• Non ci sono risultati.

The design of CO2-based working fluids for high-temperature heat source power cycles

N/A
N/A
Protected

Academic year: 2021

Condividi "The design of CO2-based working fluids for high-temperature heat source power cycles"

Copied!
8
0
0

Testo completo

(1)

ScienceDirect

Available online at Available online at www.sciencedirect.comwww.sciencedirect.com

ScienceDirect

Energy Procedia 00 (2017) 000–000

www.elsevier.com/locate/procedia

1876-6102 © 2017 The Authors. Published by Elsevier Ltd.

Peer-review under responsibility of the Scientific Committee of The 15th International Symposium on District Heating and Cooling.

The 15th International Symposium on District Heating and Cooling

Assessing the feasibility of using the heat demand-outdoor

temperature function for a long-term district heat demand forecast

I. Andrić

a,b,c

*, A. Pina

a

, P. Ferrão

a

, J. Fournier

b

., B. Lacarrière

c

, O. Le Corre

c

aIN+ Center for Innovation, Technology and Policy Research - Instituto Superior Técnico, Av. Rovisco Pais 1, 1049-001 Lisbon, Portugal bVeolia Recherche & Innovation, 291 Avenue Dreyfous Daniel, 78520 Limay, France

cDépartement Systèmes Énergétiques et Environnement - IMT Atlantique, 4 rue Alfred Kastler, 44300 Nantes, France

Abstract

District heating networks are commonly addressed in the literature as one of the most effective solutions for decreasing the greenhouse gas emissions from the building sector. These systems require high investments which are returned through the heat sales. Due to the changed climate conditions and building renovation policies, heat demand in the future could decrease, prolonging the investment return period.

The main scope of this paper is to assess the feasibility of using the heat demand – outdoor temperature function for heat demand forecast. The district of Alvalade, located in Lisbon (Portugal), was used as a case study. The district is consisted of 665 buildings that vary in both construction period and typology. Three weather scenarios (low, medium, high) and three district renovation scenarios were developed (shallow, intermediate, deep). To estimate the error, obtained heat demand values were compared with results from a dynamic heat demand model, previously developed and validated by the authors.

The results showed that when only weather change is considered, the margin of error could be acceptable for some applications (the error in annual demand was lower than 20% for all weather scenarios considered). However, after introducing renovation scenarios, the error value increased up to 59.5% (depending on the weather and renovation scenarios combination considered). The value of slope coefficient increased on average within the range of 3.8% up to 8% per decade, that corresponds to the decrease in the number of heating hours of 22-139h during the heating season (depending on the combination of weather and renovation scenarios considered). On the other hand, function intercept increased for 7.8-12.7% per decade (depending on the coupled scenarios). The values suggested could be used to modify the function parameters for the scenarios considered, and improve the accuracy of heat demand estimations.

© 2017 The Authors. Published by Elsevier Ltd.

Peer-review under responsibility of the Scientific Committee of The 15th International Symposium on District Heating and Cooling.

Keywords: Heat demand; Forecast; Climate change

Energy Procedia 129 (2017) 947–954

1876-6102 © 2017 The Authors. Published by Elsevier Ltd.

Peer-review under responsibility of the scientific committee of the IV International Seminar on ORC Power Systems. 10.1016/j.egypro.2017.09.125

10.1016/j.egypro.2017.09.125

© 2017 The Authors. Published by Elsevier Ltd.

Peer-review under responsibility of the scientific committee of the IV International Seminar on ORC Power Systems.

1876-6102

Available online at www.sciencedirect.com

ScienceDirect

Energy Procedia 00 (2017) 000–000

www.elsevier.com/locate/procedia

1876-6102 © 2017 The Authors. Published by Elsevier Ltd.

Peer-review under responsibility of the scientific committee of the IV International Seminar on ORC Power Systems.

IV International Seminar on ORC Power Systems, ORC2017

13-15 September 2017, Milano, Italy

The design of CO

2

-based working fluids for

high-temperature heat source power cycles

S. Lasala

a*

, D. Bonalumi

b

, E. Macchi

b

, R. Privat

a

and J.-N. Jaubert

a aUniversité de Lorraine, Laboratoire Réactions et Génie des Procédés, 1 rue Grandville, 54000 Nancy, France

bPolitecnico di Milano, Energy Department via Lambruschini 4, 20156 Milano, Italy

Abstract

The application of CO2 power cycles is advantageous to exploit high-temperature sources (500-800°C) in the case of available low-temperature heat sinks (15-25°C). However, their efficiency is strongly reduced for higher heat sink low-temperatures. At these temperatures, due to the low-critical temperature of CO2 (about 31°C), CO2 is in fact compressed in the supercritical vapor phase rather than in the liquid phase, thus increasing energetic demand for compression. One of the solutions envisaged to overcome this problem is the addition of one or more chemicals that allow having a mixture with a higher critical temperature than the one of pure CO2. This preserve the working fluid compression in its liquid phase, even in the case of heat sinks with temperatures greater than 25°C.

