow path modules and boundary conditions as per the preceding sec­
tion’s specifications. The approach used here is based on meanline/
streamline approach which is described in [28–30],[31]. The boundary
conditions for the design point are presented in Table 1.

5.2. Selection of optimal operating speed

The turbine and compressors’ geometric parameters such as di­


ameters, blade height, speed, and design parameters are varied to assess
turbomachinery performance sensitivity. The turbine and compressors
can be operated in either coupled or de-coupled manner. Hence, it is
essential to select the speeds that provide the best performance. The
shaft speeds corresponding to the maximum efficiencies of the main
b compressor, recompressor, and turbine are 25,018 rpm, 39,090 rpm,
and 18,120 rpm, respectively. A uniform speed of 21,000 rpm is selected
Fig. 5. Locus of maximum efficiency & net specific work on a) η-wsp chart b)
for both the compressors and the turbine for coupled operation. The
PR-ϕ chart.
speed can, however, be varied to achieve maximum overall cycle effi­
ciency for the de-coupled operation.

choosing the maximum net specific work condition would be ideal as a 5.3. Turbine performance
design point. However, a closer look at the operating conditions is
essential. In Fig. 5b, the same loci are plotted against the pressure ratio The turbines could be designed as either Impulse or Reaction type.
and recompression fraction. From the operating limits of the turbine, Since maximum turbine blade heights are small, a drum type rotor with
compressors, and other equipment, the feasible operating regime is hub reaction varying from 20% to 40% is selected. The profile loss is
highlighted as a rectangular region within Fig. 5b. Only the locus of the estimated using the modified Craig and Cox [32], loss model. The Ste­
maximum efficiency passes through this identified operating zone. This panov [33], loss model is used for secondary loss estimation. The geo­
provides a practical basis for the selection of the optimum design point. metric parameters of the turbine are presented in the Table 2. The
Accordingly, a recompression fraction of 0.25 and a pressure ratio of 2.0 overall axial length of the turbine is 215 mm.
are selected as the design point. These parameters are used in the turbine As the hub diameter to blade length ratio is significantly small, the
and compressors’ detailed design, as described in the next section. The nozzle and rotor blades for both the stages are designed using a
design point net power of the recompression cycle is 10 MW. controlled vortex method. At the design point, the turbine produces net
power of 18.3 MW at a corresponding shaft speed of 21,000 rpm and a
total-to-static pressure ratio of 2.0. The total-to-static efficiency for the
turbine is estimated to be 86.7%.
Table 1
Turbomachinery design boundary conditions.
5.4. Compressor performance
Sl. No Parameter Unit Value

Main Compressor The flow rates for both the main compressor and the recompressor
1. Inlet pressure bar 76.5 are comparatively low, which do not justify the usage of multi-stage
2. Outlet pressure bar 153.0 axial compressors. Therefore, a single-stage centrifugal compressor
3. Inlet temperature ◦
C 45.0
4. Mass flow rate kg/s 150.0
configuration is selected for both the main compressor and the recom­
Recompression compressor pressor. The centrifugal compressor includes an impeller, vane diffuser,
5. Inlet pressure bar 76.5 and volute. Though the pressure ratio is nearly identical for both the
6. Outlet pressure bar 153.0 compressors, the mass flow rate and inlet temperatures differ as the
7. Inlet temperature C 106.0
recompression cycle uses varying mass flow split. The profile loss model

8. Mass flow rate kg/s 50.0


Turbine is based on Aungier [34], and the deviation angle calculation is based on
9. Inlet pressure bar 153.0 Wiesner [35],.
10. Outlet pressure bar 76.5
11. Inlet temperature ◦
C 565.0 5.4.1. Main compressor
12. Mass flow rate kg/s 200.0
The impeller inlet hub and tip diameters obtained from the design

6
S. Sathish et al. Applied Thermal Engineering 186 (2021) 116499

Table 3
Cycle parameters for basic and high performance cycles.
Cycle Parameter Unit Basic High Performance
Cycle Cycle

