See discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/230601230 Design and Numerical Analysis of Aerospike Nozzles with Different Plug Shapes to Compare their Performance with a Conventional Nozzle Conference Paper · March 2005 CITATIONS READS 15 6,941 3 authors, including: Mehdi Nazarinia Arash Naghib-Lahouti Heriot-Watt University University of Toronto 28 PUBLICATIONS 136 CITATIONS 16 PUBLICATIONS 110 CITATIONS SEE PROFILE SEE PROFILE Some of the authors of this publication are also working on these related projects: Mapping the Urban Heat Island (UHI) effect in Dubai and its impact on building energy performance View project All content following this page was uploaded by Mehdi Nazarinia on 19 May 2014. The user has requested enhancement of the downloaded file. AIAC-11 Eleventh Australian International Aerospace Congress WC0023 Design and Numerical Analysis of Aerospike Nozzles with Different Plug Shapes to Compare their Performance with a Conventional Nozzle Mehdi Nazarinia 1, Arash Naghib-Lahouti 2 and Elhaum Tolouei 1 1 Senior Research Engineer 2 Faculty Member Aerospace Research Institute, P.O. Box 14665-834, Tehran, Iran Tel: +98-21-8362010, Fax: +98-21-8362011 E-mail: mnazarinia@ari.ac.ir Summary: The objective of the present study is to demonstrate thrust advantage of a small axisymmetric aerospike nozzle in comparison with the equivalent conventional (convergentdivergent) nozzle through numerical simulation, and to study the effects of geometrical parameters on the aerospike’s performance. Internal and external flow of the aerospike nozzle and the equivalent conventional nozzle has been numerically analyzed in different ratios of ambient pressure to design exhaust pressure (Patm/Pdes) representing optimum, under- and over-expansion working conditions. Results indicate that the aerospike nozzle can deliver as much as 5% more thrust in over-expansion conditions. To address practical aspects of using an aerospike nozzle, effects of geometric parameters including base curvature, and different values of plug truncation ranging from 0 to 75%, on thrust and base temperature distribution of the nozzle have been studied. Results show that higher values of base truncation are more appropriate for flight at altitudes higher than design altitude (under-expansion condition), while lower values are more appropriate for flight below design altitude. It has also been shown that adding curvature to the base of a truncated plug can improve local temperature distribution without any significant effect on thrust. Keywords: nozzle, aerospike, truncation, design, thrust, CFD. Nomenclature CF CP F h Isp L & m M M P PR ℜ T r x thrust coefficient specific heat at constant pressure (J/kgK) thrust (N) altitude (m) specific impulse (s) length (m) mass flow rate (kg/s) Mach number molecular weight (g/mol) absolute static pressure (N/m2) Pressure Ratio (= Patm / Pdes) Universal gas constant, 8314 J/kg.mol.K absolute static temperature (K) radial coordinate axial coordinate Eleventh Australian International Aerospace Congress Sunday 13 – Thursday 17 March 2005 Fifteenth National Space Engineering Melbourne, Victoria, Australia Symposium Greek Symbols γ ratio of specific heats, Cp/Cv Subscripts 1 stagnation atm atmosphere (external flow) b base des design e exhaust i inlet mom momentum p pressure t throat Introduction Expansion and discharge of a gas in different propulsion systems, e.g. jet engines and rockets, is always accomplished by a nozzle. Thrust of a conventional nozzle with fixed geometry, discharging in atmosphere can be expressed by the following simple relation: & Ve + (Pe − Patm )A e F=m (1) This relation indicates that for a nozzle designed to have a constant value of Pe (also known as design exhaust pressure, Pdes), thrust is affected by change of altitude. At the design altitude, where Patm = Pdes, the second term of the above relation (known as pressure thrust) is zero, and the nozzle is said to be working in “optimum condition”. At altitudes lower than the design altitude, where Patm > Pdes, pressure thrust assumes a negative value, and loss of thrust is inevitable. These conditions, which occur at altitudes ranging from ground level to the design altitude, are known as “over-expansion” conditions. Beside the inherent loss of thrust, the conventional nozzle might suffer problems including shock waves and flow separation in the divergent section, thrust oscillation, and flow asymmetry in over-expansion conditions. Ever since jet and rocket propulsion systems have emerged, researchers have invented and implemented many types of nozzles, mainly to increase the thrust performance of nozzles in off-design working conditions. Among these various designs, features of the aerospike nozzle have attracted researchers since mid-1950s [1]. Many theoretical studies of the aerospike nozzle have been carried out in 1960s. Berman and