Astronomy Reports, Vol. 44, No. 5, 2000, pp. 279Ð285. Translated from Astronomicheskiœ Zhurnal, Vol. 77, No. 5, 2000, pp. 323Ð330. Original Russian Text Copyright © 2000 by Kondrat’ev.

Dynamics and Stability of Resonant Rings in Galaxies B. P. Kondrat’ev Udmurt State University, Krasnoarmeœskaya ul. 71, Izhevsk, 426034 Russia Received December 24, 1998

Abstract—We have developed a model for a spheroidal, ring-shaped galaxy. The move in a ring with an elliptical cross section at the 1 : 1 frequency resonance. The shape of the cross section of the equilibrium ring depends on the oblateness of the galaxy itself, so that the ellipse of the ring cross section is radially extended when the oblateness of the galaxy is small. If the oblateness of galaxy exceeds some critical value, the ellipse cross section is extended along the Ox3 axis. The shape of the ring cross section is circular for a galaxy with critical eccentricity. The stability of the ring over a wide range of perturbations is studied. A fundamental bicu- bic dispersion equation for the frequencies of small oscillations of a perturbed ring is derived. Application of the model to the ring galaxy NGC 7020 shows that its ring cross section should be approximately circular. Anal- ysis of the dispersion equation demonstrates that stellar orbits in the arm are unstable (but the instability incre- ment is small). We conclude that stars in the ring of this galaxy should drift from the 1 : 1 resonance, and the ring itself should evolve. © 2000 MAIK “Nauka/Interperiodica”.

1. INTRODUCTION appearance of various commensurability ratios. Recall that our solar system is literally saturated with reso- The rings (or partial ring structures) existing in nances between the frequencies of the spinÐorbital many galaxies have long attracted the attention of theo- motions of planets and satellites [7]. rists. The abundance of such stellar systems was empha- sized, for example, in [1], which notes about 200 flat gal- Galaxies are the oldest stellar systems and have axies with ring structures. The most prominent example undergone long periods of evolution. Therefore, we can of a galaxy possessing a distinct nucleusÐring combina- assume that resonances could be formed in galaxies in tion (outwardly similar to Saturn and its rings) is much the same way as in the solar system. Taking such NGC 7702. According to [2], this galaxy has not one resonances into account when developing galactic mod- but two well-separated rings. The well-known exam- els can radically change the dynamical properties of the ples [3Ð5] indicate that ring and spiral-arm structures models.1 Therefore, when constructing models for ring have similar compositions and stellar dynamics. Inves- galaxies, we should first and foremost investigate the tigations of such structures are necessary if we wish to dynamical consequences of resonances between the elucidate the important role of secular evolution pro- frequencies of small oscillations of stars in a plane of cesses in spiral and SO galaxies. the ring cross section. The existence of ring galaxies is a serious theoreti- The present paper is concerned with a stellar cal problem. A quite simple stellar dynamical model for dynamical model for a ring galaxy with a local 1 : 1 res- a galactic ring was developed and its stability studied in onance of the frequencies of stellar motions in a circu- [6]. That paper was primarily concerned with general lar arm with an elliptical cross section. The equilibrium collective effects in the dynamics of prolate structures, model itself is developed in Sections 2Ð5, and its stabil- whereas the dynamics of individual stars within such ity is studied in Sections 6Ð10. We apply this model to structures were not studied. However, the existence of the ring galaxy NGC 7020, whose structure was studied prolate structures and circular rings is possible only if in detail in [9]. there are certain definite relations between the parame- ters of stellar orbits and the corresponding first integrals of motion. Precisely the character of individual orbits is 2. EQUATIONS OF STELLAR MOTION the decisive factor determining the particular form of IN AN ELLIPTICAL ARM the components of the velocity dispersion tensor, and Let us consider an axially symmetric galaxy pos- pressure anisotropy in the “stellar gas,” which is char- ρ acterized by these components, will considerably affect sessing a uniform circular arm with density and ellip- the stability of rings in galaxies. 1 See, for example, [8], which describes a self-consistent phase Therefore, when considering the orbits of individual model for a spheroidal galaxy with an additional resonant integral of the stellar motion. Torus-shaped motions of the cen- stars in the arms, we should first and foremost examine troids arise in this model, in agreement with both observations possible resonances. This is especially important since and numerical experiments in the framework of the N-body the evolution of dynamical systems often leads to the problem.

1063-7729/00/4405-0279 $20.00 © 2000 MAIK “Nauka/Interperiodica”

280 KONDRAT’EV tical cross section. The arm is characterized by a cross motion along the corresponding axis is described by the section with semi-axes a1 and a3: equation of a one-dimensional harmonic oscillator 2 ξ2 xúú = Ð2(α + A )x , (9) x3 ≥ 3 3 3 3 ----2 + ----2 = 1, a1 a3, (1) a3 a1 and obviously does not depend on motions along the R ξ and θ directions. (where = R – R0), and is filled with stellar orbits. Since the cross section and curvature of the generatrices of Let us simplify the problem and assume that the dis- the arm are small, we can take its inner potential at the tribution of mass in the galaxy and ring possesses cir- ξ point ( , x3) in the form cular symmetry; i.e., ϕ α ξ2 α 2 ∂ϕ p = const Ð 1 Ð 3 x3, (2) -∂---θ-- = 0. (10) which is typical for a uniform two-dimensional ellipti- cal cylinder with straight generatrices. The constant Next, let us consider a coordinate system rotating coefficients in expression (2) are equal to Ω with angular velocity c: a a α π ρ 3 α π ρ 1 θ( ) θ˜ ( ) Ω 1 = 2 G ------; 3 = 2 G ------. (3) t = t + ct. (11) a1 + a3 a1 + a3 A in the arm is also affected by the gravitational Then, the first two equations of (8) take the form field of the galaxy itself. Let us expand the galactic field ∂ϕ ú 2 potential in a Taylor series in powers of ξ and x in the Rúú = ------+ R(θ˜ + Ω ) , 3 ∂R c neighborhood of the arm. Then, with accuracy up to (12) quadratic terms, we obtain d ( 2θ˜ú ) Ω ú ---- R = Ð2R c R. ϕ Ω2 ξ ξ2 2 dt g = const Ð c R0 Ð A1 Ð A3 x3. (4) Here, Consequently, we obtain in a linear approximation with ˜ú ∂ϕ respect to ξ and θ Ω2( )  1 g c R0 = Ð------ (5) R ∂R R = R 0 úξú (Ω2 α )ξ Ω θ˜ú = c Ð 2 1 Ð 2A1 + 2R0 c (13) is the square of the angular velocity in a reference orbit with radius R0 from the galaxy center and the coeffi- and cients of the potential are equal to úú R θ˜ = Ð2Ω ξú . (14) ∂2ϕ  ∂2ϕ  0 c 1 g 1 g A1 = Ð------, A3 = Ð------. (6) η θ˜ 2 ∂ 2  2 ∂ 2  By substituting = R0 into the last two equations R R = R0 x3 R = R0 x3 = 0 x3 = 0 and using (9), the complete system of equations for the The total potential inside the arm is the sum of the stellar motion can be reduced to the form contributions: úξú = (Ω2 Ð 2α Ð 2A )ξ + 2Ω ηú , ϕ ϕ ϕ c 1 1 c = P + g (7) ηúú = Ð2Ω ξú , (15) Ω2 ξ (α )ξ2 (α ) 2 c = const Ð c R0 Ð 1 + A1 Ð 3 + A3 x3. (α ) xúú3 = Ð2 3 + A3 x3. The initial equations for the motion of a star inside θ the arm in cylindrical coordinates (R, , x3) are well- After the first integration (to within a constant, which is known: not important in our consideration), the middle equa- tion of system (15) gives ∂ϕ úú θú 2 R = ------+ R , η Ω ξ ∂R ú = Ð2 c . (16)