This research shows that the addition to CO2 of a properly selected chemical component enables to increase the critical temperature up to 45°C with relevant improvements of cycle efficiency with respect to pure-CO2 power cycles. In particular, it summarizes the most relevant criteria to be accounted for when selecting CO2-additives. Moreover, the paper warns of the thermodynamic effects deriving from adding to CO2 a second characterized by a much more high critical temperature, such as the occurrence of infinite-pressure critical points and multiple-phase liquid-liquid and vapor-liquid critical points. Finally, the paper analyses the thermodynamic properties of a high-critical temperature CO2-based mixture, suitable for these applications, that presents multiple phase critical points. In this regard, it is specified that the paper also aims at filling a knowledge gap in the study of thermodynamic properties of mixtures presenting how do enthalpy and specific volume change in response to pressure variations in the event of liquid-liquid and vapour-liquid critical points. Finally, we present the comparison between performances of power cycles which use, as working fluid, either pure CO2 or the novel designed higher temperature CO2-based mixture.

© 2017 The Authors. Published by Elsevier Ltd.

Peer-review under responsibility of the scientific committee of the IV International Seminar on ORC Power Systems.

Keywords: CO2-based mixtures; Critical points; Power cycles; High-temperature heat source

* Corresponding authors: silvia.lasala@univ-lorraine.fr;

Available online at www.sciencedirect.com

ScienceDirect

Energy Procedia 00 (2017) 000–000

www.elsevier.com/locate/procedia

1876-6102 © 2017 The Authors. Published by Elsevier Ltd.

Peer-review under responsibility of the scientific committee of the IV International Seminar on ORC Power Systems.

IV International Seminar on ORC Power Systems, ORC2017

13-15 September 2017, Milano, Italy

The design of CO

2

-based working fluids for

high-temperature heat source power cycles

S. Lasala

a*

, D. Bonalumi

b

, E. Macchi

b

, R. Privat

a

and J.-N. Jaubert

a aUniversité de Lorraine, Laboratoire Réactions et Génie des Procédés, 1 rue Grandville, 54000 Nancy, France

bPolitecnico di Milano, Energy Department via Lambruschini 4, 20156 Milano, Italy

Abstract

The application of CO2 power cycles is advantageous to exploit high-temperature sources (500-800°C) in the case of available low-temperature heat sinks (15-25°C). However, their efficiency is strongly reduced for higher heat sink low-temperatures. At these temperatures, due to the low-critical temperature of CO2 (about 31°C), CO2 is in fact compressed in the supercritical vapor phase rather than in the liquid phase, thus increasing energetic demand for compression. One of the solutions envisaged to overcome this problem is the addition of one or more chemicals that allow having a mixture with a higher critical temperature than the one of pure CO2. This preserve the working fluid compression in its liquid phase, even in the case of heat sinks with temperatures greater than 25°C.

This research shows that the addition to CO2 of a properly selected chemical component enables to increase the critical temperature up to 45°C with relevant improvements of cycle efficiency with respect to pure-CO2 power cycles. In particular, it summarizes the most relevant criteria to be accounted for when selecting CO2-additives. Moreover, the paper warns of the thermodynamic effects deriving from adding to CO2 a second characterized by a much more high critical temperature, such as the occurrence of infinite-pressure critical points and multiple-phase liquid-liquid and vapor-liquid critical points. Finally, the paper analyses the thermodynamic properties of a high-critical temperature CO2-based mixture, suitable for these applications, that presents multiple phase critical points. In this regard, it is specified that the paper also aims at filling a knowledge gap in the study of thermodynamic properties of mixtures presenting how do enthalpy and specific volume change in response to pressure variations in the event of liquid-liquid and vapour-liquid critical points. Finally, we present the comparison between performances of power cycles which use, as working fluid, either pure CO2 or the novel designed higher temperature CO2-based mixture.

© 2017 The Authors. Published by Elsevier Ltd.

Peer-review under responsibility of the scientific committee of the IV International Seminar on ORC Power Systems.

