On thermodynamic cycles for detonation enginescalculations using Cantera yields the postshock...

6
On thermodynamic cycles for detonation engines R. Vutthivithayarak, E.M. Braun, and F.K. Lu 1 Introduction Detonation engines are considered to potentially yield better performance than ex- isting turbo-engines in terms of improved thermodynamic efficiency, simplicity of manufacture and operation, and high thrust-to-weight or thrust-to-volume ratio, amongst other advantages. Much effort has been put into the development of pulsed detonation engines (PDEs), including thermodynamic cycle analysis. Thermody- namic analysis of PDEs usually makes use of one-dimensional models, based on the Chapman–Jouguet (CJ) and the Zeldovich–von Neumann–D¨ oring (ZND) theories, although increasingly sophisticated techniques, some involving numerical model- ing, have also been developed lately. It is now understood that the Humphrey cycle used to model an isochoric cycle underpredicts the performance of a PDE [1]–[4]. The so-called Fickett–Jacobs (FJ) cycle is based on the CJ model. While an im- provement over the Humphrey cycle, its reliance on the CJ model means that it fails to account for the physics espoused by the ZND model [1, 2]. In this paper, a discus- sion of the Humphrey and FJ cycles is given and the proper ZND cycle is suggested. These cycles are illustrated with a hydrogen/air mixture initially at STP. The use of a generic heat release parameter to construct the ZND cycle is provided. 2 The Humphrey Cycle Figure 1 shows the three ideal processes under discussion in both the pv and T s diagrams. The states portrayed in the plots are the total (or stagnation) states. The initial state of the reactants is (1). The hugoniot running through (1) is shown in Fig. 1 as a dashed line. The post-detonation hugoniot is also shown in the figure by Aerodynamics Research Center, Mechanical and Aerospace Engineering Department, University of Texas at Arlington, Arlington, Texas 76019, USA 1

Transcript of On thermodynamic cycles for detonation enginescalculations using Cantera yields the postshock...

Page 1: On thermodynamic cycles for detonation enginescalculations using Cantera yields the postshock pressure and specific volume as 2.8 MPa and 0.22 m3/kg respectively. The subsequent CJ

On thermodynamic cycles for detonation engines

R. Vutthivithayarak, E.M. Braun, and F.K. Lu

1 Introduction

Detonation engines are considered to potentially yield better performance than ex-isting turbo-engines in terms of improved thermodynamic efficiency, simplicity ofmanufacture and operation, and high thrust-to-weight or thrust-to-volume ratio,amongst other advantages. Much effort has been put into the development of pulseddetonation engines (PDEs), including thermodynamic cycleanalysis. Thermody-namic analysis of PDEs usually makes use of one-dimensionalmodels, based on theChapman–Jouguet (CJ) and the Zeldovich–von Neumann–Doring (ZND) theories,although increasingly sophisticated techniques, some involving numerical model-ing, have also been developed lately. It is now understood that the Humphrey cycleused to model an isochoric cycle underpredicts the performance of a PDE [1]–[4].The so-called Fickett–Jacobs (FJ) cycle is based on the CJ model. While an im-provement over the Humphrey cycle, its reliance on the CJ model means that it failsto account for the physics espoused by the ZND model [1, 2]. Inthis paper, a discus-sion of the Humphrey and FJ cycles is given and the proper ZND cycle is suggested.These cycles are illustrated with a hydrogen/air mixture initially at STP. The use ofa generic heat release parameter to construct the ZND cycle is provided.

2 The Humphrey Cycle

Figure 1 shows the three ideal processes under discussion inboth thep–v andT–sdiagrams. The states portrayed in the plots are the total (orstagnation) states. Theinitial state of the reactants is (1). The hugoniot running through (1) is shown inFig. 1 as a dashed line. The post-detonation hugoniot is alsoshown in the figure by

Aerodynamics Research Center, Mechanical and Aerospace Engineering Department, Universityof Texas at Arlington, Arlington, Texas 76019, USA

1

Page 2: On thermodynamic cycles for detonation enginescalculations using Cantera yields the postshock pressure and specific volume as 2.8 MPa and 0.22 m3/kg respectively. The subsequent CJ

2 Vutthivithayarak, Braun and Lu

another dashed line. This hugoniot was obtained using data obtained from the NASACEA code [5]. The data indicate a dimensionless heat releaseα = qρ1/p1 = 27.28.

