You are on page 1of 3

NOVEMBER 1967

ENGINEERING NOTES

1549

of momentum and pressure thrust effects, respectively. Similar effects on performance as shown in Fig. 3 were found5 for various 7*3 from 1.1 to 1.4; aopt is not strongly affected by 7 for the ranges of variations typical of most propellant systems of interest. Decreasing p** causes a slight decrease in aopt and a slight increase in CF. To determine the validity of the one-dimensional source flow model employed in this analysis, several conical nozzles were analyzed with a two-dimensional method-of-characteristics computer program. Results were in very good agreement with those of Migdal and Landis8; the differences between CVs generally were <0.25%.
Appendix: Example of the Effects of Friction

out friction, thus decreasing the difference between them. By applying these displacement thicknesses to the friction 0pt, it was found that the corrected friction aopt now slightly exceeds the frictionless aopt for the shorter L and that the corrected friction aopt lies halfway between the friction and frictionless aopt's shown in Fig. 4 at the longer L. These results justify ignoring the boundary-layer effects for the purpose of estimating nozzle geometry in a preliminary analysis of a propulsion system.
References

Downloaded by BEIHANG UNIVERSITY (CNPIEC Xi'an Branch) on April 17, 2014 | http://arc.aiaa.org | DOI: 10.2514/3.29128

As stated in the analysis, the effects of friction on the optimum cone angle can be ignored for the purpose of estimating nozzle geometry in a preliminary analysis of a propulsion system. However, the effect of friction on CF is very significant and should not be ignored. To illustrate this point, the drag loss was added to the thrust equation in an approximate manner, and the necessary condition for maximum thrust was re-examined. The thrust equation including frictional effects is given by F = \[mVs + As(ps PQ)] rASUT cosa
2

(Al)

r = r(r) = ifpT (A2) where r is an average value of the shear stress based on r between the throat and the exit. For the purpose of this example, which is to determine what effect friction can have on a0pt and CF, f is assumed to be constant and equal to 0.0005. This choice of / was made to correlate with the ACp normally caused by friction and not from an experimental correlation of skin-friction data. Defining a CF from Eq. (Al) and setting dCp/da = 0 yields the following criterion for maximum thrust: 2f/Pc s na r r0 = X (1 + Gosa)(dg/da) (1 + cosa) (deg/da) 2esur cosa dr dr esursina] ST3T (1 + cosa) Pc dr der (A3) The results obtained by using Eq. (A3) instead of Eq. (8) are shown in Fig. 4. Curves "D" and "E" show that by including friction in the design equation, a slight shift in a0pt results. The difference in aopt calculated with and without friction increases with I/*, reaching values of the order of one degree at extreme L*. From Fig. 4, it is also seen that the frictional losses are indeed significant (the differences between curve "A" and curves "B" and "C," 1.63% @ L* = 26), and that there is actually a maximum value of Cp(L*) (curves "B" and "C") This CV max corresponds to the ojopt and L* determined by Fraser et al.4 but for the constant length constraint. It is not until L* becomes very large, and slightly before CF max, that a slight increase in CF (the difference between curves "B" and "C", 0.057% @ L* = 26) is detected by designing a conical nozzle with friction included in the design equation (curve "B") as compared with just simply subtracting the drag loss from the thrust of the nozzle designed without friction in the design equation (curve "C"). The displacement thickness of the boundary layer can vary from negative values for a highly cooled wall to rather large positive values for an adiabatic wall. The boundary-layer displacement thickness for several optimum conical nozzles with a nearly adiabatic wall (Te = 6000R, Tw = 5700R) was calculated by the authors using the boundary-layer computer program developed by Elliott, Bartz, and Silver.9 The growth of the displacement thickness for nozzles with L's of 9, 15.5, and 26 in. between the throat (yt = 1.0 in.) and the exit was found to be 0.05, 0.091, and 0.143 in., respectively. These values of the displacement thickness shift the aopt obtained including friction towards the aopt obtained with-

