You are on page 1of 20
Mechanics of Particles, Bodies -and Fluids ee Up:to the present the velocities encountered in astronautics are relatively small as compared to the velocity of light, and so classical or Newtonian chanics will suffice our purpose. Newtonian mechanics is based on the ; three:laws of Newton, Galileo’s principle of relativity and the three classical conservation laws: conservation of mass, linear momentum and energy. Moreover, the assumption is made that forces behave as vectors and can be tepresented mathematically as such. We will now state Newton’s laws for particles, where a particle is defined as a constant mass concentrated at a point. 1. Every particle continues in its state of uniform motion in a straight line (or rest), unless compelled to change that state by forces acting upon it. ~ 2. The time rate of change of linear momentum of a particle is propor- tional to the force acting upon that particle and is collinear with and in the direction of that force. _ /3, The mutual forces of two particles acting upon each other are equal in magnitude and opposite in direction (action = reaction). 3.1. Newton's first law In analyzing the first law, we must realize that it is a statement about ‘ motion, which is a relative concept. If we speak about motion of an object, we have. to specify a frame of reference relative to which the object is pe eeuing, In the-first law no reference frame is specified. We know from experi- -ence that we cannot just choose any reference frame, i.e. there are special J Telerence frames wherein the first law is valid. We can in fact turn the first law around and make the following statement: There exist certain reference frames with respect to which the motion of a le; free of all external forces, is uniform (motion in a straight line with tant speed, including zero). ‘Those frames are called inertial frames. The first law then becomes in fact hi ‘inition of an inertial frame of reference. § a direct consequence of this definition that if such a frame exists, a class of inertial frames can be found. To show this, let us suppose that frame S$ is an inertial frame. If another frame S' translates uniformly, 26 Rocket Propulsion & Spaceflight Dynamics Mechanics of Particles, Bodies and Fluids 27 with velocity V, (i.e. V is constant), with respect to S and if the time in differs by only a constant from the time in S, then S' is an inertial frame tog| We will now proceed to prove this statement, known as Galileo’s principle of- relativity. q The transformation of space and time coordinates from S to S' is given bj the following relations: 3.2. Newton's second law The second law, like the first law, deals with motion and again from experience it is found that the second law holds only with respect to inertial frames. The second law can be expressed as at (3.2-1) r'=r-Vi-R, F=7 (mv). pe Here mV is the linear momentum of a particle with mass m and velocity V. The mass in this law is the so-called inertial mass, in contrast with the gravitational mass which will appear in Newton’s law of gravitation later on. Newton, however, tacitly assumed the equivalence of inertial and gravita- tional mass. Experiments by Eétvés et al. [1] in the beginning of this century, and- more recently by Roll et al. [2] and Braginsky and Panov [3] have established the constancy of the ratio of inertial to gravitational mass within 1 part in 10’? or better. In Einstein’s general theory of relativity the equivalence between inertial and gravitational mass is a consequence of the principle of equivalence [4, 5]. We have seen that from the Galilean principle of relativity and from the . first law it followed that systems translating uniformly relative to an inertial | system are inertial systems themselves. We can only use this statement if the second law is invariant under the Galilean transformation. This invariance - exists indeed if mass and force are invariant and if, moreover, the mass of the particle is a constant, as will be shown. Consider therefore again the Systems $ and S' as described before. If F is the force on the particle P in S, and F’ is the force on P in S$", then since the force is invariant, F= F’. It follows from Eqs. (3.1-1) that the time rate of change of linear momentum of the particle in the two reference systems is related by a dry dy fdr’, \)_d/_,dr'\, Vd’ a(S) "ar |" (G+ ¥) Jae (m ae)*¥ ar” Applying Newton's second law in both systems, we obtain: d dr An syste; si¥a5( 3) osm a ae) idr'\ dm’ (mi at¥ ar” We see that the second law is invariant only if the mass of the particle is a constant. This transformation of coordinates is called the Galilean transformation Herein, r and r’ are the position vectors of a particle P with respect to origins of S and S’, respectively. R is a constant vector defining the positioi of S' relative to S at time 1=0; 1 and 1’ are respectively the times in S ani S', while T, finally, is the constant time difference between S and S’. If nt forces are acting on the particle P, then according to the first law the veloci of the particle relative to S is a constant dr vo dt? The velocity of the particle relative to S’ is defined as Using Eqs. (3.1-1) we find vind tar ot yyy, dt dt! dt dt 7 Since both V, and V are constant, we conclude that V’ is a constant to According to the definition of inertial frames, S' and thus any frame whi translates uniformly with respect to S is an inertial frame. As we will see later on, this is not true for reference frames which rotatt f whether or not uniformly, relative to an inertial frame. Any such rotation motion will produce apparent accelerations. We therefore have to concludj that Newtonian mechanics requires an absolute reference for rotation. motion. Translational motion on the other