d 2 ∂ϕ ----(R θú ) = ------, (8) Substituting (16) into the right-hand side of the first dt ∂θ equation of (15), we finally obtain the equations of a ∂ϕ two-dimensional harmonic oscillator, xúú3 = ∂------. x3 úξú ω2ξ = Ð 1 , When solving this system of equations, we first note (17) ϕ xúú = Ðω2 x , the quadratic dependence (7) of on x3. As a result, 3 3 3

ASTRONOMY REPORTS Vol. 44 No. 5 2000 DYNAMICS AND STABILITY OF RESONANT RINGS IN GALAXIES 281 where we have introduced the parameters 4. SLOW EVOLUTION OF A RING GALAXY ω2 Ω2 (α ) In the case of slow evolution, the adiabatic invari- 1 = 3 c + 2 1 + A1 , (18) ants must be conserved. In accordance with (17), we ω2 (α ) ξ 3 = 2 3 + A3 . have a harmonic oscillator with respect to and x3. Consequently, as is well known [11], the products of α α In particular, when the arm is absent (i.e., 1 = 3 = 0), the squares of the amplitudes and the frequency are the first relation of (18) can be written in the equivalent conserved: form 2 2 2 2 a 3Ω + 2(α + A ) = a 2(α + A ). (26) 3 ∂ϕ ∂ ϕ 1 c 1 1 3 3 3 ------g + ------g < 0, (19) R ∂R ∂R2 In addition, the mass per unit length of the arm should be conserved: leading to the well-known criterion for the stability of ρ circular orbits (see, for example, [10]). a1a2 = const. (27) The system of equations (25), (26), and (27) deter- 3. STELLAR MOTIONS mines variations in the arm parameters during the sec- AT THE FREQUENCY RESONANCE ular evolution of the stellar system. The solution of system (17) can be written in the form 5. MODEL CALCULATIONS ξ (ω ) = C1 cos 1t + k1 , (20) As a simple equilibrium model of a galaxy, let us (ω ) x3 = C3 sin 3t + k3 . consider a nonuniform stratified spheroid with similar lay- ers and the realistic normalized density distribution [12] Next, we require that the stellar orbits be closed in the meridional plane of the arm; i.e., that they be repre- 2 1 ρ(m ) = ------, (28) sented by Lissajous figures. In this case, the ratio of the 2 3/2 [1 + βm ] oscillation frequencies must be a rational number: where ω m -----1 = --- . (21) ω n 2 2 2 2 R x3 ---- ≤ ≤ m = ----2- + 2 and 0 m 1, In addition, the condition of conservation of the a˜ 1 a˜ 3 ellipse boundary (1) must be satisfied for all stellar orbits. We shall consider here the simplest case of equal a˜ 1 and a˜ 3 are the semi-axes of the spheroidal galaxy frequencies (i.e., the 1 : 1 resonance) and initial phases: itself, and β is a constant that can be derived from pho- ω ω tometric data. 1 = 3, k1 = k3. (22) It follows from (25) that Then, the motion of a star is described by the formulas ξ( ) (ω ) Ψ t = C1 cos 1t + k1 , 1 Ð ------(23) 4πGρ(R ) ( ) (ω ) χ = ------0----, (29) x3 t = C3 sin 1t + k1 Ψ 1 + ----π------ρ----(------) and under the condition 4 G R0 C a -----1 = ----1 , (24) where C a 3 3 1 ρ(R ) = ------. (30) the stars move along ellipses similar to the boundary 0 [ β 2]3/2 ellipse (1). Such an arm, resembling a swiss roll, con- 1 + R0 sists of concentric tori with homothetic elliptical cross sec- As a result (for details, see the Appendix), the expres- tions. Consequently, the condition that the boundary (1) sion for Ψ from (25) will be equal to be conserved is satisfied. π Note that equality (22) implies the relation Ψ( ) 4 G e = ---2------2 χ e + βR Ψ Ω2 ( ) π ρ1 Ð 0 = 3 c + 2 A1 Ð A3 = 4 G ------(25) 1 + χ 2 2 (31)  1 Ð e2 e2(F Ð E) 2 Ð e + βR  × 2 ------+ E Ð ------0-. between the ratio of the semi-axes of the arm cross sec-  2 β 2 β 2 ( β 2)3/2  χ e + R0 R0 1 + R0 tion = a3/a1 and the parameters of the galaxy itself.