Turbine inlet temperature ◦


C 550 700
Compressor inlet temperature ◦
C 32
Compressor outlet pressure bar 200
Pressure ratio – 2.6
Compressor isentropic efficiency – 0.89
Recompressor isentropic – 0.89
efficiency
Turbine isentropic efficiency – 0.90

tool are 40 mm and 129.54 mm. The impeller outlet tip diameter is 208
mm. The exit blade width is 14.1 mm. The vane diffuser outlet diameter
Fig. 7. Part load cycle and turbomachinery efficiency.
is 311.7 mm. The designed compressor has a total-to-total efficiency of
85.8% at a pressure ratio of 2.0. The main compressor consumes a power
of 4.7 MW.
7. Design performance analysis and application to part load and
5.4.2. Recompressor off-design operation
The impeller inlet hub and tip diameters are 44.5 mm and 120.1 mm,
respectively. The impeller outlet tip diameter is 273 mm. The exit blade The design point is the 100% load point having design boundary
width is 7.73 mm, and vane diffuser outlet diameter is 409.1 mm. The conditions for the cycle and the components. Part load refers to the
designed compressor has a total-to-total efficiency of 76.9% at a pres­ variation in the net power to meet the flexible demand from the grid. In
sure ratio of 2.0 with a power consumption of 3.3 MW. The decrease in this case, a variation of net power from 40% to 110% is considered. The
the efficiency of the recompressor compared to the main compressor is part load analysis is carried out at the design operating conditions of the
attributable to two factors; a) lower mass flow rate resulting in shorter cycle. In the off-design analysis, two main variations in boundary con­
exit blade width, which leads to higher loss, b) higher inlet temperature ditions are assessed: i) variation in the main compressor inlet tempera­
(106 ◦ C) compared to the main compressor (45 ◦ C). The turbine and ture, ii) variation in the turbine inlet temperature. All the other
compressor performance maps are elucidated in Appendix C. boundary conditions are fixed as per the initial design, including the
cycle power output of 10 MW. In both the part load and off-design
6. Model validation analysis, the two compressors are independently driven by individual
electrical motors, while the turbine drives a generator. Both compressors
Two design conditions are considered for validation of the cycle are operated at independent speeds for optimum efficiency resulting in
model developed in Section 3. They are the “Basic” and “High Perfor­ an uncoupled recompression cycle configuration. The coupled configu­
mance” cycles described by Dostal [6],. The turbine inlet temperatures ration where the turbine drives both the generator and the main
are 550 ◦ C and 700 ◦ C for the Basic and High Performance cycles. The compressor is not viable. It violates the operating range of either the
same conditions have also been considered by Dyreby [18],. The cycle main compressor or the recompressor. At each part load/off-design
parameters for the two design conditions are shown in Table 3. point, both the compressor speeds are independently varied. The
The Basic and High Performance cycle thermal efficiencies are compressor mass flow rates and the pressure ratios are set by the turbine
shown in Fig. 6. The variation of efficiency with recompression fraction power and the cycle parameters. The optimum cycle efficiency at part
is shown for the current model and compared with other models’ opti­ load/off-design condition is arrived at by varying the compressors’
mum points. This comparison establishes the match in the efficiency speed while conforming to the operational limits dictated by their in­
value and the optimum recompression fraction for the chosen design dividual performance maps.
conditions, thus validating the current model. sCO2 Brayton cycles are non-condensing closed loop cycles with the
Validation of the turbomachinery models and its application to sCO2 pressure and temperature at the compressor inlet being independent
cycles are highlighted in [28,36]. parameters. Therefore, locking the compressor inlet pressure while the
temperature varies demands explicit control. The part load operation is
most effectively achieved by reducing the mass flow rate in the system.
Thus, inventory control plays a crucial role in enabling both part load
and off-design operation. Alternatively, Turbine by-pass control [37], is
an another strategy suggested in the literature. The inventory control
mechanism involves bleeding and feeding CO2 mass from the primary
loop to and from an external tank. Though there are different configu­
rations to implement the inventory system, the one shown in the sche­
matic Fig. D.1 is a simple and effective [6], concept. The inventory
changes due to part load and off-design operation discussed in this paper
are detailed in Appendix D.