Barrere, M., Jaumotte, A., Fraeijs de Veubeke, B., and Vandenkerckhove, J., Rocket Propulsion (Elsevier Publishing Co., Amsterdam, 1960), p. 90. 2 Rao, G. V. R., "Recent developments in rocket nozzle configurations," ARS J. 31, 1488-1494 (1961). 3 Ahlberg, J. H., Hamilton, S., Migdal, D., and Nilson, E. N., "Truncated perfect nozzles in optimum nozzle design," ARS J. 31,614-620(1961). 4 Fraser, R. P., Rowe, P. N., and Coulter, M. O., "Efficiency of supersonic nozzles for rockets and some unusual designs," Proc. Inst. Mech. Engrs. (London) 171, 553-580(1957). 5 Scofield, M. P., and Hoffman, J. D., "An analysis of optimized conical thrust nozzles," Jet Propulsion Center, Purdue Univ., TM-66-6 (July 1966). 6 Malina, F. J., "Characteristics of the rocket motor unit based on the theory of perfect gases," J. Franklin Inst. 230, 433-454(1940). 7 Landsbaum, E. M., "Thrust of a conical nozzle," ARS J. 29, 212-213(1959). 8 Migdal, D. and Landis, F., "Characteristics of conical supersonic nozzles," ARS J. 32, 1898-1901 (1962). 9 Elliott, D. G., Bartz, D. R., and Silver, S., "Calculation of turbulent boundary-layer growth and heat transfer in axisymmetric nozzles," Jet Propulsion Lab., California Institute of Technology, Rept. 32-387 (February 1963).

Internal Ballistics of the Solid-Propellant Rocket Motor


V. K. SRIVASTAVA* Indian Institute of Technology, Kanpur, India
Nomenclature = burning-rate constant (r = apn) C = mass of solid propellant = D = initial web size = / fraction of the web size remaining unburnt at time t = semiempirical factor9 A; = NC = gas mass present in the chamber n = pressure index P = chamber pressure Q = rate at which gas escapes rate of burning R = gas constant S = burning surface area of a charge t = time T = chamber temperature U = empty chamber volume W = weight of the rocket z = fraction of the charge mass burnt up to time t e = form factor [Eq. (3)] 7 = ratio of the specific heats propellant density 5 in the chamber P = gas density covolume9
a1
7 1/2 [2/(7

Received August 30, 1966; revision received July 10, 1967. The author thanks J. N. Kapur for his help and guidance, and the Director R&D Organization, Directorate of Research (Laboratories), New Delhi, for financial support. [4.24] * Department of Mathematics.

1550

J. SPACECRAFT

VOL. 4, NO. 11

\ = RT at constant T ( ' ) , ( " ) = d( )/<ft and dz( )/dt2, respectively


Subscripts

N =z Pressure-Time Relation I

(8)

0 B

= conditions at t = 0 = at all burnt


Introduction

From Eqs. (2) and (8), we have

N = -f - rSop/CX1/2
Again from Eqs. (7) and (9), we get

(9)

Downloaded by BEIHANG UNIVERSITY (CNPIEC Xi'an Branch) on April 17, 2014 | http://arc.aiaa.org | DOI: 10.2514/3.29128

ERSHNER1 developed the theory of the internal ballistics of a solid-propellant rocket under the following assumptions: 1) constant burning surface, 2) uniform chamber pressure at any instant, 3) constant chamber temperature (isothermal model), 4) covolume 77 per unit mass is zero, and 5) change of volume available to the gases in the rocket motor due to the burning of the propellant is neglected. Assumption 1 implies that 6 = 0 (i.e., tubular charge), and assumption 2 implies that 1 + (kCN/QW) is unity, since the space mean pressure distribution is p/[l + (kCN/6W) ]. In rockets the pressure distribution is approximately parabolic and is important when discussing "regression" and "erosive" burning. Here, however, pressure drop is neglected. Assumption 3 implies that T is taken as constant, 4 implies that 77 is taken as zero, and 5 implies that CZ/5 CNrj is neglected. Price2 obtained the pressure-time curve p(t) in terms of charge and motor shape parameters. Gupta and Mehta3 also obtained p(t) when the change in volume is taken into account, i.e., when only the first four of the foregoing assumptions are made. Kapur4>5 used the basic equations of internal ballistics [for the general linear law of burning as derived from recoilless (RCL) guns6 for the nonisothermal model to determine p(t), which is a correction to the Von Karman and Malina equation.7 Srivastava8 has also established p(t) for the pressure index law of burning for a nonisothermal model. In the present note, we establish p(t) in the form of a second-order, nonlinear, differential equation. Particular cases 1) 77 = 0 and 2) 77 = 1/5 have been discussed. The law of burning is the pressure index (r = apn). The system of equations which we have considered is deduced from those of the RCL gun.6 Another case is discussed which includes all of the assumptions made by Kershner, but is solved for quadratic form-function only. An exact solution has been obtained for n = 0.5. If 6 = 0, the equations can be solved analytically for every value of n.