Consider first the constant volume, Humphrey cycle [6] whichbas been proposedas a surrogate for the PDE cycle perhaps due to its simplicity[2, 3, 7]–[14]. In thisprocess, the working fluid is assumed to be compressed isochorically to state (2H)wherep2H = 0.8 MPa andT2H = 2550 K. The fluid is then expanded isentropicallyto reach (3H) wherep3H = 0.1 MPa andT3H = 1520 K. The increase in entropyfrom (1) to (3H) is∆s = 3.08 kJ/(kg·K). The cycle is closed by a fictitious isobaricprocess(3H)→ (1) of heat rejection to the open ambient conditions. A single valueof specific heat ratioγ = 1.242 appears sufficient for such an analysis but withR = 348 kJ/(kg ·K) and 396 kJ/(kg ·K) for the isochoric compression and for theisentropic expansion respectively.

3 The FJ Cycle

The FJ cycle is based on the CJ detonation model. As can be seenin Fig. 1, the cycleconsists of a compression process that brings the gas from state (1) to state (2CJ).This process is strictly a nonequilibrium one. However, within the one-dimensionalmodel of the detonation process, this process is identical to Rayleigh heating andthus can be regarded to be a process that is in local thermodynamic equilibrium[15]. Isentropic expansion occurs between (2CJ) and (3CJ) after which the cycle isclosed by a fictitious isobaric process to the initial state.

For this study, the detonation process raises the pressure to p2CJ = 1.5 MPa anddecreases the specific volume tov2CJ = 0.67 m3/kg with the same dimensionlessheat releaseα = 27.28 as for the Humphrey cycle. The isentropic expansion from(2CJ) to (3CJ) yieldsp3CJ = 0.1 MPa andv3CJ = 5.92 m3/kg respectively. Finally,a fictitious isobaric process returns both pressure and specific volume to the initialstate.

While the calculations of(p2CJ , v2CJ) are straightforward,(T2CJ , s2CJ) are morecomplicated to determine. The value of the gas constant changes from(1′) to (2CJ)as, for example, in computations using Cantera [17]. For simplicity, it was foundthat a linear variation ofR between the value at state(1′) and (2CJ) is accuratefor modeling the nonequilibrium heat release, which coincided with the equilibriumRayleigh heat release [16]. The temperature rises from the ZND value of 1545 Kto 2920 K and the entropy rises by 3.12 kJ/(kg ·K). The gas then expands isen-tropically from (2CJ) to (3CJ). State (3CJ) is different from state (3H) because theisentropic expansions arise from the different states (2CJ) and (2H) state, respec-tively. The values ofp3CJ , v3CJ and T3CJ are 0.1 MPa, 5.922 m3/kg and 1562 Krespectively.

Page 3: On thermodynamic cycles for detonation enginescalculations using Cantera yields the postshock pressure and specific volume as 2.8 MPa and 0.22 m3/kg respectively. The subsequent CJ

On thermodynamic cycles for detonation engines 3

(a) p–v diagram

(b) T–s diagram

Fig. 1 Ideal Humphrey (1→ 2H → 3H → 1), FJ (1→ 2CJ→ 3CJ→ 1) and ZND (1→ 1′ →2CJ→ 3CJ→ 1) cycles for a stoichiometric hydrogen/air mixture initiallyat STP.