Crimp’s work [2] is an example of these studies, in which issues such as analytical design methods, thrust vectoring, and integration with solid- and liquid-propellant systems have been addressed. Rao [3] has presented a more accurate method based on calculus of variations for design of the plug in 1961, and Lee and Thompson [4] have developed the first computer program for plug nozzle design based on Rao’s work in 1964. In early 1970s, thermal and strength problems of the aerospike nozzle and development of more efficient methods for fabrication of conventional nozzles led to a decline in research activities in this field. In 1990s, NASA initiated the SSTO (Single Stage to Orbit) project, which required a propulsion system with maximum efficiency in a broad range of working altitudes. The aerospike propulsion system was selected for this purpose, and extensive research and development led to successful testing of the RS2200 aerospike engine [5]. Cancellation of this project in 2001 has led to another decline in research activities in this field in the United States in the recent years, but the aerospike nozzle is still a live research topic in Europe and Japan. Hageman et al. [6, 7] have proposed application of a large scale aerospike nozzle in the post-Ariane 5 launch vehicles to DLR. Tomita et al. [8] and Sakamoto et al. [9] have carried out experimental studies of axisymmetric and linear aerospike nozzles, and Fujii and Ito [10] have studied many aspects of aerospike nozzles numerically. Most recently, two different groups in the United States have applied the axisymmetric aerospike nozzle in propulsion system of sounding rockets, and demonstrated considerable gain in the rockets’ performance [11, 12]. The aerospike nozzle is considered to have better overall performance compared to the conventional bell nozzle since expansion of the jet is not bounded by a wall and the exhaust flow can adjust itself to the environment by changing the jet boundary. In addition, the nozzle performance is considered not to be influenced by the cutting off the nozzle because the base pressure compensates the loss of thrust force. The base pressure can also be increased by injecting a secondary flow at the base which may be tapped off the exhaust flow and injected at the base [13]. For detailed discussion on the off-design performance comparisons of aerospike and conventional nozzles, the reader is kindly referred to reference [14]. Despite its remarkable advantage over a conventional nozzle, practical application of an ideal aerospike nozzle has structural and thermal problems, especially because of the sharp end of the plug. These problems can be averted to a great extent by truncating the base of the plug. Base truncation reduces weight and length of an aerospike nozzle, and increases its strength, thus facilitating practical application of this kind of nozzle. In this paper, effect of different values of base truncation on thrust performance of an aerospike nozzle has been studied in various working conditions. To further understand the mechanism by which base flow affects performance of the nozzle, variation of pressure and thrust in this area has been studied in more detail. Effects of rounding the base of a truncated plug on flow pattern, thrust performance, and temperature distribution are also studied. Numerical flow simulation This section describes numerical modeling and analysis of internal and external flow of the aerospike nozzle with different plug shapes, and the equivalent convergent-divergent nozzle using a commercial CFD code. The objective of the analysis is comparison of thrust produced by aerospike nozzles with different plug shapes with a conventional nozzle in off-design working conditions. Geometry Two methods, namely the method of characteristics [15] and an approximate method based on ideal rocket assumptions [16], have been used to design the central plug of the aerospike nozzle, which have resulted in almost identical plug shapes. Schematic of the axisymmetric aerospike nozzle, displaying the characteristic lines representing Prandtl-Meyer expansion waves is shown in Fig. 1. The two methods are implemented to design an aerospike nozzle in 4-5 kN thrust class, with specifications originally introduced in [11], assuming the following design parameters: Chamber Pressure: Design Altitude: Optimum Thrust: Specific Impulse: Mass Flow: Propellant: P1 = 2067857 N/m2 h = 3657.6 m F = 4443.9 N Isp = 235 s & = 3.25758 kg/s m Ethanol-Oxygen, γ = 1.21 The following specifications are also calculated for the nozzle using the relations derived in the approximate method [14]: Throat Angle: Throat Area: Throat Radius: Exhaust Area: Exhaust Radius: Exhaust Mach number: θt = 57.1665° At = 1.438569×10-3 m2 rt = 0.045221 m Ae = 7.04612×10-3 m2 re = 0.047888 m Me = 2.80434 Fig.1: Schematic of the axisymmetric aerospike nozzle, displaying