hand is relative in that a arbitrary uniform translation may be superimposed. Experience has prové that for all motions we can find a frame that behaves as an inertial one, Jeast within our measuring accuracy. To describe the motion of the plane! in our solar system, the heliocentric frame that is not rotating relative to th fixed stars can be considered as an inertial frame. Likewise, as for exper} ments within a laboratory on Earth a frame fixed to the laboratory can viewed as an inertial one. We therefore define an inertial frame to be one which, experimentally, the first law holds within our desired accuracy. ‘Tt was for that Teason that we stated the laws of motion for particles with _ COhStant mass, The third law, finally, is clear in itself. However, we have to add the quitement that the action and reaction forces are collinear. This collinear- ity'is essential for the conservation of mechahtical angular momentum of an 28 Rocket Propulsion & Spaceflight Dynamics Mechanics of Particles, Bodies and Fluids 29 isolated system. It is true for all Newtonian gravitational and mechanic interaction forces but does not apply to certain forces between movi book. 3.3. Non-inertial frames We have seen that the first and second of Newton’s laws are valid only inertial frames. This does not mean that studying the motion of an object a non-inertial frame prohibits us from using Newton’s laws. We will have te modify these laws in such a way that they are applicable in non-inertil frames too. In this section we will derive the laws for the said modification. | me Suppose S is an inertial frame and S' is a non-inertial one. We know bjE now that if S' is non-inertial it is moving with respect to $ with a velocitjy which is not uniformly translational. According to Chasles’ theorem [6], thi most general motion of S’ with respect to S$ consists of a translational motion and a rotational motion. By a translational motion of S’ is meaalé that in the infinitesimal time interval At every point fixed with respect to Sf undergoes a displacement Ar relative to S, where Ar is the same for a points of S’. The translational velocity is defined as . Ar Vor Tim ‘Fig. 3.1 Rotation of vector r fixed with respect to S’ relative to reference ‘frame S proportional to the vector rx@ and we can write rx Ar=_—— rsi . r jrxg] ine 48 and |rQ|=rQsing. with respect to S' undergoes an angular displacement A@ in the time interval At, where AQ is measured about a line fixed in S: the axis of rotation. Wig. now define a vector A@ with magnitude A@ and directed along this axis df: rotation with a sense according to the right-hand rule. The angular veloci OQ of S’ is then defined as Ar=rxQ At e finally get the velocity of P by dividing Ar by At and taking the limit for 0. So the velocity of P due to the rotation of S' relative to S is With this result we can derive a formula which relates the time derivative of “4 -Vector in a rotating frame to the time derivative of that vector in a Totating frame. Again, let S’ rotate with angular velocity © relative to Ife, ¢, and.e, form a Cartesian unit-vector triad in S', then any vector A be expressed as Ae i =tim 48 33 Olin 5: B34] 3.3.1 Rotation In order to concentrate a moment on rotation, suppose the origin of systes| S' coincides with the origin of S. Suppose $’ is in pure rotational moti relative to $ about an axis through the origins of both systems. If P i: point fixed in S’, with position vector r with respect to the origin of S’, th the point P describes in the time interval At part of a circle (Fig. 3.1) willg radius r sin g and lying in a plane perpendicular to the axis of rotation. Heft @ is the angle between the vector r and the axis of rotation. Neglecti dA, dA, de, de, de, Se 4 a L& oad z e ar e+ de e+A, dt + Ay dt +A, at” second order terms, the point P undergoes a displacement Ar perpendicul “ e+fAre 494s, * dt dt 7 to r and the axis of rotation and with magnitude Ar=rsin g A@. So Ar} i Vp=Qxr. (3.3-3) | esterase 30. Rocket Propulsion & Spaceflight Dynamics 5A 7 then or is the time rate of change of the vector A as viewed by an observ in the rotating system S'. For, only the components of A can change fork such an observer since the unit vectors e,, e, and e, are fixed for him. For ar} ; observer in the inertial system Sy & is the velocity of the point with position vector e, due to the rotation 2. Taseordind to Eg. (3.3-3) this velocity js ist 2 xe,. Likewise: dey at 2*4 and ap Xe So the rate of change of A for an observer in the inertial system is dA 6A 4 Gr O*A (3.3-4) Though we have assumed that S is an inertial frame, it is important tof note that nowhere have we actually used this fact. Therefore Eq. (3.3-4) generally holds for two systems rotating relative to each other. If, foi} example, a system S' is rotating with respect to a system S” with angulare velocity Qs as viewed from S$”, then for any vector A (2), (rane Bt) \ St) 5 G35, Likewise 6A 3A A ==) =(—) +0. 3-6) i). (Fr). Oe xA, G34 where Qg» is the rotation rate of 8” as viewed by an observer in S’. AS Qg.=—Qs,, both expressions are alike. 3.3.2 Newton's laws We will now turn to the problem how to modify Newton’s laws such thal} they are valid in non-inertial frames. Let S’ be a system which translates and! rotates relative to S (Fig. 3.2). If R is the position vector of the origin of SH with respect to the origin of S and r and p the position vectors of a particle} P with mass m relative to the origins of $ and S', respectively, then | r=R+p. (33-9) By differentiating Eq. (3.3-7) we obtain the absolute velocity of P _dr_dR dp ae ar dt dt” All derivatives in Eq. (3.3-8) are rates of change as viewed fro the inertial [ (3. 