OTEC Maximum Net Power Output Using Carnot Cycle and Application to Simplify Heat Exchanger Selection
Abstract
:1. Introduction
- It addresses the lack of evaluation methods to efficiently select a heat exchanger for OTEC purposes and considers the trade-off between heat transfer performance and pressure drop.
- It is easily applicable for different heat exchangers as long as geometry, heat transfer coefficient correlation, and pressure drop correlation are provided.
- It is easily applicable to different seawater temperatures.
2. Description and Analysis
- The heat transfer coefficient of the working fluid is much greater than the seawater one as the working fluid undergoes a phase change [24].
- The thermal resistance due to fouling can be neglected.
- Changes in the water thermodynamic properties in the heat exchangers due to temperature variation can be neglected.
2.1. Carnot Cycle: Concept and Equations
2.2. Rankine Cycle: Concept and Equations
3. Optimization Process
3.1. Carnot Cycle
3.2. Rankine Cycle
4. Optimization Results
4.1. Carnot Cycle Results
4.2. Rankine Cycle Results
5. Conclusions
- For the sole purpose of a heat exchanger comparison, calculations based on the Carnot cycle for any source temperature were sufficient, as the cycle and temperature difference do not have an impact on the choice of the heat exchanger even though they do change the power output and optimized operating conditions. The Rankine cycle calculations presented a maximum net power output 23–27% lower than for the Carnot cycle that dropped ~10% each time a temperature difference decrease of 1 °C was observed. The evolution of the power output as a function of the temperature difference was found to follow the same trends as found in other studies.
- The maximum net power output was found to highly depend on the chosen heat exchangers. For the highest temperature difference, the most suitable heat exchangers among the three considered led to a maximum power output 158% and 165% higher than the worst heat exchanger for the Carnot and Rankine cycles, respectively.
- Due to the trade-off that exists between the heat transfer coefficient and the pressure drop, the heat exchanger presenting the highest heat transfer coefficient is not necessarily the one that will lead to the highest maximum net power output. In this study, for the Carnot cycle, PHE 2 competed with PHE 1 as it led to a maximum net power output that was only 3.7% lower than the one of PHE 1, even though its heat transfer coefficient was 35% lower.
- Heat exchangers with a high pressure drop and those with a low pressure drop have been found to have their own advantages and drawbacks. High pressure drop heat exchangers require lower Reynolds numbers, and therefore a smaller pumping power and/or a smaller pipe diameter are needed. Low pressure drop heat exchangers are less sensitive to a change in the Reynolds numbers, which can be useful in case a change in the operating conditions is needed. This is even more important as the results showed that negative net power output can be reached for low enough Reynolds numbers.
Author Contributions
Funding
Acknowledgments
Conflicts of Interest
Nomenclature
Nomenclature | ||
A | [m2] | Heat transfer surface area |
As | [m2] | Heat exchanger surface area |
B | [K/W] | |
Cp | [J/(kg·K)] | Specific heat |
C | [W/K] | heat capacity |
D | [m] | Equivalent diameter |
f | [-] | Friction factor |
h | [J/kg] | Specific enthalpy |
L | [m] | Length of the plate |
m | [kg/s] | Mass flow rate |
Nu | [-] | Nusselt number |
P | [kPa] | Pressure |
Pr | [-] | Prandtl number |
Q | [W] | Heat transfer rate |
Re | [-] | Reynolds number |
Rf | [m2·K/W] | Resistance due to fouling |
S | [m2] | Total cross-sectional surface area |
s | [J/(kg·K)] | Specific entropy |
T | [K] | Temperature |
U | [W/(m2·K)] | Overall heat transfer coefficient |
v | [m/s] | Mean velocity |
Wi | [m] | Width of the plate |
W | [W] | Power output |
w | [W/m2] | Power output per unit of surface area |
x | [-] | Vapor quality |
α | [W/(m2·K)] | Heat transfer coefficient |
ΔT | [K] | Temperature difference |
λ | [W/(m·K)] | Thermal conductivity |
Λ | [-] | Lagrange multiplier |
μ | [Pa·s] | Dynamic viscosity |
ρ | [kg/m3] | Density |
τ | [Pa] | heat exchanger wall shear stress |
Subscripts | |
cs | Cold source |
csi | Cold source inlet |
cso | Cold source outlet |
gross | Gross |
in | Inlet |
max | Maximum |
net | Net |
opt | Optimum |
out | Outlet |
p | Plate |
w | Water |
ws | Warm source |
wsi | Warm source inlet |
wso | Warm source outlet |
wf | Working fluid |
Appendix A
References
- International Energy Agency. CO2 Emissions from Fuel Combustion statistics 2017. Available online: https://www.iea.org/publications/freepublications/publication/CO2EmissionsfromFuelCombustionHighlights2017.pdf (accessed on 17 November 2019).