the characteristic lines representing Prandtl-Meyer expansion waves. Details of the design process by both methods can be found in [14]. Results of the two methods show a maximum difference of 5% in radius. In the following sections, results of the approximate method are used as the baseline for numerical flow simulation due to their better agreement with the original plug shape. Fig. 2 shows geometry of the aerospike and the conventional nozzle and the preceding convergent sections. (a) Aerospike nozzle (inset: plug contour) (b) Equivalent conventional convergent-divergent nozzle Fig. 2: Nozzle Geometry Two methods were also examined to define the truncated plug shapes: implementation of the relations of the approximate method to design a whole new truncated plug, and cutting off the ideal aerospike nozzle at different stations of the central plug without altering the remaining portion. Exhaust and throat area of the aerospike nozzle is calculated using the following relations in the approximate method (Fig. 1): 2 2 A e = p re − rb (2) At ( p (r = − rt Cos? t e 2 2 ) ) (3) Considering Eqns 2 and 3, it can be concluded that for constant and specified values of throat area and expansion ratio (e=Ae/At), increasing base radius causes the exhaust radius to increase, thereby increasing the throat radius. Fig. 3 shows the mentioned effect for an aerospike nozzle with different values of plug truncation of 0% (ideal), 75%, 50%, and 25%. It is clearly seen in the figure that for low values of truncation (25% and 50%) the increase in radius of the aerospike nozzle is more pronounced. Considering restraints imposed by combustion chamber and the vehicle’s radius, the application of such truncated aerospike nozzles may not be practical. Fig. 3: Trend of aerospike nozzle radius increase for different plug shapes with constant and specified values of At and e Therefore, in this study in order to determine the profile of truncated aerospike nozzles with lengths of 75%, 50%, and 25% of the ideal aerospike nozzle, 25%, 50%, and 75% of the ideal plug were cut off respectively in a way similar to [10]. As for the remaining portions, instead of redesigning the aerospike nozzle contour, these parts remained unchanged to prevent changes in exhaust and throat radius. Geometric specifications of the truncated nozzles are presented in Table 1. Table 1: Geometric specifications of truncated aerospike nozzles % of plug length 100 75 50 25 Plug length (m) 0.127530 0.0942975 0.0628650 0.0314325 Base radius (m) 0.004363 0.011551 0.020910 Unlike a conventional nozzle, in which the throat (where M=1) remains at the same longitudinal position in all working conditions (optimum, over-, and under-expansion), the throat of an aerospike nozzle covers a different angular domain for each ratio of Patm/Pdes. Therefore, in modeling of an aerospike nozzle, the convergent section should also be included in the solution domain, and the throat should be allowed to establish by itself, rather than imposed at a fixed location. To achieve this objective, the solution domains for both nozzles analyzed herein include a convergent section with a length of L=0.2m and the following inlet properties: Ai = 7.212855×10-3 m2 (Ai/At = 5.01384) Mi = 0.13 Pi = 0.98984 P1 = 2045430 N/m2 Authors’ experience [17] has also shown that to satisfy the convergence criteria, far-field boundaries of the solution domain should have a distance of at least 40re from the nozzle’s exhaust surface. Considering axial symmetry of the problem, numerical solution is carried out in half of the entire domain, bounded by a boundary with axial symmetry condition. Grid The solution domain has been divided into 4 regions (labeled Fluid 1-4 in Fig. 4), to facilitate definition of different initial conditions in each region. In case of the conventional nozzle, regions 1-3 (which include convergent and divergent sections of the nozzle) are discretized using a structured grid of quadrilateral cells, while in region 4 (the surrounding domain) an unstructured grid of triangular cells has been used for discretization. The resulting grid has 50767 cells (Fig. 5a). In case of the aerospike nozzle, geometric variations in regions 1 and 2 make it impossible to generate a structured grid with acceptable quality. Therefore, the entire solution domain is discretized by an unstructured grid of triangular cells. The total number of cells for the ideal aerospike nozzle a finer grid with 340544 cells has also been generated to study the effect of grid size on results (Fig. 6), which is shown in Figure 5b, is 85136. Different geometries of the truncated aerospike nozzles cause region 2 of the solution domain (Fig. 5) to have different number of cells. Total number of grid cells for the 75%, 50%, and 25% cases are 90582, 93334, and 95892, respectively, which are comparable with the grid cells of the ideal aerospike nozzle (85136). As it was mentioned previously, adding curvature to the base of a truncated plug can improve local temperature distribution without any significant effect on thrust. To analyze this effect, a similar aerospike nozzle with a 50% truncated plug with rounded base was analyzed in the same working conditions as the three other cases. In this case, the straight line defining the base was replaced with a circular arc tangent with the contour of the remaining portion of the plug (Fig. 7). Fig. 5f shows the grid used for the 50% plug rounded end with the total number of grid cells of 95374. Fig. 4: Solution domain and boundary conditions Fluid properties Average properties of combustion products resulting from chemical reaction of ethanol and oxygen have been assumed for the fluids in all regions. These properties, which satisfy the following relation: CP = are as follows: γ = 1.21, CP = 1286.68 J/kgK, M = 37.23 g/mol. ?ℜ ? (? − 1) (4) (a) Conventional nozzle (b) Ideal Aerospike nozzle, 100% plug length (c) Truncated aerospike nozzle, 75% plug length (d) Truncated aerospike nozzle, 50% plug length (e) Truncated aerospike nozzle, 25% plug length (f) Truncated aerospike nozzle, 50% plug with rounded base Fig. 5: Grid for analysis of conventional and aerospike nozzles with different plug shapes, displayed in vicinity of the plug Fig. 6: Comparison of the results of numerical analysis using the original and refined grids for the ideal aerospike nozzle Fig. 7: Geometry of the 50% truncated aerospike nozzle with rounded base Boundary conditions The condition implied at the inlet of the convergent section is inlet with specified mass flow, with the following boundary values: & = 3.25757 kg/s m T0 = 1577.826 K P = 2045430 N/m2 Seven different cases of atmospheric conditions (Patm/Pdes) have been analyzed, which correspond to different working conditions (including over-expansion, optimum and underexpansion). Boundary values imposed at pressure farfield and pressure outlet boundaries in the 7 cases are presented in Table 2. Apart from case 4, which corresponds to the nozzles’ optimum working condition as introduced in [11], the other cases have been chosen hypothetically to represent under-expansion (cases 1-3) and over-expansion (cases 5-7) conditions. Supersonic Mach numbers of external flow have been selected for all cases in order to facilitate satisfaction of the convergence criteria. Table 2. Values imposed at farfield boundaries in different analysis cases M T (K) T0 (K) &DVH Patm/Pdes P (N/m2) 1 4.00 257736 1.5 288.15 356.04 2 3.10 200000 1.5 288.15 356.04 3 1.57 101325 1.5 288.15 356.04 4 1.00 64434 2.80434 264.378 482.689 5 0.41 26415.4 2.5 223.12 369.54 6 0.19 12037.1 3.0 216.61 421.31 7 0.10 6410 3.0 216.61 421.31 Initial conditions Numerical values of flow variables vary greatly in different regions of the solution domain in analysis of a nozzle with external flow. For example, while values close to stagnation properties prevail at the inlet of the convergent section, the external flow might involve substantially lower pressures and higher (even supersonic) Mach numbers. In such circumstances, initialization of a flow variable with a constant value throughout the entire domain can make convergence difficult or sometimes impossible. To deal with this problem, properties at nozzle inlet, throat and exhaust surface (which can be roughly estimated using one dimensional isentropic flow relations [18]) have been used to define custom field functions describing initial values of axial velocity, pressure and temperature in Fluid 1-3 regions. The initial state of the solution domain for the 75% truncated aerospike nozzle, which is generated using the above-mentioned functions, is displayed in Fig. 8, as an example. Fig. 8: Contours of Mach number at the initial state of the solution domain for the 75% truncated aerospike nozzle Analysis features Combustion products have been assumed to behave as a viscous (laminar) and compressible ideal gas (P=ρRT). The coupled implicit method described in [19] has been used for solution of the four governing equations (continuity, conservation of momentum in longitudinal and radial directions, and conservation of energy), considering severe compressibility effects existing in the solution domain. Fluxes of convected variables at cell walls are approximated by the first order upwind scheme. Courant’s number has been set to 0.7 for cases 5, 6 and 7, and to 2.0 for the other cases. Two criteria have been posed for convergence. One is reduction of the global residual of solution of all governing equations to the order of 10-5, and the other is establishment of mass balance between inlet, far-field and outlet boundaries, which is checked by integration of mass flow through the mentioned boundaries at each iteration. In all cases, the solution process has been continued until both criteria have been