4 The ZND Cycle

ZND theory captures the physics of a one-dimensional detonation wave. In this the-ory, the gas is shock compressed to the von Neumann spike, followed by exother-mic chemical reactions. The shock compression is assumed toproceed along theinert hugoniot which is equivalent to assuming local thermodynamic equilibrium.The subsequent heat release from exothermic chemical reactions, even though anonequilibrium process, is identical to that of Rayleigh heating [15] and brings the

Page 4: On thermodynamic cycles for detonation enginescalculations using Cantera yields the postshock pressure and specific volume as 2.8 MPa and 0.22 m3/kg respectively. The subsequent CJ

4 Vutthivithayarak, Braun and Lu

gas to the CJ state. (In fact, the CJ point is the tangent pointfrom the initial point tothe ZND point in thep–v diagram.)

The FJ model fails to account for the physics of a detonation process whereshock compression does not increase the pressure directly to the CJ point. Instead,the shock raises the pressure to the ZND point, commonly called the von Neu-mann spike. The pressure then decreases to the CJ value due tosupersonic Rayleighheating. It therefore appears that the ZND cycle, also called the PDE cycle in theliterature, is the most appropriate one amongst the three cycles evaluated in thispaper.

In this example where the stoichiometric hydrogen/air mixture is initially at STP,calculations using Cantera yields the postshock pressure and specific volume as 2.8MPa and 0.22 m3/kg respectively. The subsequent CJ value is the same as the FJcycle reported above. This is followed by the same isentropic expansion as the FJcycle, followed by a fictitious isobaric process to close thecycle.

For the particular example of a stoichiometric hydrogen/air mixture initially atSTP, the shock compression to the ZND point raises the temperature to 1531 K withan entropy increase to 1.366 kJ/(kg ·K). The heat addition that brings the gas fromthe ZND to the CJ point raises the temperature to 2920 K with a further increase ofentropy to 3.12 kJ/(kg ·K). From Cantera, the gas constant at these two points are397.6 and 348.22 kJ/(kg·K) respectively. The isentropic expansion to 1 atm lowersthe gs temperature to 1515 K. Finally, a fictitious isobaric process closes the cycle.

5 Comparison Between the Humphrey, FJ and ZND cycles

Experimental observations indicate that the detonation front is actually a complex,three-dimensional surface that defies any simplified analytical description. Thus,despite the one-dimensional nature of the ZND model, it is presently acceptable forengineering analysis. In this section, a comparison is provided on the discrepancy inestimating the net work by the Humphrey and FJ cycles when compared to the ZNDcycle. The cyclic specific net work produced, the cyclic specific heat input and theefficiency are given by

wout =∮

Ptdv (1)

qt,in =∮

Tt ds (2)

η = wout/qt,in (3)

In the above, the subscriptt is used to indicate that it is the total property thatis considered. This distinction may not be required in usualthermodynamic cycleanalysis. However, the kinetic energy in the detonation wave is a substantial portionof the total enthalpy that is available for energy conversion. A more detailed first-and second-law analysis exchange is planned. For the present, it can be stated thatthe ZND cycle accounts for the energy in the shock wave while the Humphrey and

Page 5: On thermodynamic cycles for detonation enginescalculations using Cantera yields the postshock pressure and specific volume as 2.8 MPa and 0.22 m3/kg respectively. The subsequent CJ

On thermodynamic cycles for detonation engines 5

FJ cycles do not, these two only accounting for heat addition. The respective data forthe stoichiometric oxyhydrogen example are displayed in Table 1. The table showsthat there is a large underestimation in performance parameters. These differenceswill likely affect the entire thermodynamic system. For example, these differencesfor a power production system will affect the design of the power generator and theheat exchanger. For a propulsion system, paramount considerations of weight andvolume may cause these to be too conservatively estimated.