Keywords: CO2-based mixtures; Critical points; Power cycles; High-temperature heat source

(2)

948 S. Lasala et al. / Energy Procedia 129 (2017) 947–954

2 S. Lasala, D. Bonalumi, E. Macchi, R. Privat, J.N. Jaubert/ Energy Procedia 00 (2017) 000–000

1. Introduction

The ever-increasing necessity to reduce the environmental impact of industrial and urban energy conversion processes has led engineers to consider the use of mixtures as working fluids of thermodynamic power and refrigeration cycles. Mixing different fluids in different amounts, in fact, allows enhancing specific thermodynamic characteristics of the mixture that may lead cycle performances to improve; moreover, the use of multicomponent fluids enables the reduction of the environmental impact and hazardous properties of some of the basic pure fluids under mixing.

For example, performances of real-gas-Brayton-cycles may significantly increase if the working fluid is compressed in the liquid phase, rather than in the gaseous one. However, when cooling sources are only available at temperatures higher than the critical temperature of the pure working fluid, the liquid-phase compression of that fluid cannot evidently take place. To allow the achievement of subcritical conditions, it has been proposed [2, 3] to increase the critical temperature of the working fluid by adding a small amount of a high-critical temperature second component. Vapor-liquid critical temperatures of resulting mixtures lie, in fact, between the critical temperatures of the two pure components.

This paper focuses on the design of working fluids for closed power cycles that exploit a high temperature heat source comprised between 500 °C and 800 °C to convert the thermal input into electricity. Possible high-temperature heat sources are, for example, nuclear and solar technologies or high-temperature waste thermal sources recovered from industrial processes. The amount of thermal energy made available by these sources is, in general, not sufficient to enable the application of efficient steam power cycles of which minimum commercial size is 10 MWel [4].

Moreover, mainly due to the thermal decomposition of organic working fluids, that for the most stable organic molecules occurs around 300 - 400 °C [5], the high-grade of these heat sources does not allow the application of ORC especially above 400°C.

CO2 cycles represent potential reliable candidates for these applications. When water at about 15°C is available to

cool these cycles, CO2 may be favourably pumped in subcritical liquid conditions. However, the efficiency of these

cycles is highly affected by the temperature of the cooling source: the low-critical temperature of CO2 (about 31°C)

entails the necessity of using entirely supercritical CO2 Brayton cycles when temperature of the available cooling

source is above 25°C, rather than the less - compression power consumer transcritical CO2-cycle.

This paper aims at showing how to increase the efficiency of CO2 cycles, used in the event of high-temperature

cooling sources, by adding to CO2 a properly selected second component. To this aim, section 2 presents the main

criteria that should be accounted for when selecting such a fluid. Furthermore, section 3 presents the thermodynamic analysis of a promising mixture under patenting. This section also shows how do enthalpy and specific volume change in response to pressure variations across the critical region, in the event of multiple vapour-liquid and liquid-liquid critical points that characterize this system. Finally, section 4 compares performance results of a power cycle operating either with pure CO2 or with the designed CO2-based working fluid.

Nomenclature

T temperature [K] P pressure [bar]

h specific enthalpy [kJ/kg] v specific volume [m3/kg]

s specific entropy [kJ/kgK] Q heat [kJ/kg]

LL liquid-liquid equilibrium VL vapor-liquid equilibrium x, y, z mole fractions of liquid, vapor and global composition, respectively η efficiency

2. Criteria for CO2-additive selection

The selection of CO2-additives should be performed so that their addition to CO2 leads to a mixture being (1)

thermally stable; (2) characterized by a critical temperature greater than 45 °C; (3) moderate mixture critical pressure; (4) environmental benign (low-GWP and zero-ODP); (5) non-toxic and non-flammable. Since pure CO2 already fulfills

(3)

S. Lasala, D. Bonalumi, E. Macchi, R. Privat, J.N. Jaubert / Energy Procedia 00 (2017) 000–000 3

all the mentioned – not exhaustive – requirements, it is evident that the non-compliance with them would be associated to the selected additive.

The simultaneous compliance with all these criteria is currently a challenge. For this reason, it is worth distinguishing between necessary and preferable conditions to be fulfilled: criteria (1-3) must be necessarily satisfied to guarantee the correct working of the power cycle, while the other mentioned criteria (4-5) generally represent a desiderata which could lead to the preferential selection between additives.

Moreover, considering criterion (1), it is specified that, in the absence of experimental and modelling results, the thermal stability of the mixture could be ensured, at an earlier stage, providing additives being non-reacting with CO2

and thermally stable (in their pure form) at the operating conditions of interest. In the following, further details are provided on criteria (2) and (3), that actually represent the only criteria on the basis of which the amount of additive is defined.

The addition of a small amount of a high-critical temperature component to CO2, to allow the desired moderate

increase in the mixture critical temperature, could be accompanied by an extreme increase in critical pressure that, for some classes of mixtures, may achieve indefinitely high values [6]. In other cases, the addition of a second component might lead to the emergence of multi-phase equilibria (e.g., Vapour-Liquid-Liquid Equilibrium) that could give rise to the coexistence of two critical points: one vapour-liquid and one liquid-liquid critical point.