satisfied. Results and discussion In this section, flow pattern of the ideal aerospike nozzle in different working conditions is compared to that of the truncated aerospike and conventional nozzles in this section. Performance of all the nozzles in off-design conditions is also compared by calculation of thrust for each case. Qualitative study of the results Exhaust flow of the aerospike nozzle is characterized by formation of a series of expansion waves, which originate from the upper lip of the convergent section (Fig. 1). Since the exhaust flow is not bounded by a solid wall, these expansion waves can adjust their intensity and domain to match the exhaust flow with the external flow. For the ideal aerospike, in over-expansion conditions, the domain covered by these waves ends before the end of the plug. At that station, flow properties are close to those of the optimum condition, which usually involve a higher Mach number and a lower pressure compared to the external flow. From this station onwards, flow encounters reflection of the expansion waves in form of a series of compression waves, which increase the pressure and reduce Mach number to a value close to that of the external flow. The above-mentioned process can be noticed in Fig. 9a in form of contours of static pressure for case 1, where Patm/Pdes = 4.00. Effect of these expansion and compression waves on pathlines (Fig. 9b) is so that the exhaust flow leaves the exhaust surface straight and without further contraction and residual radial velocity. But for the truncated aerospike nozzles, the situation is such that the expansion waves originated from the upper lip of the convergent section will face the truncated portion of the plug, while in the ideal case these expansion waves meet the plug surface. The flow facing the truncation first encounters a sharp expansion, then by continuing its way to the centre of the plug base a compression, stagnating exactly at the centre of the plug base. This phenomenon is due to the formation of two symmetric vortices in the base of the plug (Fig. 10), which counteract the effect of each other at two locations, one of which is located at the centre of the plug base, where flow conditions change into the stagnation conditions. It should be pointed out that regardless of the amount of truncation and the extent of the plug base area, the flow parameter distribution pattern is the same. Figs. 9c, 9e, and 9g clearly show the above-mentioned process in form of contours of static pressure for different plug shapes. Effect of these vortices and also the expansion and compression waves on pathlines is also seen in Figs. 9d, 9f, and 9h. The exhaust flow of a conventional nozzle, on contrary, does not have the chance to adapt itself to the condition prevailing outside the nozzle before leaving the nozzle. After leaving the nozzle, flow is compressed through a series of compression waves originating from the edge of the exhaust surface, which resemble converging shock waves. These compression waves contract the flow, and impose a radial velocity component, which contributes to loss of thrust in over-expansion conditions [14]. (a) Contours of static pressure, 100% plug length (b) Pathlines, 100% plug length (c) Contours of static pressure, 75% plug length (d) Pathlines, 75% plug length (e) Contours of static pressure, 50% plug length (f) Pathlines, 50% plug length (g) Contours of static pressure, 25% plug length (h) Pathlines, 25% plug length Fig. 9: Flow pattern of the ideal and truncated aerospike nozzles in over-expansion conditions (Patm/Pdes = 4.00) Fig. 10: Flow pattern at the base of a truncated nozzle For an aerospike nozzle in optimum condition, the domain covered by expansion waves ends right at the end of the plug. At that station, flow properties match those of the external flow, and no considerable further expansion or compression are encountered (Fig. 11a). Pathlines also indicate that the flow leaves the exhaust surface without any radial component (Fig. 11b). But for the truncated aerospike nozzles, once again the exhaust flow will face the truncated portion of the plug before expanding to optimum values. Flow pattern in the optimum condition is the same as over-expansion condition, explained in the preceding paragraphs. Figs. 11c, 11e, and 11g show the expansion process in form of contours of static pressure for different plug shapes. Effect of the base vortices and also the expansion and compression waves on pathlines is also seen in Figs. 11d, 11f, and 11h. It should be noted that in spite of existence of rotational flow at the base area, pathlines will continue their way parallel to the axis of the plug after the longitudinal position corresponding to the end of a virtual ideal plug, even for the 25% truncated aerospike nozzle (Fig. 11h). For the conventional nozzle in optimum condition, expansion waves inside the divergent