p(U - (Cf/S)

- CNvj] = CN\

(10)

Differentiating Eq. (10) twice, and using Eqs. (2) and (9), we get

[(77 -

l)pn

or [(77 -

[(77 - 1/5) (n

p2

(11)

Equation (11) is the basic second-order, nonlinear differential equation for this system. At 77 = 0, Eq. (11) becomes

[\p -

[\np-~1 - \t - (n + 2)p/5]p*

(12)

A better approximation is, however, obtained by putting 77^ = 1. ThenEq. (11) becomes
[\p *
(13)

Each of Eqs. (11-13) can be integrated to give p ( t ) . The transient exhaust curve is given by Eq. (93) of Kapur10 :

- log{(X + 77?)/(X + rjpB)}] +

1/(X + npB)} = -bCT(t - ts)/(DU)


To find the web thickness, we integrate Eq. (2) :

(14)

Basic Equations

The general equations for internal ballistics are Equation of state:

The integration will depend on the p(t) we use.


Pressure-Time Relation1II (General Form- Function)

p[U - (C/S)(I - Z) - CNy] = NCRT[l + (kCN/QW)] (1)


Rate of burning equation:
r = -Df = apn
Form-function: (2)

Here we consider the assumptions 2-5, but instead of constant burning surface, we consider the charge with quadratic form-function. The equations are

p[U - (C/S)] = CN\

(15)

Z = (1 -/)(! + Of)
Rate of nozzle flow:

(3)

and Eqs. (2, 3, and 9). Differentiating Eq. (15) and using Eqs. (2, 3, and 9), we get

N = Z - TSQp/C(RT)
Energy equation10:

(U - C/S)p = C\(Zf - rSop/CX1/2)


(4)

(16)

Let J70 = U - ((7/6), and B = C\a/DU0', then

d(NT)/dt = T<>Z -

(5)

p/B = -(-I + 0 - 20f)p - *p


If 0 = 0, then

(17)

Because of assumption 3, Eqs. (4) and (5) become identical. Hence, including assumptions 1-3, Eqs. (1, 3, and 4) reduced to

p/B = pn - $P
The solution of this equation is

(18)

p[U - (C/S) (I - Z) - CNrj] = CN\


Z = 1- /

(6)
(7)

pi-n = ^-i + Ae~*Bt

(19)

NOVEMBER 1967

ENGINEERING NOTES

1551

where A is the constant of integration. The initial conditions of integration are t = 0, / = /0, p = p0; hence A = pQl~n \l/~l} and Eq. (19) becomes

Turbulent Separation Associated with Axisymmetric Flared Bodies


THEODORE J. GOLDBERG* NASA Langley Research Center, Hampton, Va.
Nomenclature

which is the solution of the differential equation (18). Rewriting Eq. (17),

p/Bpn + \l/pl'n = 1 - 0 + 20/

Differentiating, we get
Let Y = pl~n, then Eq. (21) transforms to

y/(l - n) + $BY + 20aBY'l-*/D = 0


Downloaded by BEIHANG UNIVERSITY (CNPIEC Xi'an Branch) on April 17, 2014 | http://arc.aiaa.org | DOI: 10.2514/3.29128

(22)

Equation (22) is a second-order, nonlinear differential equation and is solvable for n = 0.5. The conditions for the integration of Eq. (22) are p = p0, i.e., F = 0, t = 0, / = /0, and, from Eq. (19),

p = po

p-o =

- ^po

(23)

At n = 0.5, Eq. (23) becomes

F + %B$Y + daBY/D = 0
or
y = Cie[
(24)

= first peak or plateau pressure coefficient for turbulent or laminar flow, respectively d, I = diameter and length of cylinder midbody section, respectively, cm Moo, Mo = Mach number at freestream and undisturbed conditions ahead of pressure rise at outer edge of boundary layer, respectively PI, Pt,& = pressure, local and freestream stagnation, respectively, Newton/m2 Rm = unit freestream Reynolds number, cm"1 RL,O, RX,O = Reynolds numbers based on undisturbed conditions at outer edge of boundary layer and axial distance from either leading edge to flare junction or from leading edge to point ahead of pressure rise, respectively s = surface distance from cylinder-flare junction, positive toward rear, cm a = angle of attack, deg

CP,p

which gives the desired pressure- time relation.