These thermodynamic singularities in the critical region of specific binary mixtures started to be classified by van Konynenburg and Scott [6]. A first classification level discerns between binary mixtures characterized either by a continuous or discontinuous locus of vapour-liquid equilibrium critical points.

To visualize the three types of critical loci that could be encountered in the context of this work, we present, in the following, three CO2-based binary mixtures, respectively characterized by these three types of critical loci, containing

one of the following high-critical temperature fluids: C6F14; NOA (an environmental friendly Non-Organic Additive

currently under patenting process); H2O.

Figures 1(a)-(c) aim at showing the resulting locus of critical points (solid black lines) of the three systems as a function of temperature, pressure and composition of the mixtures, which extremes coincide with the critical points of the two pure components (round mark), C1 and C2, end-points of their saturation curves (dotted black lines). Dotted

red lines represent equilibrium lines where three vapour-liquid-liquid phases co-exist. A more detailed description is provided in the followings.

Fig. 1(a) represents the continuous locus of critical points that characterize the mixture CO2-C6F14. As it can be

observed, the addition of a small amount of C6F14 to CO2 allows the increase of the critical temperature of the resulting

mixture, positively maintaining a moderate critical pressure. The only limit of C6F14 is represented by its thermal

stability, which temperature limit has been experimentally attested to be around 350 °C [3].

Fig. 1(b) shows the discontinuous locus of critical point of the CO2-NOA mixture. This discontinuity derives from

the emergence of a three-phase line that results in the co-existence of two critical points in VL and LL conditions (red dotted line in the zoomed area). More details about this peculiarity are given in the next section. As for the CO2-C6F14

system, the critical pressure undergoes a acceptable moderate increase in the critical temperature of the system. Fig. 1(c) shows how the addition of a very small amount of water to CO2 results in an extremely limited increase

of critical temperature of the system (from the critical temperature of pure CO2 (304.1K) to, about, 305.5K) and in an

indefinitely high-value of the critical pressure. This system and similar CO2-based mixtures characterized by

infinite-discontinuities, occurring at the low-CO2 concentrations of our interest, cannot evidently be applied.

Actually, our research for proper additives has led to the preliminary selection of the organic C6F14 and C6F6, of

NOA and of the further non-organic reacting N2O4 (which decomposes into NO2, NO and O2 at the considered

operating conditions). The effect of adding N2O4 is currently under modelling study. In the next section we focus on

(4)

950 4 S. Lasala, D. Bonalumi, E. Macchi, R. Privat, J.N. Jaubert/ Energy Procedia 00 (2017) 000–000 S. Lasala et al. / Energy Procedia 129 (2017) 947–954

Fig. 1. Global Phase Equilibrium Diagrams of CO2 (C1) + C6F14 (C2) (fig. (a)) / NOA (C2) (fig. (b)) / H2O (C2) (fig. (c)). These diagrams result

from the application of the Peng-Robinson [7] equation of state, with the following calibrated binary interaction parameters parameters: kCO2-C6F14

= 0.0151 [8], kCO2-NOA = 0.07 (references are not provided because NOA is under patenting), kCO2-H2O = 0.0504 [8]

3. Definition of mixture composition

As mentioned in section 2, mixtures under design must be finally characterized by a critical temperature greater than 45 °C, resulting from the addition to CO2 of an additive (in this case NOA). This section thus quantifies the

variation of the critical temperature of CO2+NOA mixtures, as a function of its global composition, with the aim of

(1) defining the global mixture composition to be finally used as working fluid of the analyzed CO2-based transcritical

power cycles and (2) describing the thermodynamic features of the designed mixture. It is further specified that all calculations performed in the following result from the application of the Peng-Robinson equation of state [7] calibrated over vapour-liquid equilibrium and enthalpy change due to mixing experimental data available in the literature. It is not possible to specify references for these data since the fluid is under patenting. The class of cubic equation of state models has been selected since it actually enables the most accurate representation of locus of critical points of mixtures, with respect to SAFT models [9] and multi-parameter models. The latter, in fact, would require to be optimized over a high-number of experimental data currently unavailable. Finally, it is specified that the applied standard Peng-Robinson equation of state is based on van der Waals mixing rules grounded on assumptions of ideal random mixing. More precise calculations of loci of critical points could be obtained by applying more accurate mixing rules based on local composition models [10] which would more realistically account for molecular interactions. All thermodynamic calculations reported in this paper have been performed with an in-house Fortran code.