7.1. Design point performance

The cycle design point results are updated with the turbine and
compressor efficiencies from Section 5. The design operating conditions
and the optimum recompression cycle parameters (pressure ratio 2.0
and recompression fraction of 0.25) result in a net cycle efficiency of
Fig. 6. Thermal efficiency for Basic & High Performance cycles. 35% (at 10 MW net power). The actual turbomachinery efficiencies are

7
S. Sathish et al. Applied Thermal Engineering 186 (2021) 116499

Fig. 8. Optimum ϕ and PR for part load operation. Fig. 11. Optimum ϕ and PR for varying CIT.

Fig. 9. Optimum compressor speed for part load operation. Fig. 12. Optimum compressor speeds for varying CIT.

lower than that assumed in the initial design point analysis. efficiency bandwidth is not as robust as that of the main compressor.
The following figures highlight the variation in the optimum cycle
7.2. Part load performance and component parameters across the part load operation. Fig. 8 shows
that the optimum recompression fraction varies between 0.35 and 0.25
The part load cycle and turbomachinery efficiencies are illustrated in for the part load variation from 40% to 110%. The design point
Fig. 7. The cycle efficiency varies between 28% at 40% part load to 35% recompression fraction of 0.28 is following the earlier stated result in the
at 100% part load condition. The turbine efficiency variation with part design space exploration. Similarly, the pressure ratio varies between
load is the primary contributor to the deteriorated cycle efficiency as the 1.5 and 2.2 for the part load variation between 40% and 110%. The
turbine is constrained to operate at a fixed speed to drive the generator. design point pressure ratio of 2.0 matches with that of the results from
The main compressor efficiency is relatively flat throughout the oper­ the design space exploration. The compressor speeds vary between 75%
ating range owing to the flexible speed operation. Likewise, the and 110% of the design speed, as shown in Fig. 9. The pressure ratio
recompressor speed varies across the operating range though its

Fig. 10. Off design cycle and compressor efficiency variation with CIT. Fig. 13. Off design cycle and turbomachinery efficiency variation with TIT.

8
S. Sathish et al. Applied Thermal Engineering 186 (2021) 116499

efficiency variation is 10%. The turbine efficiency is not shown as it


remains unaffected since the turbine inlet temperature is at its design
value.
The optimum recompression fraction and optimum pressure ratio are
shown in Fig. 11. The corresponding optimum compressor speeds are
shown in Fig. 12. The optimum recompression fraction variation is be­
tween 0.25 and 0.45. Like the part load, the optimal recompression
fraction is higher for lower main compressor inlet temperatures. The
optimal pressure ratio varies between 1.8 and 2.2. The corresponding
optimal compressor speed variation is 90% and 105%. Main compressor
and recompressor speed variations exhibit exactly opposite trends
attributable to a steep change in the recompression fraction. A high
recompression fraction at lower compressor inlet temperature leads to a
higher recompressor mass flow rate and higher speed.
Fig. 14. Optimum ϕ and PR for varying TIT.
7.3.2. Turbine inlet temperature variation
The cycle efficiency varies between 27% and 35% for the variation in
the turbine inlet temperature ranging from 450 ◦ C to 565 ◦ C, as shown in
decreases with decreasing part load following the swallowing capacity Fig. 13. The cycle efficiency variation is significant compared to the case
of the turbine. The recompression fraction and the compressor speed where the compressor inlet temperature was varied. Though the turbine
vary to achieve the optimum cycle efficiency respecting the operational efficiency is expected to drop with lower inlet temperature, it is nearly
constraints set by the turbine and compressor performance maps (Ap­ constant. This behaviour is due to the increase in pressure ratio with
pendix 3). decreasing turbine inlet temperature (TIT) to meet the required set
power output of 10 MW. Therefore, the cycle efficiency drop at lower
TIT is due to the non-optimal pressure ratio (1.7 at 450 ◦ C [26],)
7.3. Off-design performance
imposed by the performance map limits.
The variation in the optimal recompression fraction and optimal
Off-design performance due to the main compressor inlet tempera­
pressure ratio is shown in Fig. 14, while the corresponding optimal
ture variation and turbine inlet temperature variation is discussed in
compressor speeds are shown in Fig. 15. The optimal recompression
Sections 7.3.1 and 7.3.2, respectively. The variation in the compressor
fraction varies in a narrow range while the optimal pressure ratio varies
inlet temperature is due to the variation in the ambient temperature. The
between 2.0 and 2.5. It is observed that the pressure ratio increases as
range of main compressor inlet temperature variation considered is
the turbine inlet temperature decreases. The compressor speed varies
between 32 ◦ C and 55 ◦ C. The turbine inlet temperature may vary based
between 120% and 100%, requiring the compressors to be sped up as the
on the heat source temperature variation but is limited to 565 ◦ C ac­
turbine inlet temperature is lowered.
cording to the scope of this work concerning industrial applications and
its material and technology limits. The range of the turbine inlet tem­
8. Conclusions
perature considered for this analysis varies between 450 ◦ C and 565 ◦ C.