ASTRONOMY REPORTS Vol. 44 No. 5 2000 282 KONDRAT’EV χ 6. SMALL PERTURBATIONS OF THE ARM 3.0 Let the elliptical cross section of the arm (1) at each 2.5 time be deformed into a slightly different ellipse by the linear transformation 2.0 δξ ξ = a + bx3, (32) δ ξ 1.5 x3 = c + dx3, 1.0 where the functions of time a(t), b(t), c(t), and d(t) [whose specific form is given by (45)] are small enough 0.5 that we can neglect their squares in the subsequent analysis. The symmetric part of transformations (32) 0 δ ξ ξ δ 1 = a , 1 x3 = dx3 (33) 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 e obviously leads to changes in the semi-axes a1 a1(1 + a) and a3 a3(1 + d) and, consequently, in the χ density ρ. Dependence of the semi-axis ratio = a3/a1 of the elliptical cross section of an equilibrium ring at the 1 : 1 frequency We can easily find the new coefficients of the inner resonance on the eccentricity of the meridional cross section potential of the elliptical arm. It is obvious that e of a spheroidal galaxy with density distribution (28). The calculations were conducted for NGC 7020. We found, in a aa + da accordance with the photometric data [9], that β ≈ 100, α = 2πGρ------3------1 Ð a Ð ------1------3 ; ≈ ≈ 1 a + a a + a R0/a˜ 1 0.6, and e 0.89. The ellipse of the arm cross sec- 1 3 1 3 ≈ tion turns into a circle at ecr 0.88. i.e., the increment in the first coefficient is a  aa + da  δα π ρ------3------1 3 2 1 = Ð2 G a + ------ . a˜  a1 + a3 a1 + a3 Here, e = 1 Ð ----3 is the eccentricity of a meridional  ˜  a1 In the same way, we find cross section of the galaxy (and all its layers of equal density), R is normalized to a˜ , and F and E are ellip- δα π ρ a1  aa1 + da3 0 1 3 = Ð2 G ------d + ------ . tical integrals of the first and second kinds. a1 + a3 a1 + a3 Formulas (31) and (29) specify the ratio of the semi- On the other hand, the antisymmetric part of trans- χ formations (32) axes = a3/a1 of the arm as a function of the eccentric- β δ δ ξ ity e of the galactic layers via the parameters R0/a˜ 1 2 x3 = bx3, 2 x3 = c (34) β (the central radius of the arm) and . The calculations leads to rotation of the arm through a small angle θ and χ β showed that the dependence of on and R/a˜ 1 is quite a trivial shift of the particles along the elliptical cross weak over a wide range of their values. Therefore, it is section of the orbits. Indeed, possible that the results of our calculations, with rea- δ ξ θξ ε 2ξ sonable accuracy, are characteristic for many ring gal- 2 x3 = c = Ð a3 , (35) axies. Of course, the condition of azimuthal symmetry δ ξ θ ε 2 in these objects must be satisfied. This is true, for exam- 2 = bx3 = Ð x3 + a1 x3, ple, for the ring galaxy NGC 7020. The corresponding where θ is just the small angle of the arm rotation: plot of χ(e) is shown in the figure. 2 2 θ a1c + a3b There is an interesting characteristic feature: The = ------2------2----. cross sections of the resonant rings in weakly oblate a1 Ð a3 galaxies are radially extended, while the cross sections of the arms in galaxies whose oblateness exceeds the 7. PERTURBATION OF THE POTENTIAL critical value ecr are extended along the Ox3 axis. The χ value of ecr is determined by equation (29) for = 1. For The potential of the unperturbed elliptical arm is ≈ NGC 7020, we obtained ecr 0.88, which is very close given above by formula (2). A shift along the ellipse to the mean eccentricity e ≈ 0.89 of the isophotes in the obviously does not affect the arm potential. Therefore, galaxy itself. Consequently, in the resonance model we obtain under consideration, the ring of NGC 7020 has an ϕ˜ α (ξ θ )2 α ( θξ)2 approximately circular cross section. = const Ð 1 + x3 Ð 3 x3 Ð . (36)

ASTRONOMY REPORTS Vol. 44 No. 5 2000 DYNAMICS AND STABILITY OF RESONANT RINGS IN GALAXIES 283

The increment in the potential due to the antisymmetric Consequently, equations (42) reduce to the form part of transformations (35) is equal to 2 d 2 2 2 ------δξ + ω δξ a c + a b 2 1 δ ϕ˜ = 4πGρ----1------3-----ξx . (37) dt 2 ( )2 3 a1 + a3 2 2  aa + da  a c + a b Then, the total increment of the potential at the inner α 1 3 ξ π ρ 1 3 = Ð 2 1a + ------ + 4 G ------2 x3, point (ξ, x ) is given by the formula a1 + a3 (a + a ) 3 1 3 (44) 2  aa + da  2  aa + da  2 d 2 δϕ˜ α 1 3 ξ α 1 3 ------δ ω δ = Ð 1a + ------ Ð 3d + ------ x3 2 x3 + 1 x3 a1 + a3 a1 + a3 dt 2 2 (38) 2 2 a1c + a3b π ρ ξ  aa1 + da3 a1c + a3b + 4 G ------2 x3. α π ρ ξ ( ) = Ð 2 3d + ------ x3 + 4 G ------2 . a1 + a3 a + a ( ) 1 3 a1 + a3

8. MOTION OF A PARTICLE 9. THE DISPERSION EQUATION IN THE PERTURBED ARM ξ Let us now consider the motion of a particle. The The unperturbed solutions (t) and x3(t) from (23), equations of its dynamics in the unperturbed potential as well as the time dependences for the coefficients iωt iωt 2 L a = a0e , c = c0e , Φ = ϕ Ð ------, R = R + ξ, (39) (45) 2 0 iωt iωt 2R b = b0e , d = d0e (where ϕ is given by formula (7), and L is the integral (where ω is the frequency of small oscillations of the of angular momentum) will be perturbed arm), must be substituted into the right-hand 2 2 sides of equations (44). In addition, d ξ ∂Φ d x3 ∂Φ ------= ------, ------= ------. (40) ω 2 ∂ξ 2 ∂ δξ [ ξ( ) ( )] i t dt dt x3 = a0 t + b0 x3 t e , (46) δ [ ξ( ) ( )] iωt In the perturbed state, x3 = c0 t + d0 x3 t e , ξ ξ δξ δ ϕ ϕ δϕ + , x3 x3 + x3, + ˜ , [where ξ(t) and x (t) are also taken from (23)] must be δξ3 δ substituted for and x3 on the left-hand sides of the in place of (40), we obtain the following set of inhomoge- above equations. After these substitutions into (44) and neous, ordinary differential equations for the unknown δξ δ calculation of the second derivatives of functions (46), quantities and x3: the system (44) can be reduced after extensive manipu- lation to the equations d2 ∂2Φ ∂2Φ ∂ ------δξ Ð ------δξ Ð ------δx = -----δϕ˜ , 2 2 ∂ξ∂ 3 ∂ξ ( , , ) (ω ) dt ∂ξ x3 Q1 a0 b0 d0 cos 1t + k1 (41) ( , , ) (ω ) d2 ∂2Φ ∂2Φ ∂ + Q2 a0 b0 c0 sin 1t + k1 = 0, ------δ ------δξ ------δ ------δϕ˜ (47) 2 x3 Ð Ð 2 x3 = , ( , , ) (ω ) ∂ξ∂x ∂ ∂x Q3 b0 c0 d0 cos 1t + k1 dt 3 x3 3 + Q (a , c , d )sin(ω t + k ) = 0. where δϕ˜ is given by formula (38). For the reduced 4 0 0 0 1 1 potential Here,

2 Φ ϕ L ω2 α  a1  = Ð ------2-, (42) Q1 = a0C1 Ð + 2 11 + ------ 2 ξ  a1 + a3 2R01 + ----- R0 [ ωω ] α a3 + b0C3 2i 1 + d0C1 2 1------, we obtain with the required accuracy the expression a1 + a3 Φ Ω ξ ( α )ξ2 [ ωω ] = const Ð c R0 Ð A1 + 1 Q2 = a0C1 Ð2i 1 2  2 2 2 L ξ ξ (43) a Ð (A + α )x Ð ------1 Ð 2----- + 3----- . + b C Ð ω2 Ð 4πGρ------3------3 3 3 2 R 2 0 3 ( )2 2R0 0 R0 a1 + a3