Conclusion

Equations (11-13, 20, and 24) give the required p(t) for an ideal solid-propellant rocket motor. Equations (11-13) can be integrated numerically for the parameters of the propellant used in the rocket, whereas Eqs. (20) and (24) give the pressure- time relation by simply substituting the data for actual rockets. The most significant result is that exact solutions are possible for n = 0.5 for all charges shapes and for 6 for all values of n.
References

Kershner, "Rocket fundamentals," U.S. Dept. of Defense, National Defense Research Committee, Div. 3, Sec. H, Office of Scientific Research and Development Kept. 3711, Chap. Ill (1944). 2 Price, E. W., "Charge geometry and ballistic parameters for solid propellant rocket motors," Jet Propulsion 24, 16 (1954). 3 Gupta, S. K. and Mehta, A. K., " Variation of pressure with time in rocket motion," Proc. Natl. Inst. Sci. India, Sec. A: 20, 653 (1954). 4 Kapur, J. N., "Pressure-time curve in internal ballistics of solid fuel rockets as deduced from the theory of internal ballistics of RCL gun," Proc. Natl. Inst. Sci. India, Sec. A: 23, 150 (1957a). 5 Kapur, J. N., "Pressure-temperature curve for the ideal solid-propellant rocket motor," Proc. Natl. Inst. Sci. India, Sec. A: 28, 541 (1962). 6 Kapur, J. N., "Internal ballistics of R.C.L. Guns," Proc. Natl. Inst. Sci. India, Sec. A: 23, 541 (1957b). 7 Von Karmdn, T. and Malina, F. J., Collected Works of T. Von Kdrmdn (Butterworths Scientific Publications Ltd., London, England, 1956), Vol. IV, pp. 94-106. 8 Srivastava, V. K., "Pressure-temperature curve for the ideal solid propellant rocket motor II," Mem. Art. France 4, 923-928 (1966). 9 Corner, J., Theory of the Interior Ballistics of Guns (John Wiley & Sons Inc., New York, 1950). 10 Kapur, J. N., "Thesis on internal ballistics of guns and rockets for single composite and moderated charges," University of Delhi, Delhi, India (1958).

LARED afterbodies can provide stability at high Mach numbers without markedly increasing the weight of a body, and they provide a natural fairing between various stages of launch configurations. However, predictions of aerodynamic characteristics of such bodies are not routinely successful. To calculate these characteristics it is required that the flow on and ahead of the flare surface be analyzed in terms of local conditions. Therefore, it is necessary to have information on separation phenomena as a function of local Mach number and flare angle. Kuehn1 has obtained data with turbulent boundary-layer separation induced by flares on cylinders at a. = 0 from Mach numbers of 1.5 to 4.5. However, data available on hypersonic turbulent boundary-layer separation are limited. The present paper provides additional information on maximum flare angles for attached flow and pressures in turbulent separated regions on nose-cylinder-flare bodies at M = 6 and at Rm/cm from 0.19 to 0.33 X 106. Some data were obtained at angles of attack up to 12. While separation is affected by many parameters, one of the important parameters in flare stabilization studies is the maximum flare angle through which the flow can turn without separating. In general, the criterion used to detect the presence of separation in this investigation was the first appearance of a "hump" in the longitudinal pressure distributions as used by Kuehn.2 An example of the surface pressure distributions obtained at a = 0 with a 15 conical nose-cylinder configuration with various conical flare angles is presented in Fig. 1. Although the flare length is relatively short and not truly representative of a practical configuration, the pressures on the flares reached the inviscid conical-flow values except for those with large separation. Therefore, the maximum flare angles for attached flow determined from these data appear applicable to longer length flares. It should be noted that the pressure rise measured in a separation region is not necessarily that which will cause separation. V The maximum flare angles for attached flow at a = 0 as a function of local surface Mach number are presented in Fig. 2. These values were obtained at near adiabatic wall conditions; however, limited data obtained by Kuehn1 indicate
Presented at AIAA Sounding Rocket Vehicle Technology Specialist Conference, Williamsburg, Va., February 27-March 2, 1967 (no paper number; published in bound volume of papers of the conference); submitted March 9, 1967; revision received August 10,1967. [11.14,11.16] * Aerospace Engineer, Applied Fluid Mechanics Section, Aero-Physics Division.

You might also like