- Horowitz, C.A. Paris agreement. Int. Leg. Mater. 2016, 55, 740–755. [Google Scholar] [CrossRef]
- International Energy Agency. Key World Energy Statistics 2017. Available online: https://www.iea.org/publications/freepublications/publication/KeyWorld2017.pdf (accessed on 17 November 2019).
- Rajagopalan, K.; Nihous, G.C. An Assessment of Global Ocean Thermal Energy Conversion Resources with a High-Resolution Ocean General Circulation Model. J. Energy Resour. Technol. 2013, 135, 041202. [Google Scholar] [CrossRef]
- Transforming Our World: The 2030 Agenda for Sustainable Development. In A New Era in Global Health; Rosa, W. (Ed.) Springer Publishing Company: New York, NY, USA, 2017; ISBN 978-0-8261-9011-6. [Google Scholar]
- Wu, C. A Performance Bound for Real OTEC Heat Engines. Ocean Eng. 1987, 14, 349–354. [Google Scholar] [CrossRef]
- Ikegami, Y.; Bejan, A. On the Thermodynamic Optimization of Power Plants with Heat Transfer and Fluid Flow Irreversibilities. J. Sol. Energy Eng. 1998, 120, 139–144. [Google Scholar] [CrossRef]
- Kim, A.S.; Kim, H.-J.; Lee, H.-S.; Cha, S. Dual-use open cycle ocean thermal energy conversion (OC-OTEC) using multiple condensers for adjustable power generation and seawater desalination. Renew. Energy 2016, 85, 344–358. [Google Scholar] [CrossRef]
- Kim, A.S.; Oh, W.; Lee, H.-S.; Cha, S.; Kim, H.-J. Perspective of membrane distillation applied to ocean thermal energy conversion. IDA J. Desalin. Water Reuse 2015, 7, 17–24. [Google Scholar] [CrossRef]
- Johnson, D. The exergy of the ocean thermal resource and analysis of second-law efficiencies of idealized ocean thermal energy conversion power cycles. Energy 1983, 8, 927–946. [Google Scholar] [CrossRef]
- Kalina, A.I. Generation of Energy by Means of a Working Fluid, and Regeneration of a Working Fluid. U.S. Patent US4346561A, 31 August 1982. [Google Scholar]
- Zhang, X.; He, M.; Zhang, Y. A review of research on the Kalina cycle. Renew. Sustain. Energy Rev. 2012, 16, 5309–5318. [Google Scholar] [CrossRef]
- Uehara, H.; Ikegami, Y.; Nishida, T. OTEC System Using a New Cycle with Absorption and Extraction Processes. In Physical Chemistry of Aqueous Systems: Meeting the Needs of Industry; White, H.J., Jr., Sengers, J.V., Neumann, D.B., Bellows, J.C., Eds.; Begell House, Inc.: New York, NY, USA, 1995; pp. 862–869. ISBN 978-1-56700-445-8. [Google Scholar]
- Ikegami, Y.; Yasunaga, T.; Morisaki, T. Ocean Thermal Energy Conversion Using Double-Stage Rankine Cycle. J. Mar. Sci. Eng. 2018, 6, 21. [Google Scholar] [CrossRef]
- Dijoux, A.; Sinama, F.; Marc, O.; Clauzade, B. Working Fluid Selection General Method and Sensitivity Analysis of an Organic Rankine Cycle (ORC): Application to Ocean Thermal Energy Conversion (OTEC). 2017. hal-01653074. Available online: https://hal.archives-ouvertes.fr/hal-01653074/document (accessed on 17 November 2019).