Fig. 2 shows how does the critical temperature of the CO2-NOA mixture variates with mixture composition. In

particular, it is possible to attest that a critical temperature equal to 45°C can be obtained by mixing 97.5%mol of CO2

with 2.5%mol of NOA. As it is visible from fig. 2, such a global composition is close to a discontinuity in the locus of

critical points, already observed in fig. 1(b), resulting from the occurrence of three co-existing phases (dotted red line). It is worth observing that each one of the points of the three-phase dotted red line observed in the T-P plane of fig. 1(b) corresponds to three isothermal and isobar points on the plane T-zC1, each one characterized by the composition

of one of the three phases at equilibrium conditions (see fig. 2).

Furthermore, it is highlighted that this diagram results from the application of the Peng-Robinson-78 [7] equation of state, optimized over a scarce number of available experimental data. The optimized binary interaction parameter is thus highly uncertain (kij = 0.07 ± 0.03 at 95% conf. level). As a direct consequence, either the actual occurrence or the

precise location of such a three-phase line is uncertain. There is thus a possibility that, in the reality, the selected global composition (CO2/NOA : 97.5%mol/2.5%mol) actually leads to the emergence of multiple critical points arising from

the presence of multiple-phase equilibria. However, it is shown in the following that, also in this case, the effect on thermodynamic properties (and on the functioning of the power cycle components) is negligible.

0 1 2 3 4 5 6 7 8 9 293 323 353 383 413 443 473 P [ M Pa ] T [K] 100% C1 (a) 100% C2 0 5 10 15 20 25 30 293 343 393 443 493 543 593 643 693 P [ M Pa ] T [K] 100% C1 (b) 100% C2 9.65 9.7 9.75 9.8 322.5 322.8 323.1 ZOOM 0 50 100 150 200 250 293 393 493 593 693 P [ M Pa ] T [K] (c) 100% C2 5 6 7 8 293 298 303 308 100% C1 ZOOM

(5)

S. Lasala et al. / Energy Procedia 129 (2017) 947–954 951

Fig. 2. Global Phase Equilibrium Diagrams of CO2 (C1) + NOA(C2). Projection on T-zC1 plane.

We thus evaluate, in the following, the isothermal change of volume and enthalpy resulting from pressure variations on the CO2-NOA mixture, at one of the temperatures where the zoomed area of fig. 1(b) reveals the presence of

multiple isothermal critical points (i.e., in the range 322.55 K - 323.04 K). In particular, fig. 1(b) shows that, considering a temperature of 322.75 K, the resulting mixture is characterized by two equal-pressure vapour-liquid and liquid-liquid critical points. This is also shown in the resulting isothermal phase diagram represented in fig. 3. It is also specified that, in general, no more than two critical points could be found.

Fig. 3. Isothermal P-xy phase diagram of CO2 (C1)+NOA(C2) at 322.75 K, with zoom on critical region showing two LL and VL critical points

(on the left). Figure on the right shows the pressure of critical points and of the three-phase line corresponding to the considered isotherm.

Fig. 4 and Fig. 5 respectively show how do isothermal volume and enthalpy change across a critical region characterized by multiple critical points, as a consequence of pressure variations. In particular, it is highlighted that the presence of NOA in CO2 enhances the increase of density around the critical region, which is characterized by

high-density liquid-liquid critical points, enabling the further reduction of power consumed to compress this fluid in the vicinity of such region. These considerations, on the one hand, and the necessity of correctly calculating the thermodynamic properties of these fluids, on the other, reveal that the occurrence of such discontinuities should be carefully predicted when designing binary mixtures characterized by a specific critical point temperature, for the mentioned industrial application.

300 350 400 450 500 550 600 650 0 0.2 0.4 0.6 0.8 1 T [K ] zC1[-] 322 322.5 323 323.5 0.85 0.9 0.95 1 ZOOM 0 1 2 3 4 5 6 7 8 9 10 0 0.2 0.4 0.6 0.8 1 P [ M Pa ] x,yC1[-] T = 322.75 K 9.69 9.70 9.71 9.72 0.85 0.9 0.95 1 ZOOM 9.66 9.67 9.68 9.69 9.7 9.71 9.72 9.73 9.74 9.75 9.76 9.77 322.5 322.65 322.8 322.95 323.1 P [ M Pa ] T [K] Detail of fig. 1(b) T = 322.75 K

(6)

952 6 S. Lasala, D. Bonalumi, E. Macchi, R. Privat, J.N. Jaubert/ Energy Procedia 00 (2017) 000–000 S. Lasala et al. / Energy Procedia 129 (2017) 947–954

Fig. 4. Isothermal P-v phase diagram of CO2 (C1) + NOA(C2) at T = 322.75 K (solid blue curve) with isothermal (T = 322.75 K) P-v curves at

zC2 = 0.025 (dotted black line) and zC2 = 0.075 (solid black line). The zoom on the critical region shows two LL and VL critical points.