3-8) Mechanics of Particles, Bodies and Fluids 31 Fig. 3.2 The position vectors of a particle, P, with respect to a fixed frame, S, anda moving frame, S’ | frame. Applying Eq. (33-4) for the vector p we can rewrite Eq. (3.3-8) as e a +o (3.3-9) Vans dr R Oxp In order to explain the different terms in Eq. (3.3-9), let P’ be a point fixed Re oxp is the velocity of P’ . : 5p. telative to S due to the translation and rotation of S'. The term 3 is the in S' coinciding with P at the instant 4 Then velocity of P relative to P’ for an observer in S’ as well as for an observer in 5. This velocity is called the relative velocity of P, Vier Ifno force is applied to P then, according to Newton’s first law, V, “aps has to be a constant. The velocity of P in S’ is given by V, fret 7 aR = Vass ar OQXp. This velocity can only be a constant for any point P in $ if and only if 2=0 a at and “a =constant, Indeed we see that only those systems translating uni- formly with respect to S are inertial systems. » By differentiating Eq. (3.3-9) we get the absolute acceleration of P a@R dQ de, +5 (2), ae a Oar ar ar Fabs Again using Eq. (3.3-4), we obtain 2 ptr (xp) +20x 2422. (3.3-10) ét bt? ‘ ) 32 Rocket Propulsion & Spaceflight Dynamics The sum of the first three terms of the right-hand side of Eq. (3.3-16) represents the acceleration of the fixed point P’ for an observer in S. This q 2 called the dragging acceleration, ay. The term “3.5 is the absolute acceler tion of the origin of S'. The term x (@ xp) is due to the rotation of S’ ani dQ represents a centripetal acceleration. The term ——%p, often called of dt tangential acceleration is due to the rate of change of rotational velocity The sum of the last two terms in Eq. (3.3-10) 7 the acceleration of #} relative to P' for an observer in S. The term 22 «ie is called the Cori acceleration, a,. The term ee an observer in S' and is called the relative acceleration, a,.. So, in general, we can write finally, is the acceleration of P as viewed bj Agus = Ag, +A, + Are If F is the actual force applied to the particle P, then according to Newtons second law F= ma,y, = may,+ ma, + ma, - We now define two apparent forces: a dragging force | Fy, =—ma,, (3.3-11} and a Coriolis force E= —ma,. Thus we can write the second law in the form ln BFR, +E tg, 3-0 According to Eq. (3.3-13), we can apply Newton’s second law in not inertial frames if for the total force acting on the particle we take the sum di the external force and the two apparent forces. 3.4 Dynamics of particle systems With the laws and formulae stated and derived in the previous sections, tht} motion of a single particle under influence of a known force can bé calculated in a fairly straightforward way. By applying Newton’s second I a second-order vector differential equation is obtained. This vector equati can be split into 3 scalar second-order differential equations for motion if physical space. These differential equations, together with appropriate initial conditions, have to be solved, either analytically or numerically. In this wal] one gets velocity and position as a function of time. In Section 3.6 we wilh. consider the motion of a particle in a gravitational force field, as this analys Mechanics of Particles, Bodies and Fluids 33 will be very useful in the treatment of satellite orbits and ballistic trajec- ‘tories. ‘First, we will consider the dynamics of a group of interacting particles, uwhere we are interested only in the over-all aspects of the motion and not in “the motion of each individual particle. The formulae as derived in this section are particularly useful for the analysis of the motion of continuous particle systems, ic. bodies, whether rigid or not. Although fluids can be treated as particle systems too, it is customary to treat fluids using a different approach (Section 3.7). 3.4.1 Systems of discrete particles Consider a mass-system S$ consisting of N particles P,, each with a constant mass m, and position vector r, with respect to the origin O of an inertial frame XYZ, as shown in Fig. 3.3. In the most general case the system will have a variable mass which is expressed by N being a function of time. ‘The forces as applied to P, consist of an external force, F,, arising from sources external to the system S, and internal forces due to interactions among the particles. If F, is the internal force acting on P, due to the particle P,, we know from Newton’s third law Fy. = Fs (3.4-1) We have already pointed out that we will suppose F;, and F,; to be collinear, ie. they act along the straight line connecting the particles. In agreement with Eq. (3.4-1), F,=0, which means that particle P, cannot exert a force on itself. Fig. 3.3 The particle system, S, with position vectors of, and forces on the Particles P, and Py 34 Rocket Propulsion & Spaceflight Dynamics Mechanics of Particles, Bodies and Fluids 35 The linear momentum of P, relative to XYZ is ' of. mass and were driven by an external force equal to the forces external to dr, “BE the system. J,=m,—. ar (3.4-2)—- The equation of rotational motion of the system S$ can be obtained by ectoral multiplication of Eq. (3.4-3) with r,: Applying Newton’s second law on P, we get ad; x ay, XE + Sonemennx (3.4-8) " ee the moment of F; with respect to O is defined as Now summing Eq. (3.43) over all N particles which form the system S af =1XF, time f, we obtain si the Sadan momentum of P, is X a, dard y Re yo. 6.44 se nXJ,= ma, «Zt (3.4-9) Now let z F,=F,: the total external force applied to the system S at time fF “Then, the right-harid side of Eq. (3.4-8) can be written as From Eq. (3.4-1) we see that So Eq. (3.4-4) can be written in the form (3.4-10) Re Xap 8 dy : SQ Ge 2 ae . By summing Eq. (3.4-10) over alll particles N, we obtain This equation is known as the translational equation of system S. t M+ y 5 1X Fy = yo (3.4-11) For an invariable system, i.e. a system where no particles enter or leave the system, the translational equation can be written in a simpler and mort] familiar form. In that case, summation and differentiation can be intet changed, leading to sees Emegute as (2, mr) gp Mn = MSS isLke1 Now i M, = Ms: the total moment about O due to the external forces on fhe system S. Because of Eq. (3.4-1) and the collinearity of the internal orces “oN N YY xR =0, is KE where M is the total (constant) mass of the system, given by M= x m, anf as can be seen by taking the terms in pairs together Tom is the aval vector of the center of mass of the system, defined as + 1X Ei + 1, X Fy = (1) — 1) X Fy, = Me X Fy, = 0. at rit ; G. Finally, the rotational equation for the system is found to be Mit ; X aB, @r, My= Y Gig is & i ZL mir, (3.4-12) Again, for an invariable system this equation can be written in a more ‘amiliar form by defining the total angular momentum relative to O as B=Y mr, x2, Sm a The translational equation can be written as Fem B=Mae". (3.4 This equation states that the motion of the center of mass of the system if - the same as if the entire mass of the system were concentrated at the centel 36 Rocket Propulsion & Spaceflight Dynamics Mechanics of Particles, Bodies and Fluids 37 and as summation and differentiation can be interchanged, we get for such] equation is the general translational equation for an arbitrary body with system ible mass. ian ‘The rotational equation as given in Eq. (3.4-15) can be simplified by o,=25. g:the center of mass of the system, instead of a fixed point in inertial at "as reference point for calculating the external moment and angular 3.4.2 Bodies entum. If ro and Po are the location vectors of the point of application total external force with respect to O and the center of mass, respectively, then Mg = 10% Fs = (Tem + Po) X Es otal moment due to the external force relative to the center of mass is Mon = PX Fs = Mg — Trem X Es. After having derived the equations of motion for mass systems consisting: a finite number of discrete particles we now can quite easily give equations for continuous mass systems, i.e. for bodies. They can be thou; of as consisting of an infinite number of particles with an infinitesimal m: In that case the summation signs in Eqs. (3.4-5) and (3.4-12) must replaced by integration sign, thus leading to: substituting the expressions for My and F, into the above equation and r= “am "Eq. (3.4-17) we obtain, as Tq is independent of the three dummy So a de” bles of the integration, M,=[ rx aM. (3.4-20) he one has to note that the integrals are in fact triple integrals foi three-dimensional system. For dM can be written as pd¥Y where p is. mass density-and dY a volume element. This volume integral will indicated by Sry. The Equations (3.4-14) and (3.4-15) can be written in a form more suil for our applications. If r,,, is the location of the center of mass of the sys defined by Man=[ ox (Sexo) am+ px{Qx (2x p)}dM+ 2 +2| ox (ax) ams [ px2 fam. (3.4-21) Ie bt hee at is the general equation of rotational motion for an arbitrary flexible body with variable mass. im| aM=[ rm, 3 7 7 [, a ¢ 4.3. Rigid bodies ; . 5p_ ap —<5=0 and the translational equation takes the and p is the position vector of a point P of the system with respect to y * 6t 8t? center of mass, then T= Tom +P. (3.4 For the absolute acceleration of P, we obtain by using Eq. (3.3-10) Pr _ rem, AD dp 8p 3 +x p+0x(O se oP 7 ae ae a xet x Xp)+ 20x +a, (3. (3.4-22) a rigid body the total angular momentum of the body relative to the ‘er of mass is, according to Eq. (3.3-9) where © is the rotational velocity of the system relative to XYZ. Substil B..=[ x(Qxp) dM. , (3.4-23) dM tion of this expression for o into the translational equation, Eq. (3.4- en fil and taking into account that fy, p dM =0, leads to en find dB. dQ & 8 2, = SB em ‘da _ R=M. "=+20%| ams | 0 au oa ht [© (Gexe) ames [ px{Qx(Qxp)}dM. (3.4-24) dr Im ot Ine Of? 38 Rocket Propulsion & Spaceflight Dynamics Evaluating the quadruple vector product of the second integral on t right-hand side of Eq. (3.4-24) and using Eq. (3.4-23), we obtain d Let us now take a reference system xyz, fixed to the body and with origin ( coinciding with the center of mass of that body. If the unit vectors in this Man | ox (xe) dM+QXxB,,, - Ie reference frame are e,, e, and e,, and if x, y, z are the coordinates of point} in this frame, then 4 p= xe, + ye, + ze, (3.4-264) and O= pe, +4e, + re, (3.4-260)f where p, q, r ate the components of © in this frame. We now can evaluate Eq. (3.4-23) as : 3 Bon= [ [e{(y?+27)p - xyq~ xzr}+ Ia teyf —yxp +(x*+27)q~ yzr}+ (3.4-21] te{ —-zxp - zyq+ (x?+y?)r}]dM. We now define: L=[ (?+27)dM, Ty= f (x?-+2?) dM, ha IM I= [ (2+y2)dM, — Iy=Ie= [ ~xy dM, (3.4-28 he ha These are the nine components of the inertia tensor I; I... Ijy and I,, at called the momenis of inertia, and I,,, inertia. Hence, we may write | Bom = €,Bz + eB, +¢,B., “Lye Ins, etc., are called the products df (4-28) : Mechanics of Particles, Bodies and Fluids 39 where B, = Lept Lyt Lat B,=Lyp+1,q+ 1.1, (3.4-30) B,= Lap t Inqt Let. Thus, Eq. (3.4-23) can also be written as 2. [ px(Qxp)dM=1-. (3.4-31) Im . This last equation is nothing but a formalization of Eq. (3.4-30) and it shows us that the dot product of the tensor I with the vector @ again is a vector. . dQ, Generalizing this result, and replacing & by awe can evaluate the first integral of Eq. (3.4-24), and obtain Mon =1-2r0xn.,. (3.4-32) Jf one resolves M,,, in Eq. (3.4-32) into its components M,, M,, M,, one can write the vector equation for rotational motion as three scalar equations: dp dq dr i M, Tax Gy t ey Gy tee gy t Be: TBy, dp, dq,, ar =r oP a 894 Ue pK 3.4-33 ly bgp t by apt beg