ASTRONOMY REPORTS Vol. 44 No. 5 2000 284 KONDRAT’EV

Solutions of the dispersion equation Here, we have introduced the notation

2 e χ ω2 ω2 ω2 χ( χ) χ 1 2 3 γ 2 + γ γ 1 1 = ------, 2 = ------, 3 = ------, (1 + χ)2 (1 + χ)2 (1 + χ)2 0.25 0.029431 5.341189 5.133680 Ð0.176781 (52) χ 0.5 0.139000 5.442480 5.005362 Ð0.140321 γ 2 + 1 4 = ------2 , 0.75 0.461141 5.439000 4.615930 Ð0.078983 (1 + χ) 0.85 0.810901 5.224364 4.292386 Ð0.060398  a  2 0.88 0.998620 5.091055 4.151278 Ð0.060223 and χ = ----3 . In addition, ω2 and ω are normalized  a  1 0.90 1.169900 4.970489 4.037862 Ð0.063190 1 to 4πGρ: 0.95 1.995270 4.467905 3.633548 Ð0.093558 ω2 ω2 ≡ ------; 4πGρ 2 (53) a1 Ω2 + c C Ð4πGρ------, (48) 2 1 A χ 3 A1 0 3 ( )2 ω = ------+ ------3----- = ------+ ------c--- + ------. a1 + a3 1 1 + χ 4πGρ 1 + χ 4πGρ 4πGρ

2 π ρ a3 10. ANALYSIS OF THE DISPERSION EQUATION Q3 = b0C1 Ð4 G ------( )2 2 a1 + a3 Since ω ≠ 0, the dispersion equation (50) can be reduced to the bicubic equation: 2 ω2 π ρ a1 [ ωω ] ω6 ω4 ω2 + c0C1 Ð Ð 4 G ------+ d0C3 2i 1 , + K1 + K2 + K0 = 0. (54) (a + a )2 1 3 We calculated the coefficients of this equation for the β same parameters and R/a˜ 1 as in the figure. Solutions α a1 Q4 = a0C3 2 3------of equation (54) for several values of e are presented in a1 + a3 the table. We can see that there are two positive and one neg- [ ωω ] ω2 α  a3  + c0C1 Ð2i 1 + d0C3 Ð + 2 31 + ------ . ative square of the frequency at each value of e. This a1 + a3 implies the existence of imaginary frequencies and, therefore, the instability of the stellar orbits in the arm Since equations (47) must be satisfied at any instant around the galaxy. We can also see that the absolute of time t, we impose the requirement value of the negative square of the frequency is small everywhere. Consequently, the instability increment of Q = Q = Q = Q = 0. 1 2 3 4 (49) the stellar orbits is also a small quantity. Equality (49) represents a system of four linear alge- braic equations for the unknown quantities a0, b0, c0, 11. CONCLUSION and d0. Since the right-hand sides of these equations are equal to zero, they have nontrivial solutions only if the As can be seen in the figure, the most interesting determinant of this system is equal to zero. This condi- consequence of our assumption that the frequencies of tion yields the basic dispersion equation of the problem: stellar oscillations in an equilibrium ring have a reso- nant character is that the shape of the ring cross section ω2(ω6 + K ω4 + K ω2 + K ) = 0 (50) depends on the oblateness of the spheroidal galaxy 1 2 0 itself. In the framework of the model considered, rings with the coefficients whose cross sections vary from flat to circular and fur- ther to prolate in the direction perpendicular to the sym- γ γ γ γ ω2 metry plane of the galaxy are possible from the dynam- K1 = 2 + 3 Ð 1 Ð 4 Ð 8 1; ical point of view. In particular, the ring in the galaxy K = 16ω4 Ð 4ω 2(γ + γ Ð γ Ð γ ) NGC 7020 should have an approximately circular arm 2 1 1 2 3 1 4 cross section in the case of the 1 : 1 resonance. We have γ γ γ γ (γ γ )(γ γ ) studied the stability of rings in the general case of linear + 1 4 Ð 2 3 Ð 1 + 4 2 + 3 ; (51) affine perturbations. Our solution of the dispersion equation for NGC 7020 shows that orbits in the ring ω2(γ γ γ γ γ γ ) will be slightly unstable. We conclude that the stellar K0 = 4 1 1 2 + 3 4 Ð 2 2 3 orbits in the ring of this galaxy will drift from the 1 : 1 (γ γ )(γ γ γ γ ) + 2 + 3 1 4 Ð 2 3 . resonance, and the ring will slowly evolve as a whole.

ASTRONOMY REPORTS Vol. 44 No. 5 2000 DYNAMICS AND STABILITY OF RESONANT RINGS IN GALAXIES 285 Ω2 REFERENCES Finally, c can be expressed in terms of the incom- 1. S. N. Nuritdinov and M. V. Usarova, Problems in the plete elliptical integrals of the first and second kinds: Physics and Dynamics of Stellar Systems (Tashkent, 1989), p. 49. 2 2 4πG 1 Ð e 2. R. Buta, Astrophys. J. 370, 130 (1991). Ω = ------[F(ϕ, k) Ð E(ϕ, k)], (A.4) c β 2 2 β 2 3. B. A. Vorontsov-Vel’yaminov, Astron. Zh. 37, 381 R0 e + R0 (1960). 4. B. A. Vorontsov-Vel’yaminov, Extragalactic Astronomy where (Nauka, Moscow, 1978). 2 β 5. C.-Y. Wong, Astrophys. J. 190, 675 (1974). ϕ 1 Ð e = arccos ------, k = R0 ------. 6. V. A. Antonov and S. N. Nuritdinov, Astrofizika 19, 547 1 + βR2 e2 + βR2 (1983). 0 0 7. H. Alfven and G. Arrenius, Evolution of the Solar System Next, the second derivative of the potential in the (NASA, Washington, 1979). equatorial plane of the galaxy is equal to 8. B. P. Kondrat’ev, Astron. Zh. 74, 845 (1997). 9. R. Buta, Astrophys. J. 356, 87 (1990). ∂2ϕ g Ω 2 π 2 β 2 = Ð c + 6 Ga˜ 1a˜ 3 R0 10. S. Chandrasekhar, Principles of Stellar Dynamics (Univ. ------2- Chicago Press, Chicago, 1942; Inostrannaya Literatura, ∂R R = R0 x3 = 0 Moscow, 1948). ∞ (A.5) 11. A. S. Bakaœ and Yu. P. Stepanovskiœ, Adiabatic Invariants du (Naukova Dumka, Kiev, 1981). × ∫------. [ 2 β 2 ]5/2 ( 2 )( 2 ) 12. B. P. Kondrat’ev, Dynamics of Ellipsoidal Gravitating Fig- 0 a˜ 1 + R0 + u a˜ 1 + u a˜ 3 + u ures (Nauka, Moscow, 1989). 2 β 2 –1/2 After the substitution x = [a˜ 1 + R0 + u] and a series APPENDIX of transformations, we obtain the following expression Since the ring in NGC 7020 is an inner ring [2], we for 2A1 from formula (6): shall use the formulas for the inner potential of nonuni- 2 form stratified ellipsoids from monograph [12], ch. 2, 2 1 Ð e 2A = Ω Ð 4πG ------when calculating the quantities appearing in expression 1 c 2 2 e + βR (25) for Ψ. In the case under consideration, the layers 0 with constant density are spheroids that are similar to  1 1 Ð e2 each other. This simplifies the calculations. × ------(A.6) ( β 2)3/2 2 β 2 First, we must find the derivative 1 + R0 e + R0