section affect the flow so that properties at the intersection of nozzle axis and the exhaust surface match those of the external flow [14]. (a) Contours of static pressure, 100% plug length (b) Pathlines, 100% plug length (c) Contours of static pressure, 75% plug length (d) Pathlines, 75% plug length (e) Contours of static pressure, 50% plug length (f) Pathlines, 50% plug length (g) Contours of static pressure, 25% plug length (h) Pathlines, 25% plug length Fig. 11: Flow pattern of the ideal and truncated aerospike nozzles in optimum conditions (Patm/Pdes = 1.00) For an aerospike nozzle in under-expansion conditions, the domain of the expansion waves covers an area larger than the plug itself, and the flow continues to expand after the plug. Effect of these waves on the exhaust flow can be seen in Fig. 12a in form of contours of static pressure for case 7, where Patm/Pdes = 0.10. As seen in Fig. 12b, pathlines also keep diverging for some distance after the plug. For truncated plug nozzles in under-expansion, flow pattern at the plug base and pathlines are same as those seen in over-expansion and optimum conditions. It should be mentioned that parts of the expansion waves encounter the rotational base flow, hence drifting the base vortices, especially the second stagnation point seen in Fig. 9, away from the nozzle axis. Figs. 12c, 12e, and 12g show the process in form of contours of static pressure for different plug shapes. Effect of base vortices and also the expansion and compression waves on pathlines is also seen in Figs. 12d, 12f, and 12h. In case of the conventional nozzle in under-expansion conditions, expansion of the flow is continued outside the nozzle through a series of expansion waves originating from the edge of the exhaust surface. This expansion causes pathlines to diverge before adapting to the external flow, in a manner similar to the aerospike nozzle [14]. (a) Contours of static pressure, 100% plug length (b) Pathlines, 100% plug length (c) Contours of static pressure, 75% plug length (d) Pathlines, 75% plug length (e) Contours of static pressure, 50% plug length (f) Pathlines, 50% plug length (g) Contours of static pressure, 25% plug length (h) Pathlines, 25% plug length Fig. 12: Flow pattern of the ideal and truncated aerospike nozzles in under-expansion conditions (Patm/Pdes = 0.10) Effect of aerospike base truncation on thrust performance Thrust is the best measure of performance for a nozzle intended for use in a propulsion system. Thrust of an axisymmetric convergent-divergent nozzle is calculated by the following integral on the exhaust surface: F = ∫ ?Vx2 dA + ∫ (P − Patm )dA = Fmom + FP (5) It is clearly seen that the second term, known as pressure thrust, will be negative in overexpansion conditions, causing a considerable loss of thrust [14]. For an axisymmetric aerospike nozzle, thrust is calculated using the following relation: F = ∫ ( ρVdA)Vx + ∫ (P − Patm )Cos? t dA + ∫ (P − Patm )Sin? t dA = F1 + F2 + F3 (6) The first two terms are integrated on the throat of the convergent section, while the third term is integrated on the entire plug surface. This relation shows that the pressure component of thrust of the aerospike nozzle never assumes a negative value in over-expansion conditions. Therefore, loss of thrust of an aerospike nozzle is always less than the conventional nozzle in similar conditions. Figure 13 shows that the ideal aerospike nozzle can deliver as much as 5% more thrust in over-expansion conditions in comparison with the conventional nozzle [14]. Effect of base truncation can also be seen in Fig. 13. The thrust calculation procedure is the same as that of the ideal aerospike nozzle, noting that the term F3, also contains the thrust produced by the base as well. It should be pointed out that according to Fig. 13, base truncation, contrary to what may be expected, does not have any significant effect on thrust in under-expansion conditions. This graph shows that for under-expansion conditions, thrust of all nozzles are similar to each other. But in over-expansion conditions the 25% plug has thrust equal or less than the conventional nozzle. In these conditions the 75% and 50% plugs have less thrust compared to the ideal aerospike nozzle, but present better overall performance than the conventional nozzle. Fig. 13: Variation of thrust delivered by different nozzles studied herein with external flow pressure In order to understand the reason why thrust delivered by different truncated aerospike nozzles remains unchanged in under-expansion conditions but differs in over-expansion conditions, it is necessary to approach the thrust components differently. Dividing thrust into the three following components, explains this phenomenon more clearly: 1) Thrust produced by the nozzle convergent section (F1+F2) 2) Thrust produced by the plug surface (part of F3) 3) Thrust produced by the