t Be pB., ¢ ) dp, dq,. dr M, = 1," + Ly— + Le 7 Ty at ant at pB,~ qB,. i “Tt can be proved that it is always possible to find a coordinate system with its origin coinciding with the center of mass of the body such that the Products of inertia are zero simultaneously [7]. The three coordinate axes : for this case are known as the principal axes and the corresponding moments Of inertia are called the principal moments of inertia. In that case the Eqs. @.4-33) reduce to . dj M,= Le GP (se ly) ah M, = ty S44 (L,—LadPn (34-34) M, = f+ (ly ~ La) PA 40 Rocket Propulsion & Spaceflight Dynamics 4 Mechanics of Particles, Bodies and Fluids 41 These equations are known as Euler’s equations of motion. They are widel: -0 we obtain the so-call areal velocity: d dA = 1p de _ H dt dt 2° G6 dA. and we conclude that Gy 8 2 constant. So, the radius vector sweeps ou! equal areas in equal times, which verifies Kepler’s second law. A seconé| constant of motion can be obtained by scalar multiplication of Eq. (3.6 , ar, with a dr @’r__w, dr aed? ar or, id (dr dr wld d (wu sare a)” ~p2aee)= ale) So after integration, we obtain wt = € = constant. Attraction Fig. 3.4 The position of a particle movirig in an inverse-square force field times t and t+At Mechanics of Particles, Bodies and Fluids 47 2 a y' Equation (3.6-8) is the law of conservation of energy, since z is the kinetic energy per unit mass and, as derived in the foregoing section, - is the potential energy per unit mass. 3.6.2 Trajectory geometry $ the motion takes place in a plane, the vector equation (3.6-2) can be “written as two scalar second-order differential equations. We therefore use a feference system with the origin in the attraction center and with three mutually perpendicular axes: the r, @ and z-axis. The r-axis joins the “particle with the attraction center, the g-axis lies in the plane of motion, positive in the direction of the tangential velocity. The z-axis, finally, is normal to the plane of motion and is positive in the direction of the angular momentum. Owing to the tangential velocity of the particle this reference system is rotating with angular velocity o-Z en. We can write (3.6-9) d dr de sae +1Oxe =—e+r—e,. e+ 1OX 6, =F GFT ep By differentiating again a drdp, da 3) + + +e 7 Ox + ox 7 (aise "ae )% 7 fos or @r [dr fde @e,, drde Lots SP Stee 3.6-10 _ de (ae A) ke +r a? na ale ; ) ‘Substitution of Eqs. (3.6-9) and (3.6-10) into Eq. (3.6-2) leads to ar (dey? _ a(t) oe 7 Se, drdg_ "a errers dt nae ‘Equation (3.6-12) can be integrated directly after multiplication by r. The (3.6-11) (3.6-12) 7 d 7 : : Tesult is then eo =constant. This was already found, since according to 48 Rocket Propulsion & Spaceflight Dynamics d Eq. (3.6-6), a is the absolute value of the angular momentum, which is 4 constant of motion. Equation (3.6-11) can be used to determine the shape of the trajectory: # ie. r as a function of g. In order to solve Eq. (3.6-11) for r as a function off @ we replace t by @ using the relation d_dgd a drag (3.6-13): With the use of Eq. (3.6-6) we can write d_Hd a dg (3.6-14) In order to avoid mathematical difficulties it is useful to substitute 1 r=-, (3.6-15) u d/l du Hu? © (=)\)--ypS4 _ wee () aS (3.6-16 ar, du ae -Hu de® G.6-17) Substitution of these relations into Eq. (3.6-11) and dividing the resultin, equation by H7u? leads to the second-order linear differential equation Pu be tus, 6- ite (36-18 The general solution of this equation is u=+y+ Cos (—0) @6-19)) Ft Coos (eo), (36-19) | where C and w are integration constants which have to be determined from | the inital conditions. Let the initial conditions at ¢= fp be given by: : V(to)= Vo, (to) =o, (to) = ro and (to) = go. | Since the energy € is a constant of motion we can write 1] 7 + Mechanics of Particles, Bodies and Fluids 49 y using Eq. (3.6-16) and Eq. (3.6-6) we find 2 2. 2. du\? QZ Ve= Vit Vem EH? (Zo) + Hu, and the energy equation can be written as [eV -mee Substitution of Eq. (3.6-19) in this expression results in Ele “] SL \e-Hl-¢ 2 [e Ht [ from which equation we obtain C as (3.6-20) ‘We have chosen the positive root for C since any possible sign change can _. be handled by the choice of «. Substitution of Eq. (3.6-20) into Eq. (3.6-19) gives ; unde [r+(, 21) cos (o-«) | _ The constant w can be determined from the above equation together with Eq. (3.6-16) at #= tp. However, since the reference line from where 9 is measured can be chosen arbitrarily we make the substitution @ = g—w. The orbit equation then becomes r= — le (3.6-21) 1+ ( 8 +1) cos 6 B . We see that r is minimal when @=0, so 6 is measured from a line drawn \ from the center of attraction to the closest point of the trajectory, the -peticenter. The point farthest away is called the apocenter. Also the terms Periapsis and apoapsis are used. The angle 6 is called tue anomaly, while the angle , determining the Polar angle of the pericenter, is known as the arguménj.af pericenter. The Cquation of the orbit is the representation in polar coord.2tes of a so-called "onic section as we will prove now. In general, a conic section is the locus of Points with a constant ratio 5 (Fig. 3.5), where r is the distance to a given Point F, the focus, and d is the distance to a given line |, the directrix. The Onstant ratio is called the (numerical) eccentricity of the conic and is Genoted by ¢ From Fig. 3.5 we see that dy=d+rcos@ and since 50 Rocket Propulsion & Spaceflight Dynamics Mechanics of Particles, Bodies and Fluids Major Axis "Fig. 3.6 The ellipse Fig. 3.5 The general conic section . . “constant ratio Tv the ellipse can also be defined as the locus of points such _ that the sum of the distances from the two foci F and F’ is a constant. This constant is equal to the length of the line that connects the two