The paper provides a comprehensive analysis of a recompression


7.3.1. Main compressor inlet temperature variation
sCO2 cycle in a de-coupled configuration developing 10 MW of shaft
The cycle efficiency varies between 32% and 35%, with a peak
power. The cycle is designed for operating in tropical environments with
occurring at the design point main compressor inlet temperature of
a design main compressor inlet temperature of 45 ◦ C. The maximum
45 ◦ C (in Fig. 10). The efficiency variation is relatively limited when
cycle temperature is restricted to 565 ◦ C with a maximum cycle pressure
compared to the part load operation. One would intuitively expect that
of 200 bar to facilitate conventional steel alloys. The design point se­
the lower compressor inlet temperature would lead to higher cycle ef­
lection methodology adopted in this work establishes a clear, logical
ficiency, but it does not turn out to be so. This outcome is because the
structure that makes it lucidly explainable against automated optimi­
optimal pressure ratio at 32 ◦ C should have been 2.5 [26], against 1.9
zation routines. The resulting value of the optimum pressure ratio is 2.0,
imposed by the turbomachinery performance maps. Further, the main
and the optimum recompression fraction is around 0.25. The optimum
compressor efficiency variation is insignificant, while the recompressor
recompression fraction is different from 0.4 reported earlier [6], as they
considered low ambient design conditions. The proposed cycle employs
a two-stage axial turbine and single-stage compressors. The net cycle
efficiency achieved after incorporating the designed turbomachinery
efficiencies is 35%. An inventory control system is proposed to inject or
extract CO2 from the system as need be, when operating under part load
or off design conditions at varying ambient temperatures.
Optimal operational strategy is developed for part load (40% to
110%) and off-design operation (CIT 32 ◦ C to 55 ◦ C, TIT 450 ◦ C to
565 ◦ C) by applying the cycle and component models developed in this
work. The turbine’s constraint to run at fixed speed conditions to drive
the generator results in cycle efficiency drop at part loads. In contrast,
the variable speed operation resulted in the compressors’ robust per­
formance across the part load regime. Hence, the turbine becomes a
critical factor in deciding the efficiency of the cycle. The recompression
fraction and the pressure ratio vary across the part load range to achieve
the optimal value, which maximizes the cycle efficiency. This work’s key
outcome highlights the significance of individual component design
Fig. 15. Optimum compressor speeds for varying TIT. point considerations to the off-design cycle performance. Hence, the

9
S. Sathish et al. Applied Thermal Engineering 186 (2021) 116499

holistic design approach encompassing part load and off-design opera­ the work reported in this paper.
tional strategies developed for tropical climate conditions is unique. In
summary, the analysis and application presented in the paper shed light Acknowledgments
not only on the design of the recompression cycle but also on the
operational envelope constrained by the practical limits of component The authors acknowledge Department of Science and Technology,
operation. Government of India for financial support vide sanction order(s): TMD/
CERI/CSP/2020/1(G) dated 21-09-2020, and TMD/CERI/Clean Coal/
Declaration of Competing Interest 2017/034 (IISc) (G) dated 13-09-2018. The inputs from Triveni Tur­
bines Limited and Tata Consulting Engineers (TCE), Bangalore in car­
The authors declare that they have no known competing financial rying out this work is also gratefully acknowledged.
interests or personal relationships that could have appeared to influence

Appendix A

Flowchart for Cycle Thermodynamic Calculations


(See Fig. A1)

Fig. A1. Flowchart for the estimation of recompression cycle state points.