plug base (other part of F3) Fig. 14 shows the contribution of the above-mentioned components to the total value of thrust for under- and over-expansion conditions. In under-expansion conditions, when the plug is truncated, its lateral area decreases. Therefore the pressure thrust produced by the plug reduces. On the other hand, the thrust generated by the base region increases because of the increase of the base area. These two effects compensate each other, and the total nozzle thrust becomes almost the same for different nozzle truncation. This effect can be seen most clearly for the 25% plug. But in over-expansion conditions, the situation is totally different. In these conditions, as the nozzle length becomes shorter, hence decreasing the plug area, thrust produced by the plug still decreases, while as the atmosphere pressure is higher than the exhaust pressure thrust produced by the base pressure would have a negative value. So by increasing truncation, the negative value of base thrust will increase, hence decreasing total thrust in over-expansion conditions. Fig. 14b shows that for the 25% plug, total thrust is lowest. Conv+Plug+Base Conv Conv+Base 0.7 Plug Base Conv 1 0.9 0.6 0.8 0.5 F / Fdes 0.7 F/Fdes 0.6 0.4 0.5 0.3 0.4 0.3 0.2 0.2 0.1 0.1 0 25 50 75 100 0 25 50 75 %Plug %Plug (a) Patm / Pdes = 0.10 (b) Patm / Pdes = 4.00 100 Fig. 14: Proportion of the thrust for each truncated nozzle To describe the reason with more detail, average base pressure for the 75, 50 and 25% plug nozzle is plotted versus external flow pressure in Fig. 15. The solid line represents the atmospheric pressure which decreases as the pressure ratio (Patm/Pdes) decreases or the altitude becomes higher. At low altitudes (i.e., over-expansion conditions), base pressure linearly increases as atmospheric pressure increases. Therefore, the pressure thrust which is produced by the pressure difference between the atmosphere and the base becomes small and even negative in most cases at low altitudes. On the other hand, at high altitudes, pressure at the base remains constant despite variation of altitude. As altitude is increased, atmospheric pressure decreases and the difference between base pressure and atmospheric pressure increases, hence increasing the base thrust. 10 6 Base Pressure Plug-75%Base P Plug-50%Base P Plug-25%Base P Patm 10 5 10 4 103 1 2 3 4 P atm/P des Fig. 15: Variation of the averaged base pressure with the external flow pressure To compare behavior of the portion of thrust produced by the plug’s base pressure (base thrust), the non-dimentionalized base thrust coefficient is defined as follows: FBase = CF Base 1 ?P M 2 S e Base 2 des (7) This coefficient is plotted against Patm/Pdes for different values of plug truncation in Fig. 16. This figure indicates that base thrust, which has a positive value in under-expansion conditions, decreases as atmospheric pressure is increased. This decrease continues in overexpansion condition for the 25% plug, explaining its lower thrust performance in these conditions. For the 50% and 75% plugs, base thrust begins to increase again beyond a certain value of atmospheric pressure, and even assumes a positive value in case of the 75% plug. This behavior explains the superior performance of plugs with lower amount of truncation in over-expansion conditions. 0.5 Plug-75%CFBase Plug-50%CFBase Plug-25%CFBase 0.4 0.3 0.2 CFBase 0.1 0 -0.1 -0.2 -0.3 -0.4 -0.5 1 2 3 4 Patm/P des Fig. 16: Variation of the base thrust coefficient with the external flow pressure Effect of base curvature of a truncated plug on thrust, and temperature distribution As it was previously mentioned, adding curvature at the base of a truncated plug can improve local temperature distribution without any significant effect on thrust. To show this, the rounded base nozzle introduced in Figs. 5f and 7 was analysed in the same working conditions as the other nozzles. Fig. 17 shows, thrust of the 50% rounded base and truncated plug nozzles against external flow pressure. It can be seen that there is no difference in thrust performance of a rounded base and a truncated plug nozzle. Therefore, adding curvature to the base has a negligible effect on thrust of the nozzle. 