vertices, the major axis, which is denoted by 2a. The minimum value of r, known as the pericenter distance is P r : Eaqne we can also write Bg e a. +rcos 0, and by solving for r we obtaiti 0 P — 3.6.2 TT re cos 0 6 ‘This is the general equation for a conic section. The radius at 0=90° nase, (3.6-25) called the semi-latus rectum and is denoted by p. Comparing Eqs. (3.6-21) te and (3.6-22), we see that they are the same and we conclude that th} while the maximum value of r, the apocenter distance, is given by trajectory of a particle in an inverse-square force field is a conic section with eccentricity and semi-latus rectum given by : EZ P. (3.6-26) = 2eF (3.6-23) i-major ax e= 7 7 | The semi-major axis, a, and the eccentricity, e, follow from H 1 a=h(r,+7,)=—2 =the pa (3.6-24)} 2(t +a) ine °7 7, ri (3.6-27) This statement is Kepler’s first law. The distance between the two foci F and F" is | 2f=2a-21, = Pe = 200. (3.6-28) 3.6.3 Classification and characteristics of conic sections ‘The semi-minor axis, b, then follows from b=Va?= f= avi-e?. (3.6-29) Parabola. For a given value of p we see from Eq. (3.6-27) that if e ‘pptoaches unity the semi-major axis approaches infinity. In the limit, if €=1, the conic is not a finite closed curve anymore and has become a Parabola. This also follows from Eq. (3.6-22), for, if e=1, r approaches Infinity for @—> 180°, From the general theory of conic sections, we know that conics are classified depending upon the value of the eccentricity. If 01 the conic is a hyperbola. We will now briefly discuss thos, characteristics of the various conic sections, which will be needed in subsequent chapters. Ellipse. (Fig. 3.6). Although defined here as the locus of points with 52 Rocket Propulsion & Spaceflight Dynamics The pericenter distance is related to the semi-latus rectum by Eq. (3.6 25), which for e=1 yields 1, =4p. (3.6-30) So, we can write the equation for a parabola as 2r, =_4h_ ~ 1¥cos 0" (3.6-31)) Hyperbola. If e>1, the denominator on the right-hand side of Eq. (3.6-22} changes sign for cos o=2, The result of. this is that the hyperbola has tw branches with asymptotes making an angle 6, = arcoss () with the linea joining the two vertices. This line is the major axis. In order to use the same || equations for all conics, the value of the semi-major axis is taken negative Jt f for a hyperbola. So, the Eqs. (3.6-25) to (3.6-27) remain valid for hyperbola, while the distance between the two foci is given by 2f= -2a+2r, = ott = —2ae. The semi-minor axis is imaginary for a hyperbola. (3.6-32) Returning to the orbit of the particle we can calculate the semi-major axis of the trajectory with Eqs. (3.6-23), (3.6-24) and (3.6-27) resulting in The semi-major axis thus is only dependent on the total energy per unit mass of the particle, ie. only a function of Vo and rp. The energy equatior (3.6-8) can now be written in the following form: (3.6-34) This is known as the vis-viva integral. We can conclude directly from Eq. (3.6-33) that the trajectory is a ellipse if $<0, a parabola if = 0 and a hyperbola if 8 > 0, which of cours! follows from Eq. (3.6-23) too. Expressed in the initial conditions Vo and 1 this means that if 2p <2 Yo< 5 the trajectory is an ellipse, the trajectory is a parabola, (3.6- 33) E Mechanics of Particles, Bodies and Fluids the trajectory is a hyperbola. : 0, increasing the speed will gradually change a closed conic into an open . The minimum speed that is required to escape permanently from the ge (3.6-35) To special case of an elliptic orbit occurs for e = 0. For such a circular orbit e find from Eq. (3.6-23) = Expressing $ and H in rp, Vo and Yo we obtain 2, (4 EytroVa cos 19)? +1= 0. wa ‘This equation can also be written as vayz sin? 9+ cos? ¥ (1-2%8) =0. “As both terms in this equation are pure squares, a requirement for a circular Orbit is yy=0 and Vo= Je . For the so-called circular velocity we a obtain V.= (3.6-36) 7 From Eqs. (3.6-35) and (3.6-36) follows the relation between circular and escape velocity Vere =V2Ven (3.6-37) Finally, we will derive Kepler’s third law for elliptic orbits. The period, T, of the orbit is found by dividing the area of the ellipse by the areal velocity nab_2na?Vi-e* dA dt T= From Eqs. (3.6-24) and (3.6-27) we see that ~Vap= Seal, 54 Rocket Propulsion & Spaceflight Dynamics and we find for the period T=20— B which verifies Kepler’s third law. A more extensive discussion of motion in a gravitational field will be give! in Chapter 16 where the two-body problem is treated. (3.6-38) 3.7. Mechanics of fluids : The same laws of Newton as applied to particles and bodies in the previous ~ sections, govern the mechanics of fluids. Our treatment will ignore gravity, friction and electromagnetic forces, as their effects are of minor importance, or completely absent in our applications. Friction comes, among others, into play in heat transfer problems (Chapter 8), and will be dealt with there. There are two ways of looking at a fluid. The first one is concerned wit what happens to the fiuid particles in the course of time, what paths the describe, what velocities they possess, etc. This is called the Lagrangiai description. The second one, is the Eulerian description. It attempts to fin out what happens in a given point in the fluid field. We will apply both methods here. 3.7.1 Conservation of mass Consider an undeformable volume, ¥, fixed with respect to an inertial system and bounded by a permeable surface, S, Fig. 3.7. The unit vectot normal to S, directed outward, is called the unit normal, n. The fluid is assumed to have a mass density, p, and a velocity, V, both of which are a function of position and time. In general, the mass contained by ¥ will change due to the flow of matter ‘ through S. The influx of matter