Fig. 5. Isothermal P-h phase diagram of CO2 (C1) + NOA(C2) at T = 322.75 K (solid blue curve) with isothermal (T = 322.75 K) P-v curves at

zC2 = 0.025 (dotted black line) and zC2 = 0.075 (solid black line). The zoomed area on the critical region shows two LL and VL critical points.

4. Performance of CO2 versus CO2+NOA power cycles

This section presents preliminary results of the performance of recuperative CO2-Brayton cycles and recuperative

CO2-NOA transcritical cycles. The heat source has been considered at constant temperature. In such case the best

configuration should be the one with split-flow and recompression, but for this study a simpler layout has been selected. Calculations have been performed considering assumptions reported in Table 1. The minimum pressure for the pure CO2 cycle is the one that maximizes the efficiency.

Table 1. Assumptions for calculations on recuperative power cycle configurations

CO2 CO2-NOA (2.5%mol)

Maximum T of the cycle (Tmax) [K] 673.15 – 873.15

Minimum T of the cycle (Tmin) [K] 313.15

Maximum P of the cycle (Pmax) [MPa] 30

Minimum P of the cycle (Pmin) [MPa] Value which maximizes power efficiency Saturation pressure at Tmin

Isentropic turbine efficiency [-] Values from correlations reported in [11] for three-stages turbines (about 0.92)

Isentropic compressor efficiency [-] 0.85

Pressure drops (ΔP/Pin) Recuperator: 0.04 / Primary heat exchanger: 0.02 / Condenser: 0.02

Minimum ΔT between hot and cold streams in recuperator [K] 15

Electro-mechanical efficiency [-] 0.95 2 4 6 8 10 12 0 0.0002 0.0004 0.0006 0.0008 0.001 P [ M Pa ] v [m3/mol] 9.60 9.65 9.70 9.75

5.0E-05 7.0E-05 9.0E-05 1.1E-04

P [ M Pa ] v [m3/mol] ZOOM 2 4 6 8 10 12 -750000 -650000 -550000 -450000 -350000 P [ M Pa ] h [J/mol] 9.60 9.65 9.70 9.75 -480000 -460000 -440000 -420000 -400000 P [ M Pa ] h [J/mol] ZOOM

(7)

S. Lasala et al. / Energy Procedia 129 (2017) 947–954 953 6

Fig. 4. Isothermal P-v phase diagram of CO2 (C1) + NOA(C2) at T = 322.75 K (solid blue curve) with isothermal (T = 322.75 K) P-v curves at

zC2 = 0.025 (dotted black line) and zC2 = 0.075 (solid black line). The zoom on the critical region shows two LL and VL critical points.

Fig. 5. Isothermal P-h phase diagram of CO2 (C1) + NOA(C2) at T = 322.75 K (solid blue curve) with isothermal (T = 322.75 K) P-v curves at

zC2 = 0.025 (dotted black line) and zC2 = 0.075 (solid black line). The zoomed area on the critical region shows two LL and VL critical points.

4. Performance of CO2 versus CO2+NOA power cycles

This section presents preliminary results of the performance of recuperative CO2-Brayton cycles and recuperative

CO2-NOA transcritical cycles. The heat source has been considered at constant temperature. In such case the best

configuration should be the one with split-flow and recompression, but for this study a simpler layout has been selected. Calculations have been performed considering assumptions reported in Table 1. The minimum pressure for the pure CO2 cycle is the one that maximizes the efficiency.

Table 1. Assumptions for calculations on recuperative power cycle configurations

CO2 CO2-NOA (2.5%mol)

Maximum T of the cycle (Tmax) [K] 673.15 – 873.15

Minimum T of the cycle (Tmin) [K] 313.15

Maximum P of the cycle (Pmax) [MPa] 30

Minimum P of the cycle (Pmin) [MPa] Value which maximizes power efficiency Saturation pressure at Tmin

Isentropic turbine efficiency [-] Values from correlations reported in [11] for three-stages turbines (about 0.92)

Isentropic compressor efficiency [-] 0.85

Pressure drops (ΔP/Pin) Recuperator: 0.04 / Primary heat exchanger: 0.02 / Condenser: 0.02

Minimum ΔT between hot and cold streams in recuperator [K] 15

Electro-mechanical efficiency [-] 0.95 2 4 6 8 10 12 0 0.0002 0.0004 0.0006 0.0008 0.001 P [ M Pa ] v [m3/mol] 9.60 9.65 9.70 9.75