∞ ∂ϕ ρ[ 2( )] (2 + k2)F(ϕ, k) Ð 2(1 + k2)E(ϕ, k)  g 2 m u du ------ ------= Ð2πGa˜ a˜ R ------. (A.1) + 2 . ∂ 1 3 ∫ 2 β R ( 2 ) 2 R0  0 a˜ 1 + u a˜ 3 + u By substituting here the density distribution (28) in the In the same way, after quite simple but cumbersome form calculations, we obtain π ρ[ 2( )] 1 4 G m u = ------3/2----, (A.2) 2A3 = ------ 2 2  2 β 2 R x e + R0 1 + β------+ ------3------ 2 2 (A.7)  ˜ ˜  2 a1 + u a3 + u  1 1 Ð e  × ------+ ------[F(ϕ, k) Ð 2E(ϕ, k)].  β 2 e2 + βR2  Ω 2 1 + R0 0 we obtain for c from formula (5)

∞ Substituting expressions (A.4), (A.6), and (A.7) into Ω 2 π 2 du c = 2 Ga˜ 1a˜ 3 ------.(A.3) (25), we finally obtain the required formula (31). ∫ 2 3/2 2 βR 0 ( 2 ) 2 0 a˜ 1 + u a˜ 3 + u 1 + ------˜ 2 a1 + u Translated by Yu. Dumin

ASTRONOMY REPORTS Vol. 44 No. 5 2000

Astronomy Reports, Vol. 44, No. 5, 2000, pp. 286Ð297. Translated from Astronomicheskiœ Zhurnal, Vol. 77, No. 5, 2000, pp. 331Ð344. Original Russian Text Copyright © 2000 by Tyul’bashev, Chernikov.

Physical Conditions in Steep-Spectrum Radio Sources S. A. Tyul’bashev1 and P. A. Chernikov2 1Pushchino Radio Astronomy Observatory, Astro Space Center, Lebedev Institute of Physics, Leninskiœ pr. 53, Moscow, 117924 Russia 2Sternberg Astronomical Institute, Universitetskiœ pr. 13, Moscow, 119899 Russia Received July 22, 1998

Abstract—Interplanetary scintillation observations of the radio sources 4C 31.04, 3C 67, 4C 34.07, 4C 34.09, OE 131, 3C 93.1, OF 247, 3C 147, 3C 173, OI 407, 4C 68.08, 3C186, 3C 190, 3C 191, 3C 213.1, 3C 216, 3C 237, 3C 241, 4C 14.41, 3C 258, and 3C 266 have been carried out at 102 MHz. Scintillations were detected for nearly all the sources. The integrated flux densities and flux densities of the scintillating components are estimated. Nine of the 21 sources have a low-frequency turnover in their spectra; three of the sources have high- frequency turnovers. The physical parameters are estimated for sources with turnovers in the spectra of their compact components. In most of the quasars, the relativistic-plasma energy exceeds the magnetic-field energy, while the opposite is true of most of the radio galaxies. Empirical relations between the size of the compact radio source and its magnetic field and relativistic-electron density are derived. © 2000 MAIK “Nauka/Inter- periodica”.

1. INTRODUCTION 2. OBSERVATIONS Observations at 2.7 GHz [1, 2] have shown that Our 102-MHz interplanetary-scintillation observa- about 30% of strong radio sources have angular sizes tions were carried out in 1995Ð1997 on the Large less than ″ and steep spectra. Thus, there exists a spe- Phased Antenna (LPA) of the Lebedev Physical Insti- 2 tute. The effective area of the antenna in the zenith cial sample of compact radio sources with steep spectra direction is 3 × 104 m2. The receiver time constant is above 1 GHz. These sources have been actively studied τ = 0.4 s and the receiver bandwidth is about 200 kHz. with high angular resolution and high sensitivity at cen- The rms confusion for the LPA due to extended (non- timeter and decimeter wavelengths, while observations scintillating) sources is 1 Jy. Here, we present results at meter wavelengths are virtually absent. We decided to for 21 sources. conduct meter-wavelength interplanetary-scintillation observations of these compact steep-spectrum (CSS) We calibrated the observations using radio sources from the 3C catalog. As a rule, data for no fewer than radio sources in order to acquire a better understanding five calibration sources were recorded in each observ- of their nature. The basis for these observations was the ing session. All flux-density estimates were obtained sample of 62 CSS sources published in [3]. using the scale of Kellermann [5]. The observation We presented results for 12 sources in [4], our first reduction method is described in [6, 7]. The method paper in this cycle. We took primarily strong sources used enables the detection of weak scintillating sources from the initial sample, for which there were abundant whose scintillation dispersions are smaller than the high-frequency observations. This enabled us to con- noise dispersion. struct the spectra of compact (< 0.1″) features. The We present a detailed description of the results for spectra for the compact features of all 12 sources had each source. As a rule, the sources have complex struc- low-frequency turnovers. Assuming that these were the tures, and our observations separate out the main com- result of synchrotron self-absorption, we obtained esti- ponent (or components) contributing to the scintilla- mates of the magnetic field, number density of relativ- tions. Unless otherwise indicated, scintillations were istic electrons, and energies in the magnetic field and in reliably detected. Unfortunately, we are not able to pub- relativistic electrons. In most cases, equipartition of lish all the spectra (they are accessible in [8]); there- energy between the magnetic field and relativistic par- fore, the spectra were divided into groups with charac- ticles was violated. teristic behaviors. The present paper is the second in this cycle. Our main goal is to study the physical conditions in the 4C 31.04 compact (≤1″) features of the CSS sources studied and 4C 31.04 is a radio galaxy with redshift z = 0.059. to search for differences in these physical conditions Observations at 327 MHz with resolution 0.064″ × for the quasars and radio galaxies in the sample. 0.041″ [9] showed that it has a double structure with

1063-7729/00/4405-0286 $20.00 © 2000 MAIK “Nauka/Interperiodica”