10
S. Sathish et al. Applied Thermal Engineering 186 (2021) 116499

Appendix B

LT and HT recuperator pinch analysis

The design point temperature profiles of HT and LT recuperators are shown in Figs. B.1 and B.2, respectively. The hot and cold fluid stream
temperatures vary across the heat exchanger’s length with a pinch (minimum temperature difference between the hot and cold streams) occurring at
the hot outlet end for both the recuperators.

Fig. B1. HT recuperator temperature profile.

Fig. B2. LT recuperator temperature profile.

Appendix C

Turbomachinery performance maps

C.1 Turbine performance map


The total-to-static efficiency and total-to-static pressure ratio as a function of the corrected mass flow rate at 100% speed is represented in Fig. C.1.
The efficiency penalty for the lowest corrected mass flow rate is 15% compared to the maximum efficiency point. This penalty is evident in the part
load operation analysis in Section 7. The corrected mass flow rate increases with pressure ratio.

C.2 Main compressor performance map


The total-to-total efficiency and total-to-total pressure ratio as a function of the corrected mass flow rate at different speeds (50% to 120% of the
design speed) is represented in the Figs. C.2 and C.3, respectively. At any given speed, there is an efficiency penalty with higher mass flow rates (in
Fig. C.2), and the operating pressure ratio also decreases with higher mass flow rates (in Fig. C.3).

C.3 Recompressor performance map


The total-to-total efficiency and total-to-total pressure ratio as a function of the corrected mass flow rate at different speeds (50% to 120% of the
design speed) is represented in the Figs. C.4 and C.5, respectively. Like the main compressor, at any given speed, there exists an efficiency penalty with
higher mass flow rates (C.4), and the operating pressure ratio also decreases with higher mass flow rates (C.5).

11
S. Sathish et al. Applied Thermal Engineering 186 (2021) 116499

Fig. C1. Turbine performance map for 100% speed.

Fig. C5. Recompressor performance map for pressure ratio.

Appendix D

Inventory system for part load and Off-Design operation

The schematic of the inventory system is shown in Fig. D.1. The in­
ventory tank is connected between the compressor discharge (2′ ) and
gas cooler inlet (8).
Part load operation with fixed low side pressure is achieved by
bleeding or feeding the CO2 loop mass to or from the inventory tank.
Pressure is directly proportional to density in ideal gas cycles which
eases the part load operation through inventory control [38],. But CO2
properties vary significantly, particularly in the vicinity of the critical
point where the main compressor operates. Hence part load thermody­
Fig. C2. Main Compressor performance map for efficiency. namic state points which are evaluated using Refprop [23], are used in
inventory calculations [39],. The inventory balance for the system is
shown in the equation below.
( )
Σmj=1 ρj Vj = M
total
(D.1)

In a system with m components the total inventory is a product of


individual component inventory. The individual component inventory is
a product of density ρ and volume V. While the density and hence, the
individual component inventory may vary with time, the total inventory
remains the same. The inventory system provisioning relative to the
design condition are shown in Fig. D.2 for part load operation under
quasi-static conditions. The decreasing pressure ratio shown in Fig. 8 at
lower power levels demands reduction of the inventory in the
compressor discharge-end and vice-versa for the gas cooler inlet-end.
Thus, under part load conditions with varying compressor speed, the
inventory needs to be essentially shifted from the compressor discharge-
end to the gas cooler inlet-end.
Fig. C3. Main Compressor performance map for pressure ratio. The inventory provisioning for off-design condition with variation in
the main compressor inlet temperature is shown in Fig. D.3. The demand
for inventory on the gas cooler inlet-end increases when the temperature
is lowered from the design condition of 45 ◦ C. The pressure drops as the
temperature is lowered and hence more inventory is provisioned to
maintain the fixed compressor inlet pressure. The inventory rises on the
compressor discharge-end too, since the compressor inlet temperature
drops significantly compared to the design condition. But a significant
demand from the inventory on the discharge-end is balanced by the
reduced high side pressure as the pressure ratio drops as revealed in
Fig. 11 at lower compressor inlet temperature.
The variation in the inventory for varying turbine inlet temperature
is shown in Fig. D.4. Since the compressor inlet pressure is held at design
conditions, the inventory variation on the gas cooler inlet-end is
considerably low. The increasing pressure ratio with reducing turbine
inlet temperature shown in Fig. 14, necessitates inventory addition on
the compressor discharge end.
Fig. C4. Recompressor performance map for efficiency.
12
S. Sathish et al. Applied Thermal Engineering 186 (2021) 116499