- Bernardoni, C.; Binotti, M.; Giostri, A. Techno-economic analysis of closed OTEC cycles for power generation. Renew. Energy 2019, 132, 1018–1033. [Google Scholar] [CrossRef]
- Sinama, F. Thermodynamic analysis and optimization of a 10MW OTEC Rankine cycle in Reunion Island with the equivalent Gibbs system method and generic optimization program GenOpt. Appl. Ocean Res. 2015, 53, 54–66. [Google Scholar] [CrossRef]
- Sun, F.; Ikegami, Y.; Jia, B.; Arima, H. Optimization design and exergy analysis of organic Rankine cycle in ocean thermal energy conversion. Appl. Ocean Res. 2012, 35, 38–46. [Google Scholar] [CrossRef]
- Yasunaga, T.; Noguchi, T.; Morisaki, T.; Ikegami, Y. Basic Heat Exchanger Performance Evaluation Method on OTEC. J. Mar. Sci. Eng. 2018, 6, 32. [Google Scholar] [CrossRef]
- Solotych, V.; Lee, D.; Kim, J.; Amalfi, R.L.; Thome, J.R. Boiling heat transfer and two-phase pressure drops within compact plate heat exchangers: Experiments and flow visualizations. Int. J. Heat Mass Transf. 2016, 94, 239–253. [Google Scholar] [CrossRef]
- Morisaki, T.; Ikegami, Y. Performance Evaluation of Heat Exchangers in OTEC Using Ammonia/Water Mixture as Working Fluid. Open J. Fluid Dyn. 2013, 3, 302–310. [Google Scholar] [CrossRef]
- Nilpueng, K.; Wongwises, S. Two-phase gas–liquid flow characteristics inside a plate heat exchanger. Exp. Fluid Sci. 2010, 34, 1217–1229. [Google Scholar] [CrossRef]
- Uehara, H.; Ikegami, Y. Optimization of a Closed-Cycle OTEC System. J. Sol. Energy Eng. 1990, 112, 247–256. [Google Scholar] [CrossRef]
- Incropera, F.P.; Lavine, A.S.; Bergman, T.L.; DeWitt, D.P. Boiling and Condensation. In Fundamentals of Heat and Mass Transfer; John Wiley & Sons: Hoboken, NJ, USA, 2006; pp. 619–668. ISBN 978-0-471-45728-2. [Google Scholar]
- Incropera, F.P.; Lavine, A.S.; Bergman, T.L.; DeWitt, D.P. Heat Exchangers. In Fundamentals of Heat and Mass Transfer; John Wiley & Sons: New York, NY, USA, 2006; pp. 689–694. ISBN 978-0-471-47528-2. [Google Scholar]
- Ibrahim, O.M.; Klein, S.A.; Mitchell, J.W. Effects of Irreversibility and Economics on the Performance of a Heat Engine. J. Sol. Energy Eng. 1992, 114, 267–271. [Google Scholar] [CrossRef]
- Lemmon, E.W.; Bell, I.H.; Huber, M.L.; McLinden, M.O. NIST Standard Reference Database 23: Reference Fluid Thermodynamic and Transport Properties-REFPROP; Version 10.0; National Institute of Standards and Technology: Gaithersburg, MD, USA, 2018.
- Kushibe, M.; Lkegami, Y.; Monde, M.; Uehara, H. Evaporation Heat Transfer of Amonia and Pressure Drop of Warm Water for Plate Type Evaporator. Trans. Jpn. Soc. Refrig. Air Cond. Eng. 2005, 22, 403–415. [Google Scholar]
- Rody Oldenhuis orcid.org/0000-0002-3162-3660. “Minimize” Version 1.7, 2017/April/06. MATLAB Minimization Algorithm. Available online: http://nl.mathworks.com/matlabcentral/fileexchange/24298-minimize (accessed on 17 November 2019).