PHYSICAL CONDITIONS IN STEEP-SPECTRUM RADIO SOURCES 287 separation 0.07″ in the southeastÐnorthwest direction. S, Jy Both components make roughly equal contributions to 4ë31.04 the scintillations at 102 MHz, since their flux densities dif- ~ fer by less than a factor of two and they are both equally Radio lobes compact (0.03″ × 0.02″ for the southeast component and 0.043″ × 0.024″ for the northwest component). 1.5-GHz 1 observations with resolution ~1″ and ~7′ [9, 10] show an extended (~2″) halo whose flux density is less than 0.01 0.1 1 10 6% of the total flux density. ν, GHz We observed 4C 31.04 over four days at elongations ~26°. Strong scintillations were recorded. Taking into Fig. 1. Integrated spectrum of 4C 31.04 (filled triangles) based on points at 31.4 GHz (0.71 Jy) [11], 10.7 GHz (1.03 Jy) account the angular size of the source and its elonga- [12], 4.85 GHz (1.57 Jy) [13], 2.7 GHz (1.93 Jy) [14], tion, we estimate the flux density of the scintillating 1.4 GHz (2.68 Jy) [15], 0.408 GHz (3.6 Jy) [15], 0.178 GHz component to be Sscin = 2.6 Jy. We estimate the inte- (2.4 Jy) [16], and 0.102 GHz (4.7 Jy) [this paper]. Spectrum for the compact (scintillating) component (hollow triangles) grated flux density to be Sint = 4.7 Jy. The integrated spectrum of the source (Fig. 1) shows a clear flattening based on points at 5 GHz (1.48 Jy) [10], 1.5 GHz (2.46 Jy) [10], 0.327 GHz (3.3 Jy) [9], and 0.102 GHz (2.6 Jy) [this toward lower frequencies; given the large errors in the paper]. flux density due to the effect of confusion, there may even be a turnover in the integrated spectrum. The spec- trum of the compact radio emission also has a maxi- ~0.1″ [19] show that it is a very compact source sur- mum near 200 MHz and a turnover at lower frequen- rounded by an extended halo. The halo contributes cies. Note that, in connection with the small number of about 2% of the integrated flux density at 5 GHz and observations with high angular resolution (~0.01″), the about 20% of the integrated flux density at 1.67 GHz. inferred spectrum of the compact emission corresponds Since the halo is rather large (~2″ at 1.67 GHz [19]), we to the sum of the flux densities of both compact compo- do not expect it to make a significant contribution to nents. In this case, the turnover in the spectrum could scintillations at 102 MHz. 1.67-GHz observations with reflect a turnover in the spectrum of either one or both resolution ~0.01″ [20] indicate that more than 50% of components. the total flux density at this frequency is associated with a component 0.006″ × 0.003″ in size. We expect that 3C 67 precisely this component will dominate the scintilla- tions at 102 MHz. This source is a radio galaxy with redshift z = 0.31. The radio source is double, with northern and southern We observed 4C 34.07 over three days at elonga- components separated by 2.4″. The southern compo- tions near 33°. There were large variations in the flux nent contributes about 60% of the total flux at 1.6 GHz density from day to day (∆S = 1.4 Ð 2.4). This may be due [3]. At 0.6 GHz, the northern component is only half as to the small angular size of the scintillating radio source. strong as the southern component and is twice as large. We estimate Sscin = 2.2 Jy. The integrated spectrum shows This suggests that the contribution of the southern com- some steepening at low frequencies (Sint = 6 Jy). This over- ponent will dominate at 102 MHz. This component is estimation of the integrated flux density is probably the 0.06″ × 0.02″ in size and is surrounded by a diffuse halo. result of confusion. The spectrum of the scintillating radio The flux density of the halo is about 80% of the flux source has a maximum at 300Ð500 MHz and turnover at density of the southern component at 1.6 GHz [17]. lower frequencies (similar to the spectrum in Fig. 1). 0.6-GHz observations with resolution ~0.1″ [18] yield a size for the halo of 0.09″ × 0.09″. We adopted this size for the characteristic size of the source. The optical 4C 34.09 object that has been identified with the radio source This source is a radio galaxy with redshift z = 0.020. coincides with the southern component, which is prob- ″ ably the radio core. 5-GHz observations with resolution 0.4 [3] show that the source has two components separated by 1.1″ in the 3C 67 was observed over five days at elongations southeastÐnorthwest direction. 1.6-GHz observations ~30°. We estimate the flux density of the scintillating ″ component to be S = 5.2 Jy, and the integrated flux with resolution 0.08 [3] confirm that the source has two scin main components surrounded by a weak halo. The halo is density to be Sint = 14.6 Jy. The integrated spectrum of ″ 3C 67 and the spectrum of its southern component do large (~3 ) and extended in the southeastÐnorthwest direc- not show a low-frequency turnover. tion. Its contribution to the 102-MHz scintillation is insig- nificant. The southeast component contributes ~80% of the integrated flux density at 1.6 GHz, and we expect that 4C 34.07 this same feature will dominate the 102-MHz scintilla- ″ × ″ The redshift of 4C 34.07 is unknown, though optical tions. Its angular size is 0.15 0.075 . observations suggest that it is probably a quasar [3]. We observed 4C 34.09 over three days at elonga- Observations at 1.66, 5, and 15 GHz with resolution tions ~40°. We estimate Sscin = 3.8 Jy and Sint = 11 Jy.

ASTRONOMY REPORTS Vol. 44 No. 5 2000

288 TYUL’BASHEV, CHERNIKOV

S, Jy We observed 3C 93.1 over three days at elongations ~31°. We estimate Sscin = 3.9 Jy and Sint = 11 Jy. The OF 247 radio spectrum is flat.