1.1 1.05 F/Fdes-50% F/Fdes-50%-Rounded 1 F/Fdes 0.95 0.9 0.85 0.8 0.75 0.7 0 1 2 3 4 P atm/P des Fig. 17: Comparison of the variation of thrust with external flow pressure for a 50% rounded and truncated plug However, this modification can have a positive effect on base temperature distribution. Table 3 compares average values of base temperature for the 50% rounded base and truncated plugs, calculated by integration of base surface temperature distribution. This table shows that average base temperature of the rounded base plug is considerably lower than that of the truncated plug, hence making the former more attractive from fabrication and structural point of view. Table 3. The averaged base temperature of the 50% rounded and truncated plug Plug Patm/Pdes 0.10 4.00 Rounded Truncated 945.38 K 1128.80 K 1377.91K 1435.01 K Conclusion Flow structure and the performance of truncated aerospike nozzles in off-design conditions has been compared with conventional convergent-divergent and ideal aerospike nozzles with the same design values, by numerical analysis. The results clearly indicate that the aerospike nozzle is capable of generating as much as 5% more thrust in over-expansion conditions. This amount of excess thrust is particularly valuable for a launch vehicle, where it can contribute to a remarkable advantage in the vehicle’s mass performance. As the base pressure thrust compensates the loss of thrust in under-expansion conditions, plug truncation has minor effect on the loss of thrust in these conditions. But in over-expansion conditions, thrust loss will increase with the increase of truncation. Base pressure thrust is closely related to variation of base pressure with atmospheric pressure. Base pressure is constant in under-expansion conditions, but increases with the increase of the atmospheric pressure in over-expansion conditions. Adding curvature to the base of a truncated plug does not influence its thrust, but significantly improves local temperature distribution at the base of the plug. Based on the observed behavior of the thrust delivered by aerospike nozzles with different amounts of plug truncation, it can be concluded that selection of the amount of plug truncation depends on the flight regime of the vehicle which will use this propulsion system. If most of the thrusted flight phase is going to be at altitudes lower than the design altitude (overexpansion conditions) it is recommended to use lower values of truncation, but if most of the thrusted flight phase is going to be at altitudes higher than the design altitude (underexpansion conditions) it is recommended to use higher values of truncation References 1. Hearth, D. P. and G. C. Gorton, “Investigation of Thrust and Drag Characteristics of a Plug-Type Exhaust Nozzle”, NACA RM E53L16, 1954. 2. Berman, K. and F. W. Crimp Jr., “Performance of Plug-Type Rocket Exhaust Nozzle”, ARS Journal, Jan. 1961. 3. Rao, G.V.R., “Spike Nozzle Contour for Optimum Thrust”, Ballistic Missile and Space Technology, Vol. 2, C. W. Morrow (Ed.), Pergamon Press, New York, 1961, pp. 92-101. 4. Lee, C. and Thompson D., “FORTRAN Program for Plug Nozzle Design”, NASA TM X-53019, 1964. 5. Korte, J., J. Dunn, A. Salas, N. Alexandrov, W. Follett, G. Orient and A. Hadid, “Multidisciplinary Approach to Linear Aerospike Nozzle Optimization”, 33rd Joint Propulsion Conference, Seattle, USA, July 1997. 6. Hagemann, G., H. Immich, T. V. Nguyen and D.E. Dumnov, “Advanced Rocket Nozzles”, Journal of Propulsion and Power, Vol 14, No 5, pp. 620-634, AIAA, 1998. 7. Hagemann, G., A. Preuss and H. Immich, “Advanced Nozzle Concepts for Future Rocket Engine Applications”, CNES-AAAF-ESA Conference, Liege, Belgium, Dec.2002. 8. Tomita, T., Tamura H. and Takahashi M., “An Experimental Evaluation of Plug Nozzle Flow Fields”, AIAA Paper 96-2632, 1996. 9. Sakamoto, H., Takahashi M., Sasaki, M., Tomita, T., Kusaka K. and Tamura H., “An Experimental Study on a 14KN Linear Aerospike Nozzle Combustor”, AIAA Paper 992761, 1999. 10. Ito, T., K. Fujii and A.Koich-Hayashi, “Computations of the Axisymmetric Plug Nozzle Flow Fields”, 17th AIAA Applied Aerodynamics Conference, Norfolk, USA, June 1999. 11. Besnard, E., H. H. Chen, T. Mueller and J. Garvey, “Design, Manufacturing and Test of a Plug Nozzle Rocket Engine”, AIAA Paper 2002-4038, 2002. 12. Anon., “New Nozzle Shows Potential for Increased Efficiency”, Aviation Week and Space Technology, March 31, 2003. 13. Ito, T., and Kozo Fujii, “Flow Field Analysis of the Base Region of Axisymmetric Aerospike Nozzle”, AIAA-2001-1051, 2001. 14. Naghib-Lahouti, A., Nazarinia, M., and Tolouei, E., “Design and CFD Analysis of an Aerospike Nozzle to Compare Its Off-Design Performance with a Conventional Nozzle”, Proceedings of IMEC2004, International Mechanical Engineering Conference, IMEC2004-FM014-CP, December 5-8, 2004, Kuwait. 15. James E. A. John, Gas Dynamics, 2nd Edition, Prentice-Hall, 1990. 16. Angelino G., “Approximation Method for Plug Nozzle Design”, AIAA Journal, Vol. 2, No. 10, Oct. 1964, pp. 1834-1835. 17. A. Naghib-Lahouti, M. Nazarinia and E. Tolouei, “Design, numerical flow analysis and investigation of the effects of geometric parameters on performance of an aerospike nozzle in off-design conditions”, Technical Report No. ARI-83-21-ASN1-4-2-1, Aerospace Research Institute, Tehran, Iran, September 2004. 18. Anderson J. D., “Modern Compressible Flow: With Historical Prospective”, 2nd Edition, McGraw-Hill Publishing Company, 1990. 19. Anon. “FLUENT 5.23 User’s Guide”, FLUENT Inc., Lebanon, NH, USA, 1999. View publication stats