through a surface element dS is given by: p p Bets Apert as ‘Surface S Fig. 3.7 The control volume, ¥ Mechanics of Particles, Bodies and Fluids 55 “pV +ndS. The minus sign occurs because n is positive in the outward direction. By integration over the entire surface we find the total flux of ‘matter, m, through S: n=-| pV-nds, Is : which must equal the time rate of change of the mass contained by V, i.e. 2 par=—[ pV-nds. (3.7-1) at ly 5 Equation (3.7-1) is called the conservation of mass equation and states that the time rate of change of mass contained by a volume of fixed shape is equal to the flux of matter through the boundaries of that volume as long as no mass is created spontaneously. | It is practical to derive some other forms of Eq. (3.7-1). According to - Gauss’ Theorem as given in Appendix 1, the right-hand side of Eq. (3.7-1) can be transformed into a volume integral [ pv-nas=[ (v- pyar. Is As ¥ is assumed constant, the order of differentiation and integration can be ‘interchanged. So Eq. (3.7-1) can be written as [ av——[ (V+ pv)dr, at or [ [fev ; ov) dv=0. (3.7-2) As Eq. (3.7-2) must hold for any volume, the integrand must be zero, fay =(pV)=0. @.7-3) ‘Equation (3.7-3) is known as the law of conservation of mass, or the equation of continuity. The reader himself can derive the special forms of Eq. (3.7-3) such as for steady-state flow, (2-0), or for incompressible media (p = constant). oy t A pure one-dimensional form of Eq. (3.7-3) is 3, 90, apV_ (3.7-4) ot ax . 56 Rocket Propulsions & Spaceflight Dynamics n Sidewatt S nd Surface an A Surface ‘Az nt m . : Y V2 Fig. 3.8 The streamtube This form applies only to strictly one-dimensional flows. In many applica: tions, however, we have small deviations from one-dimensionality. In thest cases a quasi one-dimensional form of Eq. (3.7-1) for steady-state flow, which takes into account the variation of the cross-sectional area is ve useful. Consider a streamtube of variable cross-sectional area as given i Fig. 3.8. According to the concept of the streamtube there is no ma transport through the sidewall, S, but only through the surfaces, A, and | Ag, at both ends of the tube. These surfaces, A, and Ao, are arbitrarily’ chosen, and are not necessarily planar. As there is no mass-transpor through S, only the surfaces, A; and Ag, contribute to the surface integral o! Eq. (3.7-1), and this equation can be written as (3.7-5) a [ ware—[ pv-ndA,~| pV-ndA,, Iss and in case of a steady-state flow through the tube, { pV-ndA, --[ pV-ndAy. (3.7.6 Jar laa Often A, and Aj are assumed to be planar surfaces over which both tht density, p, and the normal velocity component, Vy, are constant. In that | case, Eq. (3.7-6) yields the well-known expression, P1 Vins = P2 VanAa- (3.7-7) This equation is often referred to as the one-dimensional steady-state equation of continuity. It should be emphasized, however, that this is only quasi one-dimensional form as the foregoing will have made clear. 3.7.2. Conservation of linear momentum Because Newton’s second law hoids only for invariable mass systems, we consider a material volume, Y, bounded by a fluid surface, S. The material # volume is defined such that it is always composed of the same fluid particle: and so its mass is constant with time. The surface, S, however, may changt in time, due to the displacement of particles. According to Eq. (3.4-5) w’ Mechanics of Particles, Bodies and Fluids 57 (3.7-8) In passing from a system of discrete particles to a fluid, considered as a © continuum, the sum in Eq. (3.7-8) changes into the integral -| VdM. Ja Sinc we neglect gravity, viscosity and electromagnetic effects, the external _ force, F, in.Eq. (3.7-8) consists only of pressure forces on the fluid surface, Ss, -| pn ds. ls ‘The minus sign occurs, because the pressure force is opposite to the unit normal on S. Application of Newton’s second law to the material volume, V, containing fluid of density p gives a [ evar = -| pn ds. (3.7-9) ls dt J, , The total derivative in this equation is, as we will show, dependent on the path. Therefore, consider a scalar or vector quantity in the fluid field, which is.a function of position and time, in general. Using Cartesian coordinates, A=A(t,x, y,z), and the total derivative is given by dA _ dA dAdx dAdy aAdz dt ar ax dt dy dt az dt’ - It is clear that this derivative is dependent on the path, given by x=x(t), y=y(), z= 2(t). ‘As we follow the particles with the material volume, Y, we take a special Path such that a at where u, » and w are the components of the fluid velocity, V. This special total derivative is called the material time derivative or substantial derivative. It is denoted by D a a =, (3.7-10) Db wo ¢ ) and it is easily seen that this is equivalent to pratv'¥ (3.7-11) 58 Rocket Propulsion & Spaceflight Dynamics Returning now to Eq. (3.7-9), we note that we consider a material volum ¥, and so for the total derivative, the substantial derivative has to be taker With the help of the Reynolds’ transport theorem, (Appendix -2), E (3.7-9) can be written as { (E+ ovew “Wt(v- viev)) av= -| pn ds. ia Is Using Gauss’ theorem we can transform the surface integral into a volum integral and obtain [ (g (eV) +pV(V-V)+(¥ wov)+ Vo} av=0. As this must hold for any volume, Y, it follows that 2 (v)+ pvr + V)+(V + ¥)(pv) +¥p =0. B71 Multiplication of the continuity equation, Eq. (3.7-3), by V and subtracting | this form from Eq. (3.7-12) leads to the Euler equation, or the equation of } conservation of linear momentum, av 1 4 Get VIVE Vp =O. (3.7-13).f 3.7.3 Conservation of energy The third relation to be derived, is the energy balance for a flowing flui For an adiabatic system, the first law of thermodynamics (Appendix 3): states: 5E+5W=0. The 6 indicates that we deal