- Yeh, R.-H.; Su, T.-Z.; Yang, M.-S. Maximum output of an OTEC power plant. Ocean Eng. 2005, 32, 685–700. [Google Scholar] [CrossRef]
- VanZwieten, J.H.; Rauchenstein, L.T.; Lee, L. An assessment of Florida’s ocean thermal energy conversion (OTEC) resource. Renew. Sustain. Energy Rev. 2017, 75, 683–691. [Google Scholar] [CrossRef] [Green Version]
Heat Exchanger | PHE 1 | PHE 2 | PHE 3 |
---|---|---|---|
Length L (mm) | 960 | 718 | 1765 |
Width Wi (mm) | 576 | 325 | 605 |
Thickness t (mm) | 0.7 | 0.5 | 0.6 |
Space between plates δ (mm) | 4.00 | 3.95 | 2.68 |
Equivalent diameter D (mm) | 8.00 | 7.90 | 5.36 |
Material | SUS316 | Titanium | Titanium |
Thermal conductivity λp (W/(m·K)) | 16.3 | 21 | 21 |
Pattern | Herringbone (72°) | Herringbone (30°) | Fluting and drainage |
Number of plates | 120 | 20 | 52 |
Total heat transfer area A (m2) | 100.3 | 3.96 | 40.6 |
Total cross surface area S (m2) | 0.140 | 0.012 | 0.041 |
Heat Exchanger | d | γ | n | β | ξ |
---|---|---|---|---|---|
PHE 1 | 0.111 | 0.8 | 1/3 | 1.4863 | −0.0540 |
PHE 2 | 0.058 | 0.8 | 1/3 | 6.5059 | −0.3292 |
PHE 3 | 0.051 | 0.8 | 1/3 | 0.7371 | −0.1274 |
Heat Exchanger | αws (W/m2·K) | αcs (W/m2·K) | fws (−) | fcs (−) |
---|---|---|---|---|
PHE 1 | 20 569 | 14 450 | 0.913 | 0.945 |
PHE 2 | 13 259 | 9 334 | 0.307 | 0.379 |
PHE 3 | 11 191 | 7 843 | 0.242 | 0.263 |
PHE 1 | PHE 2 | PHE 3 | ||||
---|---|---|---|---|---|---|
Carnot | Rankine | Carnot | Rankine | Carnot | Rankine | |
wnet (W/m2) | 613 | 471 | 590 | 451 | 237 | 178 |
Rews | 8361 | 7876 | 10,690 | 10,121 | 6255 | 5753 |
Vws (m/s) | 0.837 | 0.788 | 1.08 | 1.03 | 0.934 | 0.859 |
Recs | 4383 | 4381 | 5618 | 5762 | 3274 | 3277 |
Vcs (m/s) | 0.832 | 0.831 | 1.08 | 1.11 | 0.927 | 0.928 |
© 2019 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license (http://creativecommons.org/licenses/by/4.0/).
Share and Cite
Fontaine, K.; Yasunaga, T.; Ikegami, Y. OTEC Maximum Net Power Output Using Carnot Cycle and Application to Simplify Heat Exchanger Selection. Entropy 2019, 21, 1143. https://doi.org/10.3390/e21121143
Fontaine K, Yasunaga T, Ikegami Y. OTEC Maximum Net Power Output Using Carnot Cycle and Application to Simplify Heat Exchanger Selection. Entropy. 2019; 21(12):1143. https://doi.org/10.3390/e21121143
Chicago/Turabian StyleFontaine, Kevin, Takeshi Yasunaga, and Yasuyuki Ikegami. 2019. "OTEC Maximum Net Power Output Using Carnot Cycle and Application to Simplify Heat Exchanger Selection" Entropy 21, no. 12: 1143. https://doi.org/10.3390/e21121143
APA StyleFontaine, K., Yasunaga, T., & Ikegami, Y. (2019). OTEC Maximum Net Power Output Using Carnot Cycle and Application to Simplify Heat Exchanger Selection. Entropy, 21(12), 1143. https://doi.org/10.3390/e21121143