The inventory provisioning discussed in this work is in addition to the compressor speed variation. Together they alleviate the cycle efficiency
penalty in the part load and off-design operation of the cycle. The detailed inventory provisioning and control system design needs dynamic simulation
of the system which is beyond the scope of this work.

Fig. D1. Inventory system schematic for the sCO2 recompression cycle.

Fig. D2. Part load inventory change relative to the 100% load (10
MW) condition.
Fig. D4. Off-design inventory change for varying TIT, relative to the design
TIT (565 ◦ C).

References

[1] J.J. Pasch, T.M. Conboy, D.D. Fleming, G.E. Rochau, Supercritical CO2
recompression Brayton cycle- Completed assembly description, Albuquerque, NM,
and Livermore, CA (United States) (2012), https://doi.org/10.2172/1057248.
[2] T. Conboy, S. Wright, J. Pasch, D. Fleming, G. Rochau, R. Fuller, Performance
characteristics of an operating supercritical CO2 brayton cycle, in: Proc. ASME
Turbo Expo, American Society of Mechanical Engineers Digital Collection, 2012:
pp. 941–952. Doi: 10.1115/GT2012-68415.
[3] J. Marion, M. Kutin, A. McClung, J. Mortzheim, R. Ames, The step 10 MWe SCO2
pilot plant demonstration, in: Proc. ASME Turbo Expo, American Society of
Mechanical Engineers (ASME), 2019. Doi: 10.1115/GT2019-91917.
[4] R. Allam, S. Martin, B. Forrest, J. Fetvedt, X. Lu, D. Freed, G.W. Brown, T. Sasaki,
M. Itoh, J. Manning, Demonstration of the Allam Cycle: An Update on the
Development Status of a High Efficiency Supercritical Carbon Dioxide Power
Process Employing Full Carbon Capture, in, Energy Procedia, Elsevier Ltd (2017)
5948–5966, https://doi.org/10.1016/j.egypro.2017.03.1731.
[5] T.J. Held, Initial Test Results of a Megawatt-Class Supercritical CO2 Heat Engine,
Fig. D3. Off-design inventory change for varying CIT, relative to the design in: The 4th International Symposium-Supercritical CO2 Power Cycles, 2014.
CIT (45 ◦ C). https://www.echogen.com/_CE/pagecontent/Documents/Papers/initial-test-
results-of-a-megawatt-class-supercritical-co2-heat-engine-held.pdf (accessed July
19, 2020).