Core? 1 OF 247 This source is a radio galaxy with redshift z = 0.219. It was unresolved in observations at 1.67, 5, and 15 GHz with resolution ~0.1″ [19], so that it is very compact. 0.01 0.1 1 10 1.67-GHz observations with resolution ~0.01″ [24] ν, GHz indicate that the structure of OF 247 is asymmetric and that there is weak emission from a region to the north- Fig. 2. Integrated spectrum of OF 247 (filled triangles) based east. The characteristic angular size of the source indi- on points at 31.4 GHz (0.38 Jy) [11], 15 GHz (0.92 Jy) [25], cated by the radio contours in [24] is 0.02″ × 0.01″. 10.7 GHz (1.44 Jy) [15], 5 GHz (2.47 Jy) [15], 2.7 GHz (3.18 Jy) [15], 1.41 GHz (3.89 Jy) [15], 0.960 GHz (4.2 Jy) We observed OF 247 over three days at elongations [15], 0.635 GHz (3.96 Jy) [15], 0.430 GHz (3.42 Jy) [15], ~35°. We estimate Sscin = 1.1 Jy; we were not able to and 0.318 GHz (2.65 Jy) [15]. Spectrum for the compact measure the integrated flux density. Figure 2 presents (scintillating) component (hollow triangles) based on points the spectrum of OF 247. The general appearance of the at 15 GHz (1.06 Jy) [19], 5 GHz (2.34 Jy) [19], 1.6 GHz (3.77 Jy) [19], and 0.102 GHz (1.1 Jy) [this paper]. spectrum indicates that the role of the weak halo should be completely negligible, even at the lowest frequen- cies. The spectrum has a peak at 1 GHz and a turnover The integrated spectrum clearly steepens at low fre- at lower frequencies. Therefore, OF 247 is more aptly quencies, probably due to confusion. The spectrum of classified as a GHz-peaked-spectrum (GPS) source the scintillating component does not show a low-fre- rather than a CSS source. quency turnover. 3C 147 OE 131 This is a quasar with redshift z = 0.545. Observa- tions at 22.5, 15, 8.4, 5, and 1.6 GHz with resolution Not much information about OE 131 is available. Its ~0.1″ [19, 22, 26Ð29] show a weak, extended compo- optical counterpart has been identified with a quasar with nent to the north and a stronger, more compact structure redshift z = 2.67. Observations at 5 and 1.66 GHz [20, 21] to the south, stretching 0.25″ from northwest to south- revealed several components with sizes ~ 0.001″; how- east. Observations with resolution ~0.02″ [30Ð34] indi- ever, most (60%) of the flux density at this frequency is cate that the souther component has a coreÐjet structure from a compact region 0.0012″ × 0.0002″ in size. There surrounded by a weak halo. The core, which is at the is no significant halo component of OE 131 at least to southern end of the jet, is very compact (≈0.005″ × 1 GHz. 0.002″) and contributes most of the flux density from We observed OE 131 over six days at elongations compact emission at 1.67 GHz [34]. Superluminal expan- 23° and 32°. There were large jumps in the measured sion has been observed in the core [35]. A comparison flux-density fluctuations, ∆S = 0.35Ð1.3 Jy. This may be of the core and jet at 1.67 GHz, 609 MHz, and 329 MHz due to the small angular size of the compact radio emit- [32] showed that the shape and size of the jet are the same at these three frequencies, but its relative contri- ting region. We estimate Sscin = 0.8 Jy and Sint = 3.6 Jy. The spectrum of the scintillating component has a max- bution to the total flux changes. At 1.67 GHz, the core imum near 400 MHz and a turnover at lower frequen- contributes most of the compact flux density, while the cies (similar to the spectrum in Fig. 1). flux densities from the core and the two knots dominat- ing the jet emission are comparable at 329 MHz. Unfor- tunately, many papers dealing with high-resolution 3C 93.1 observations do not present estimates for the flux den- sities of individual features, forcing us to estimate these This source is a radio galaxy with redshift z = 0.244. flux densities from the radio maps presented. Observations at 1.67 and 5 GHz with resolution ~0.1″ We observed 3C 147 over six days at elongations [19, 22, 23] show a compact component with character- ~29°. This is the strongest source in our sample, both in ″ × ″ istic angular size 0.3 0.2 and a weak extended halo, terms of its integrated flux density (Sint = 84 Jy) and the whose contribution to the scintillations at 102 MHz will flux density of the scintillating component (Sscin = 39 Jy, be negligible. 1.67-GHz observations with resolution for θ ≈ 0.2″). Figure 3 presents an example of a record- ~0.01″ [20] indicate that ~70% of the emission at this ing for 3C 147. The scintillations are clearly visible, frequency comes from a compact region 0.19″ × 0.19″ even in the sidelobes of the antenna. There is an appre- in size. This region should also dominate the 102-MHz ciable turnover in the integrated spectrum, due to a scintillations. turnover in the spectrum of the compact features. How-

ASTRONOMY REPORTS Vol. 44 No. 5 2000 PHYSICAL CONDITIONS IN STEEP-SPECTRUM RADIO SOURCES 289

3C 147

5Jy

Fig. 3. Analog recording of strong scintillations from 3C 147. ever, we were not able to construct a trustworthy spec- arated by 0.052″. Both components have roughly equal trum including our measurement. It appears that this is angular sizes of 0.008″ × 0.004″ and should probably one of relatively few cases in which the components make comparable contributions to 102-MHz scintilla- dominating the scintillations are negligible at high fre- tions. A comparison of the integrated flux densities at quencies. Therefore, we used the spectrum presented in 1.4 GHz (resolution ~10″, 1.57 Jy [39]) and 1.67 GHz [32] to estimate the physical conditions in this source. (resolution 0.007″, 1.37 Jy [3]) indicate that ~90% of 3C 173 This source is a radio galaxy with redshift z = 1.035. S, Jy 1.67-GHz observations with resolution 0.1″ [19] show that it has a double structure with separation 2″ in the 3ë 173 northeastÐsouthwest direction. The northeast compo- 10 nent is an extended (~0.7″) region of low surface brightness and contributes ~10% of the total flux den- sity at 5 GHz [23]. Thus, we expect the 102-MHz scin- tillations to be dominated by the southwest component, ″ × ″ which has angular size 0.2 0.1 . 1 We observed 3C 173 over three days at elongations ~30°. We estimate Sscin = 4 Jy and Sint = 18 Jy. A modest flattening can be seen in the spectrum of the compact component (Fig. 4). However, there is an evident flat- tening of the integrated spectrum at 38 MHz, suggest- ing there is a turnover in the spectrum of the compact 0.1 component near 38 MHz. 0.01 0.1 1 10 ν, GHz OI 407 Fig. 4. Integrated spectrum of 3C 173 (filled triangles) based The redshift of this radio source is unknown; optical on points at 4.85 GHz (0.4 Jy) [13], 1.4 GHz (1.666 Jy) [36], observations indicate that it is probably a quasar [3]. 0.408 GHz (5.01 Jy) [37], 0.178 GHz (9 Jy) [16], 0.102 GHz The source was essentially unresolved in 5-GHz obser- (18 Jy) [this paper], and 0.038 GHz (17 Jy) [38]. Spectrum ″ for the compact (scintillating) component (hollow triangles) vations with resolution ~0.1 [3]. VLBI observations at based on points at 15 GHz (0.098 Jy) [29], 8.4 GHz (0.223 Jy) 1.6 GHz with resolution 0.007″ [3] show that the [29], 5 GHz (0.373 Jy) [23], 1.6 GHz (1.14 Jy) [19], and source is an eastÐwest double with the components sep- 0.102 GHz (4 Jy) [this paper].