5.0E-05 7.0E-05 9.0E-05 1.1E-04

P [ M Pa ] v [m3/mol] ZOOM 2 4 6 8 10 12 -750000 -650000 -550000 -450000 -350000 P [ M Pa ] h [J/mol] 9.60 9.65 9.70 9.75 -480000 -460000 -440000 -420000 -400000 P [ M Pa ] h [J/mol] ZOOM

S. Lasala, D. Bonalumi, E. Macchi, R. Privat, J.N. Jaubert / Energy Procedia 00 (2017) 000–000 7

Results obtained for the two thermodynamic cycles, in the specific case of Tmax = 873.15 K, reveal that the use of

the CO2-NOA mixture as working fluid enables the increase of cycle efficiency, from 35.1% to 39.5%.

Figure 6 shows the T-s diagrams of these two power cycles, having the same minimum temperature of 313.15 K, together with the calculated thermodynamic properties. At lower maximum temperatures, the improvement in cycle efficiency is still more pronounced, as shown in Fig. 7. To gain a better understanding of the advantages brought about by the use of the proposed mixture, a comparison of efficiency penalties obtained by a second-law analysis with reference to an ideal reversible (Carnot) cycle operating between the maximum and the minimum cycle temperature is performed for pure and mixture cycles. The real efficiency ŋ of a closed cycle is computed by:

ηII= ηrev - Tmin ∑ ∆SNi=1 i⁄Qin=ηrev -∑ ∆Ni=1 ηII,i

where η𝑟𝑟𝑟𝑟𝑟𝑟 is the efficiency of a reversible cycle operating with constant temperature Tmax heat source and constant

temperature Tmin heat sink and ΔηII,i are 2nd-law efficiency penalties related to the N-cycle irreversible processes [1].

As shown in Fig. 8, the performances of both power cycles improve with maximum temperature for both working fluids. The merits of mixture are more evident at lower maximum temperatures, where real gas effects are more pronounced. In particular, efficiency penalties taking place in the heat rejection and in the compression decrease significantly by adopting a condensing mixture. On the opposite, the efficiency penalties related to heat exchanges in the recuperator and in heat introduction increase, but their effect is much lower than the former advantages.

In addition to these considerations about cycle efficiencies, it is observed that higher logarithmic mean temperature difference between streams within the recuperator, resulting from the use of the CO2-NOA mixture rather than of pure

CO2, enables the reduction of heat exchanger surface. A final remark, that also favors the use of the CO2-NOA mixture,

concerns the lower temperature value of the hot stream entering the primary heat exchanger (point numbered with “3” in Fig. 6), which allows a major cooling of the heat source – whether available at variable temperature.

Fig. 7. Trend of cycle efficiency (solid line: mixture CO2-NOA, dashed line: pure CO2) as a function of the cycle maximum temperature, at

maximum pressure of 30 MPa

25% 29% 33% 37% 41% 45% 650 725 800 875 950 1025 ηI Tmax[K]

Fig. 6. T-s diagrams of CO2+NOA cycle (on the left) and of pure CO2 cycle (on the right), both at Tmax = 873.15 K and under assumptions

reported in Table 1. 290 380 470 560 650 740 830 920 -80 -60 -40 -20 0 20 T [ K] s [J/molK] 1 2 3 4 5 6 Critical point of the mixture P T h s v

MPa K J/mol J/mol/K m3/kmol

1 8.26 313.15 -411599 -59.68 0.0749 2 30.00 357.69 -409916 -58.98 0.0606 3 28.80 629.27 -390318 -16.89 0.1831 4 28.22 873.15 -376122 2.36 0.2738 5 8.78 730.25 -383416 3.24 0.6951 6 8.43 372.69 -403014 -33.75 0.2735 Point number 290 380 470 560 650 740 830 920 -70 -50 -30 -10 10 T [ K] s [J/molK] 1 2 3 4 5 6 Critical point of CO2 P T h s v

MPa K J/mol J/mol/K m3/kmol

1 8.47 313.15 -398959 -47.533 0.1263 2 30.00 396.53 -396570 -46.625 0.0791 3 28.80 659.06 -379621 -12.991 0.1954 4 28.22 873.15 -367588 2.9932 0.2737 5 9.00 741.07 -374202 4.6002 0.6905 6 8.64 411.53 -391152 -25.337 0.3376 Point number

(8)

954 S. Lasala et al. / Energy Procedia 129 (2017) 947–954

8 S. Lasala, D. Bonalumi, E. Macchi, R. Privat, J.N. Jaubert/ Energy Procedia 00 (2017) 000–000

Fig. 8. Second-law efficiency losses, relative to Carnot efficiency ηCarnot = 1 – Tmin/Tmax (Tmin = 313.15 K) . The lowest the sum of the efficiency

penalties, the highest is the thermodynamic quality of the power cycle. 5. Conclusions and future works