13
S. Sathish et al. Applied Thermal Engineering 186 (2021) 116499

[6] V. Dostal, A supercritical carbon dioxide cycle for next generation nuclear reactors, [23] E.W. Lemmon, M.L. Huber, M.O. McLinden, NIST Standard Reference Database 23:
Doctoral dissertation, Massachusetts Institute of Technology, 2004. https://dspace. Reference Fluid Thermodynamic and Transport Properties-REFPROP, Version 9.1,
mit.edu/handle/1721.1/17746 (accessed July 19, 2020). Natl, Inst. Stand. Technol. (2013), https://doi.org/10.18434/T4/1502528.
[7] SCO2 Power Cycles | Department of Energy, (n.d.). https://www.energy.gov/sco2- [24] Turbomachinery Solutions | Turbomachinery Design Software, (n.d.). https://
power-cycles (accessed July 19, 2020). www.softinway.com/software/ (accessed August 2, 2020).
[8] sCO2-flex - sCO2flex, (n.d.). https://www.sco2-flex.eu/ (accessed July 19, 2020). [25] V. Pandey, P. Kumar, P. Dutta, Thermo-hydraulic analysis of compact heat
[9] G. Angelino, Perspectives for the liquid phase compression gas turbine, J. Eng. Gas exchanger for a simple recuperated sCO2 Brayton cycle, Renew. Sustain. Energy
Turbines Power 89 (1967) 229–236, https://doi.org/10.1115/1.3616657. Rev. 134 (2020), 110091, https://doi.org/10.1016/j.rser.2020.110091.
[10] G. Angelino, Carbon dioxide condensation cycles for power production, J. Eng. Gas [26] S. Sathish, P. Kumar, A.N. Namburi, P.C. Gopi, M.D. Carlson, C.K. Ho, Optimization
Turbines Power 90 (1968) 287–295, https://doi.org/10.1115/1.3609190. of Operating Parameters of a Recompression sCO2 Cycle for Maximum Efficiency,
[11] E.G. Feher, The supercritical thermodynamic power cycle, Energy Convers. 8 in: Proc. ASME Turbo Expo, American Society of Mechanical Engineers (ASME),
(1968) 85–90, https://doi.org/10.1016/0013-7480(68)90105-8. 2017. Doi: 10.1115/GT2017-64625.
[12] L. Santini, C. Accornero, A. Cioncolini, On the adoption of carbon dioxide [27] P. Garg, P. Kumar, K. Srinivasan, A trade-off between maxima in efficiency and
thermodynamic cycles for nuclear power conversion: A case study applied to specific work output of super- and trans-critical CO2 Brayton cycles, J. Supercrit.
Mochovce 3 Nuclear Power Plant, Appl. Energy 181 (2016) 446–463, https://doi. Fluids. 98 (2015) 119–126, https://doi.org/10.1016/j.supflu.2014.12.023.
org/10.1016/j.apenergy.2016.08.046. [28] L. Moroz, B. Frolov, M. Burlaka, O. Guriev, Turbomachinery flowpath design and
[13] F. Crespi, G. Gavagnin, D. Sánchez, G.S. Martínez, Supercritical carbon dioxide performance analysis for supercritical CO2, in: Proc. ASME Turbo Expo, American
cycles for power generation: A review, Appl. Energy 195 (2017) 152–183, https:// Society of Mechanical Engineers (ASME), 2014. Doi: 10.1115/GT2014-25385.
doi.org/10.1016/j.apenergy.2017.02.048. [29] L. Moroz, Y. Govoruschenko, P. Pagur, Axial turbine stages design: 1D/2D/3D
[14] M. Mecheri, Y. Le Moullec, Supercritical CO2 Brayton cycles for coal-fired power simulation, experiment, optimization - Design of single stage test air turbine
plants, Energy. 103 (2016) 758–771, https://doi.org/10.1016/j. models and validation of 1D/2D/3D aerodynamic computation results against test
energy.2016.02.111. data, in: Proc. ASME Turbo Expo, American Society of Mechanical Engineers
[15] C.S. Turchi, Z. Ma, J. Dyreby, Supercritical carbon dioxide power cycle Digital Collection, 2005: pp. 1137–1146. Doi: 10.1115/GT2005-68614.
configurations for use in concentrating solar power systems, in: Proc. ASME Turbo [30] L. Moroz, Y. Govorushchenko, P. Pagur, K. Grebennik, W. Kutrieb, M. Kutrieb,