Table 1 Performance Comparisons

Humphrey Fickett Zel’dovichJacobs von Neumann

Doring

Work (MJ/kg) 0.709 0.834 2.08Heat (MJ/kg) 1.07 1.3 2.95Efficiency (%) 66.5 64.3 70

6 Conclusions

An evaluation of the relative merits and shortfalls of threedifferent models forengineering analysis of pulse detonation engines was briefly given. While simpleto implement, the constant volume, or Humphrey, cycle does not adequately cap-ture the physics of the detonation phenomenon to provide a realistic estimate of thework. A more sophisticated model to account for the pressurerise in a detonationwave, known as the Fickett–Jacobs model, also underestimates the work. Finally,the Zel’dovich–von Neumann–Doring model appears to be the most appropriateone for use in cycle analysis of pulse detonation engines. While the shock processis a nonequilibrium one, the assumption of local thermodynamic equilibrium makesthe cycle analysis tractable. Similarly, the heat release due to exothermic reactionsbetween the ZND and CJ points was found to follow supersonic Rayleigh heatingand thus could be modeled as an equilibrium process. Future work includes develop-ing a general cycle analysis for either an airbreathing or a rocket propulsion systembased on the approach found in Mattingly [1].

References

1. Mattingly JD (2006) Elements of Propulsion: Gas Turbines andRockets, AIAA, Reston, Vir-ginia.

2. Heiser WH, Pratt DT (2002) Thermodynamic Cycle Analysis of Pulse Detonation Engines. JPropul Power 18(1):68–76.

Page 6: On thermodynamic cycles for detonation enginescalculations using Cantera yields the postshock pressure and specific volume as 2.8 MPa and 0.22 m3/kg respectively. The subsequent CJ

6 Vutthivithayarak, Braun and Lu

3. Kentfield JAC (2002) Fundamentals of Idealized Airbreathing Pulse-Detonation Engines. JPropul Power 18(1):77–83.

4. Wu Y, Ma F, Yang V (2003) System Performance and Thermodynamic Cycle Analysis ofAirbreathing Pulse Detonation Engines. J Propul Power 18(4):556–567.

5. Gordon S, McBride BJ (1994) Computer Program for Calculation of Complex Chemical Equi-librium Compositions and Applications. I. Analysis. NASA RP-1311. http://cea.grc.nasa.gov/

6. Humphrey HA (1909) An Internal-Combustion Pump, and Other Applications of a New Prin-ciple. Proc Inst Mech Eng 77(1):1075–1200.

7. Bussing TRA, Pappas G (1996) Pulse Detonation Engine Theory and Concepts. In: Develop-ments in High-Speed Vehicle Propulsion Systems, Murthy SNB and Curran ET (eds), AIAA,Reston, Virginia, 421–472.

8. Eidelman S, Yang X, (1998) Analysis of the Pulse Detonation Engine Efficiency. AIAA Paper98–3877.

9. Hutchins TE, Metghalchi M (2002) Energy and Exergy Analyses of the Pulse DetonationEngine. J Eng Gas Turbines Power 125(4):1075–1080.

10. Talley DG, Coy EB (2002) Constant Volume Limit of Pulsed Propulsion for a ConstantGamma Ideal Gas. J Propul Power 18(2):400–406.

11. Harris PG, Stowe RA, Ripley RC, Guzik SM (2006) Pulse Detonation Engine as a RamjetReplacement. J Propul Power 22(2):462–473.

12. Wintenberger E, Shepherd JE (2006) Model for the Performance of Airbreathing Pulse-Detonation Engines. J Propul Power 22(3):593–603.

13. Bellini R, Lu FK (2010) Exergy Analysis of a Pulse Detonation Power Device. J Propul Power26(4):875–878.

14. Li JL, Fan W, Wang YQ, Qiu H, Yan CJ (2010) Performance Analysis of the Pulse DetonationRocket Engine Based on Constant Volume Cycle Model. Appl ThermalEng 30(11–12):1496–1504.

15. Rao S (2010) Effect of Friction on the Zel’dovich–von Neumann–Doring to Chapman–JouguetTransition. MSAE thesis, Univ Texas Arlington.

16. Vutthivithayarak R (2011) Analysis of Pulse Detonation Turbojet Engines. Ph.D. dissertation,Univ Texas Arlington.

17. Goodwin D (2010) Cantera: Object-Oriented Software forReacting Flows.http://code.google.com/p/cantera/