This paper shows that the increase the critical temperature of the working fluid from the one of CO2, by adding to

CO2 a properly selected additive, enables the efficient application of CO2-based power cycles for high-temperature

heat sources even when the heat sink temperature is above 25°C. Performance calculations are shown considering results obtained from the use of a specific CO2-based mixture under patenting, being stable up to - at least - 600°C,

environmentally friendly, non-toxic, non-flammable and characterized by a moderate critical pressure (lower than 100 bar). According to results here presented, the novel mixture enables the increase of cycle efficiency up to 2.9 percentage points compared with the pure CO2 cycle.

The promising results of the mixture are worthy of further investigation related to the mixture composition, alternative layouts of the power plant and the evaluation of the design of their components.

Acknowledgements

The authors wish to thank ASME IGTI ORC Power Systems – KCORC for supporting this research.

References

[1] Macchi E. Theoretical basis of the Organic Rankine Cycle, In: Organic Rankine Cycle (ORC) Power Systems. Woodhead Publishing Series in Energy. 2017

[2] Angelino G, Invernizzi CM, The role of real gas Brayton cycles for use of liquid natural gas physical exergy, Applied Thermal Engineering, 2011, 31: 827-833.

[3] Lasala S, Invernizzi C, Iora P, Chiesa P, Macchi E. Thermal stability analysis of perfluorohexane. Energy Procedia 2015, 75:1575-1582 [4] Invernizzi CM, Iora P, Bonalumi D, Macchi E, Roberto R, and Caldera M. Titanium tetrachloride as novel working fluid for high temperature

Rankine Cycles: Thermodynamic analysis and experimental assessment of the thermal stability. Applied Thermal Engineering 2016. 107: 21-27

[5] Invernizzi CM, Bonalumi D. Thermal stability of organic fluids for Organic Rankine Cycle systems. In: Organic Rankine Cycle (ORC) Power Systems. Woodhead Publishing Series in Energy. 2017

[6] Privat R, Jaubert JN, Classification of global fluid-phase equilibrium behaviors in binary systems. Chem. Eng. Res. Des. 91 (2013) 1807–1839. [7] Robinson DB, Peng DY. The Characterization of the Heptanes and Heavier Fractions for the GPA Peng − Robinson Programs; Gas Processors

Association: Tulsa, OK, 1978.

[8] Lasala S. Advanced cubic equations of state for accurate modelling of fluid mixtures. Application to CO2 capture systems. PhD thesis. Milano.

2016

[9] Sugata PT, Hertanto A, Maciej R. Recent Advances and Applications of Statistical Associating Fluid Theory. Industrial & Engineering Chemistry Research 2008, 47: 8063–8082.

[10] Lasala S, Chiesa P, Privat R, Jaubert JN. VLE Properties of CO2-Based Binary Systems Containing N2, O2 and Ar: Experimental Measurements

and Modelling Results with Advanced Cubic Equations of State. Fluid Phase Equilibria, 2016, 428: 18−31.

[11] Macchi E, Astolfi M. Axial flow turbines for Organic Rankine Cycle applications, In: Organic Rankine Cycle (ORC) Power Systems. Woodhead Publishing Series in Energy. 2017

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 MIX

673.15 KCO2 773.15 KMIX CO2 873.15 KMIX CO2 973.15 KMIX CO2

ΔηII /ηCa

rn

ot

Riferimenti

Documenti correlati

Con ordinanza del 16 maggio 2002 la Suprema Corte federale tedesca ha rimesso alla Corte di giustizia alcune questioni interpretative emerse nel corso di un procedimento

Per quanto riguarda le funzioni sanitarie, si posso identificare due cluster: Lombardia, RER e Tosca- na che assegnano la definizione della strategia alle singole aziende (capogrup-

Significa lavorare sulla qualificazione delle “opere d’arte” della nuova infrastruttura, con particolare attenzione ai punti più delicati, all’incrocio con le altre infrastrutture

Knowing the maximum and actual stream of accumulated air, we can determine the process' efficiency (equation 6) being the relationship between actually

Data of the two heat pumps working during one heating season was analysed to determine the influence of heat sink temperature on the Seasonal Performance Factor.. Results of

c L represents the total amount of active sites (DFO or 3,4-HP) in the solid phase involved in Fe(III) complexation (mmol/g) and α L is the ratio between the total ligand and its

The time intervals with the highest flaring activity (grey bands) used in the analysis of MAGIC data are defined following the Bayesian block method.. The arrow refers to the peak