with small increments, during a time interval, 6t. The work, 5W, done by the material volume, 7; ' Fig. 3.7, neglecting terms of order 6? and higher, equals: SW =p 57, where | 6Y is the change in the volume, ¥, during the time interval, 6t. For ‘the. material volume, Y, moving along with the fluid, the change in an eleme! tary volume element, dY, bounded by a surface element, dS‘, equals 5 dv =dS'n-V 5t, and as n- V on neighboring elements cancel, the total amount of work, 5 amounts to ow-| pn-VdS5t, : (3.7-148) ls where S is the outer surface of the control volume, V7. In the first law of thermodynamics, E stands for the total internal ener of a system. A flowing fluid not only possesses a specific internal energy, Mechanics of Particles, Bodies and Fluids 59 but also a specific kinetic energy, 3V - V. Thus the energy contained by ¥ is ; e=[ p(et+iV-Vv) dV, and the change in the energy content during the time interval, 51, is oB=5. | ole+iv-vyar 6t. (3.7-14b) Combining the Eqs. (3.7-14a) and (3.7-14b) yields: Bi] pet wars [ pn +“VdS=0. With the help of Gauss’ theorem and Reynolds’ transport theorem, this equation is easily transformed into [ {Rbte+iv-v}+¥-ovye+tv-v+py} av=0, or, with the same reasoning as before Jeol +h¥ + VY +V-(e+3V + V\(pV}+: (p¥)=0. (3.7-15) This is the expression for the conservation of energy. However, simpler expressions can be obtained. Multiplying the continuity equation, Eq. (3.7-3) by e+4V- V and subtracting the result from Eq. (3.7-15) gives a 05, (e+2V-V) + (pV: V)(e+3V°V)+¥ (pV) =0. (3.7-16) Now V-ov)=5v-(ov)+ev-v)(2). By using the equation of conservation of mass, Eq. (3.7-3), we find Vos PBs av-vy(2), / and the energy equation becomes a x(eteav- V)+v-v(c+2eiv-v) 22 . (7-17) ot pat With the concept of enthalpy, h, defined as (Appendix 3), h=e+?, 3.7-18 7 ¢ ) 60 Rocket Propulsion & Spaceflight Dynamics Eq. (3.7-17) can be written as a lop =—(h+3V-+V)+V° 3V-V)--~=0. : u (h4+3V-V)+V-V(h+3V-V) pat : For ideal gases Eq. (3.7-19) can be simplified. Scalar multiplication of Eq, (3.7-13) by V leads to FAV-V)EV-VEV-V)4oV-Vp=0. Subtracting this equation from Eq. (3.7-19) and substituting the perfect gat i law 2 p=pRT, and h=ho+ cT, gives 1 fa _lfep | a{enty Veet} raed Vp} =0. As =? of R= > we find My—1)> WKy—Ds Fin (= - \pev-vfin F \}=0, P P which, after integration yields T= cp (3.7-20: Or, using the perfect gas law, P= cp", T=¢3p™, where c,, c, and cy are constants. These famous results, known as the Poiss relations, are usually derived for a static system in thermodynamics. Fro the derivation given here, it follows that this law also holds for movil} gases. It should be kept in mind that the simple forms Eq. (3.7-20) of the ener} equation hold only if 1. There is no dissipation of energy and no heat is added to or withdray from the system. Therefore the process must be isentropic. Mechanics of Particles, Bodies and Fluids 61 2. The gases must satisfy the perfect gas law, and c, must be independent ‘of the temperature. si iven in cases where these assumptions are questionable, we will often use ihe relations Eq. (3.7-20) instead of the more complicated form of Eq. 7-19), because in practice the differences may be very small. 7.4 Conservation of angular momentum _ Although not an independent equation, we will derive the equation for “conservation of angular momentum, because it will be used in Chapter 10. From Section 3.4, we know that for an invariable system of particles the rate of change of angular momentum with respect to an inertially fixed reference “point, O, is equal to the moment due to the external forces. The angular momentum, B, of the fluid contained by the material volume, Y, with respect to O is, Fig. (3.7), B= [ rxVpdy, (3.7-21) where r is the position vector of an element, dY, with respect to O. The moment, Mg, due to external forces on the material volume, Y, follows from M,=- [ rx np dS, (3.7-22) , and, according to Eq. (3.4-13), » D oo = [rx Vp ev=-[ rXnp ds. (3.7-23) Di , By using the Reynolds’ transport theorem, (Appendix 2), Eq. (3.7-23) can be transformed into L 2 (expv) ars| {ex pV V) + (VV) (rx pv} a { rXup dS. (3.7-24) Js ‘Transformation of the second volume integral into a surface integral, with the help of Gauss’ theorem, leads to { Sox pv) dv+ [ {(pV - n)(rx V)+rx np} dS =0. (3.7-25) Is For many applications, such as turbines and pumps, Eq. (3.7-25) proves to _be very useful. ‘Application of Gauss’ theorem to the surface integral in Eq. (3.7-24) ds to the partial differential equation for the conservation of angular 62 Rocket Propulsion & Spaceflight Dynamics momentum 2 (expV)+ (XVI: pV)+ (V-Wrx V)-Vxrp=0. (3.7-2 We have discussed those aspects of dynamics needed in the followi chapters. Many interesting aspects of dynamics that we will not encoun have purposely been omitted. References 1 Eétvos, R. V., Pekar, D. and Fekete, E. (1922), Beitrage zum Geset der Proportionalitét von Tragheit und Gravitat, Ann. d. Physik, 6 11-66. Roll, P. G., Krotkov, R. and Dicke, R. H. (1964), The equivalence lj inertial and passive gravitational mass, Ann. Phys. 26, 442-517. | Braginsky, V. B. and Panov, V. I. (1971), Verification of the equivalen of inertial and gravitational mass, Zh. Eksp. Teor. Fiz.,61, 873-879 Russian). Einstein, A. and Grossmann, M. (1913), Entwerf einer verallgemein | ten Relativitatstheorie und einer Theorie der Gravitation, Zeit. Math Phys., 62, 225-261. Mgiller, C. (1952), The Theory of Relativity, Clarendon Press, Oxford, 220. Corben, H. C. and Stehle, P. (1965), Classical Mechanics, John Wil New York, p. 139. Slater, J. C. and Frank, N. H. (1933), Introduction to Theoretical Physi McGraw-Hill, New York, p. 96. Chambers, L. G. (1969), A Course in Vector Analysis, Chapman an Hall, London, p. 100.

You might also like