Expo, American Society of Mechanical Engineers Digital Collection, 2012: pp. Integrated Environment for Gas Turbine Preliminary Design, in: Proc. Int. Gas
967–973. Doi: 10.1115/GT2012-68932. Turbine Conf., IGTC2011-0007, 2011.
[16] Y. Ahn, S.J. Bae, M. Kim, S.K. Cho, S. Baik, J.I. Lee, J.E. Cha, Review of [31] L. Moroz, L. Moroz, Y. Govorushchenko, P. Pagur, S. Inc, A uniform approach to
supercritical CO2 power cycle technology and current status of research and conceptual design of axial turbine / compressor flow path”, The, Futur. Gas
development, Nucl. Eng. Technol. 47 (2015) 647–661, https://doi.org/10.1016/j. Turbine Technol.3 RD Int. Conf. 11-12 Oct. 2006. http://citeseerx.ist.psu.edu/
net.2015.06.009. viewdoc/summary?doi=10.1.1.510.1258 (accessed July 20, 2020).
[17] J. Dyreby, S. Klein, G. Nellis, D. Reindl, Design considerations for supercritical [32] H.R.M. Craig, H.J.A. Cox, Performance Estimation of Axial Flow Turbines, Proc.
carbon dioxide brayton cycles with recompression, J. Eng. Gas Turbines Power. Inst. Mech. Eng. 185 (1970) 407–424, https://doi.org/10.1243/PIME_PROC_1970_
136 (2014), https://doi.org/10.1115/1.4027936. 185_048_02.
[18] J.J. Dyreby, Modeling the Supercritical Carbon Dioxide Brayton Cycle with [33] Stepanov, Turbomachinery Cascades Hydrodynamics, Gos. Izd. Phys. Mat. Lit.
Recompression, Doctoral Dissertation, University of Wisconsin-Madison, 2014. (1962).
https://search.proquest.com/openview/8948e313dfadaa310efcc8de04ad8ce2/1? [34] R.H. Aungier, Centrifugal Compressors: A Strategy for Aerodynamic Design and
pq-origsite=gscholar&cbl=18750&diss=y (accessed July 19, 2020). Analysis, ASME Press (2019), https://doi.org/10.1115/1.800938.
[19] S.A. Wright, R.F. Radel, M.E. Vernon, P.S. Pickard, G.E. Rochau, Operation and [35] F.J. Wiesner, A review of slip factors for centrifugal impellers, J. Eng. Gas Turbines
analysis of a supercritical CO2 Brayton cycle., Albuquerque, NM, and Livermore, Power. 89 (1967) 558–566, https://doi.org/10.1115/1.3616734.
CA (United States), 2010. Doi: 10.2172/984129. [36] F. Crespi, Thermo-Economic Assessment of Supercritical CO2 Power Cycles for
[20] J. Floyd, N. Alpy, A. Moisseytsev, D. Haubensack, G. Rodriguez, J. Sienicki, Concentrated Solar Power Plants, Doctoral Dissertation, University of Seville,
G. Avakian, A numerical investigation of the sCO2 recompression cycle off-design 2019. https://www.scarabeusproject.eu/wp-content/uploads/2020/02/PhD-
behaviour, coupled to a sodium cooled fast reactor, for seasonal variation in the Crespi-Final-1.pdf (accessed October 10, 2020).
heat sink temperature, Nucl. Eng. Des. 260 (2013) 78–92, https://doi.org/ [37] A. Moisseytsev, K.P. Kulesza, J.J. Sienicki, Control system options and strategies for
10.1016/j.nucengdes.2013.03.024. supercritical CO2 cycles, Argonne, IL (2009), https://doi.org/10.2172/958037.
[21] S. Duniam, A. Veeraragavan, Off-design performance of the supercritical carbon [38] N.A. Carstens, Control Strategies for Supercritical Carbon Dioxide Power
dioxide recompression Brayton cycle with NDDCT cooling for concentrating solar Conversion Systems, Doctoral dissertation, Massachusetts Institute of Technology,
power, Energy. 187 (2019), 115992, https://doi.org/10.1016/j. 2004. https://dspace.mit.edu/handle/1721.1/41295 (accessed December 11,
energy.2019.115992. 2020).
[22] MATLAB - MathWorks - MATLAB & Simulink, (n.d.). https://in.mathworks.com/ [39] B.S. Oh, J.I. Lee, Study of Autonomous Control Systems for S-CO2 Power Cycle, in:
products/matlab.html (accessed August 2, 2020). 3rd European supercritical CO2 Conference, 2019.

14

You might also like