ASTRONOMY REPORTS Vol. 44 No. 5 2000 290 TYUL’BASHEV, CHERNIKOV

S, Jy lying in the southeastÐnorthwest direction [29]. How- ever, higher resolution observations show a more com- OI 407 plex structure [18, 28, 29, 40, 41]. The brightest com- ponent at 15 GHz, which has been identified as the core 1 Radio lobes ? [40], is absent from a 0.6-GHz image [18], suggesting the presence of synchrotron self-absorption [28]. Observations with higher resolution reveal additional components. Three components are visible in images ″ ν 0.1 with resolution ~0.1 ( = 15 GHz) [40], while six com- 0.01 0.1 1 10 ponents can be distinguished in 1.6-GHz images with ν, GHz resolution ~0.025″ [41]. Considering all the available radio observations, the radio source has an S-shaped Fig. 5. Integrated spectrum of OI 407 (filled triangles) based structure with the core at the center; the core flux den- on points at 4.85 GHz (0.625 Jy) [13], 1.4 GHz (1.44 Jy) sity is negligible at frequencies below 1 GHz. There is [36], and 0.408 GHz (1.9 Jy) [37]. Spectrum for the compact (scintillating) component (hollow triangles) based on points a jet consisting of a number of weak knots leading to a at 5 GHz (0.61 Jy) [2], 1.67 GHz (1.37 Jy) [3], 1.4 GHz radio lobe to the northwest. There is no jet visible to the (1.568 Jy) [39], and 0.102 GHz (0.25 Jy; upper limit) [this southeast, though the southeast lobe has a compact hot paper]. spot (~0.03″). All the observed components are com- pact and should therefore scintillate at 102 MHz. How- ever, the lowest frequency (ν = 609 MHz) observations the emission is from compact components, and the con- indicate that 2/3 of the energy in compact components tribution of halo emission is negligible. is emitted by the southeast lobe and its hot spot [18]. We observed OI 407 over five days at elongations Thus, we expect the contribution of the southeast lobe ~31°. No scintillations were detected. The extended (θ = 0.2″ × 0.04″) and its hot spot (θ = 0.05″ × 0.015″) component was also not detected. We estimate the to dominate in 102-MHz scintillations. upper limit S ≤ 0.25 Jy. Figure 5 shows the spectrum, scin We observed 3C 186 over four days at elongations which has a maximum at 300Ð500 MHz and a rather ≈ sharp turnover at lower frequencies (α ≈ –1.5, where 37°. Averaging over the days of observations yielded −α S ~ ν ). an estimate of the integrated flux density Sint = 33 Jy, while averaging of the observations of flux-density fluctuations with allowance for the angular size and 4C 68.08 elongation of 3C 186 yielded an estimate of the flux This is a quasar with redshift z = 1.139. 1.67-GHz density for the scintillating radio source Sscin = 7.7 Jy. observations with resolution 0.08″ [3] show that the The spectrum of the southeast lobe and hot spot source has a triple structure oriented in the southeastÐ remains steep to 102 MHz. At the same time, the spec- northwest direction. The central component includes trum of the northwest lobe and jet has a maximum near several features, the brightest and most compact of 1.5 GHz and a turnover at lower frequencies (Fig. 6). which is probably the core. Two extended (~0.6″) regions of emission are located 0.4″ to the northwest and southeast of the central component. The northwest 3C 190 component contributes ~65% of the total flux density at This is a quasar with redshift z = 1.197. Observa- 1.67 GHz, and we expect this component to dominate tions at 15, 5, and 1.6 GHz with angular resolution in 102-MHz scintillations. Most of the energy flux from ~0.2″ [22, 28, 41] show four components extending this region is contributed by a component with angular ″ ″ × ″ roughly 3 in the southwestÐnortheast direction. There size 0.17 0.10 . is also a weak, diffuse halo surrounding this structure. We observed 4C 68.08 over two days at elongation 609-MHz observations with comparable resolution 47°. Due to the large elongation, it was necessary to [18] only detect three of the components visible at make large corrections in order to estimate the flux den- higher frequencies. The angular size of the 609-MHz sity of the scintillating source. If ∆S = 0.6 Jy, then components (~0.3″) are in agreement with those for the Sscin = 1.5 Jy. We estimate Sint = 3 Jy. We were unable to corresponding components detected at the higher fre- construct an accurate spectrum using our measure- quencies. Nan et al. [18] suggest that there is a low-fre- ments. However, in our view, it is unlikely that there is quency turnover in the spectrum of the component that a low-frequency turnover in the spectrum of the com- was not detected at 609 MHz, which is probably the pact emission. core (0.03″ × 0.017″ at 1.66 GHz) [41]. The coordinates of this component are also closest to those of the optical quasar [22]. The spectra of the remaining three compo- 3C 186 nents are steep (α ~ 1), and their expected 102-MHz This source is a quasar with redshift z = 1.063. In flux is 6 Jy. Due to their small sizes, all three of these low-resolution images (~1″), it has two components components should scintillate, so that the observed

ASTRONOMY REPORTS Vol. 44 No. 5 2000 PHYSICAL CONDITIONS IN STEEP-SPECTRUM RADIO SOURCES 291

102-MHz scintillations should correspond to the sum S, Jy of their flux densities. We observed 3C 190 over four days (in the middle 3ë186 of June 1996) at elongations near 40°. The source showed appreciable scintillation, and ∆S ≈ 2 Jy. How- ever, as noted above, these scintillations probably corre- 10 spond to all three compact components. Taking into account the direction of the solar wind, the component sizes, and the elongation, we obtained the crude estimate Southeast Sscin = 2 Jy. We estimate Sint = 24.5 Jy. The spectrum of radio lobe + the core turns over at low frequencies, as noted in [18]. 1 hot spot The spectra of the three remaining components essen- tially merge due to their similar flux densities. Both the integrated spectrum and the spectrum of the compact Northwest radio lobe + emission show a clear flattening. There is probably a 0.1 turnover at lower frequencies (similar to that in Fig. 6). jet

3C 191 This is a quasar with redshift z = 1.956. Few obser- 0.01 vations with high resolution are available. A northÐ south triple structure can be seen in images at 15, 8.4, 5, 0.01 0.1 1 10 and 1.6 GHz [22, 29, 40]. There is a compact hot spot ν, GHz in the northern component. The central component is a relatively weak core, whose spectrum may be flat, but Fig. 6. Integrated spectrum of 3C 186 (filled triangles) based probably has a maximum near 6Ð7 GHz and a turnover at on points at 4.85 GHz (0.308 Jy) [13], 1.4 GHz (1.261 Jy) lower frequencies. A long (~3″) strongly polarized jet [36], 0.408 GHz (5.55 Jy) [37], 0.102 GHz (33 Jy) [this paper], and 0.038 GHz (49 Jy) [38]. Spectrum for one com- extends to the south of the core [29]. The northern compo- pact (scintillating) component (hollow triangles) based on nent is ~0.1″–0.15″ in size and should scintillate strongly points at 15 GHz (0.009 Jy) [40], 8.4 GHz (0.041 Jy) [29], at 102 MHz. The contribution of the core to the scintilla- 1.67 GHz (0.52 Jy) [41], 0.61 GHz (1.019 Jy) [18], and tions should be negligible. The southern jet may or may 0.102 GHz (7.7 Jy) [this paper]. Spectrum for the other not scintillate, depending on its orientation on the sky rel- compact (scintillating) component (hollow squares) based on points at 15 GHz (0.015 Jy) [40], 8.4 GHz (0.042 Jy) ative to the solar wind during the 102-MHz observations. [29], 1.67 GHz (0.5 Jy) [41], and 0.61 GHz (0.315 Jy) [18]. We observed 3C 191 in the middle of June 1996 at elongation 33°. The effective angular size of the jet was ~1.5″, so that its contribution to the scintillations 1.6-GHz observations [19, 23, 29]. The flux density of should have been negligible. Thus, the scintillations the southern feature in the northern component is twice were dominated by the northern hot spot. We estimate that of the northern feature at 1.6 GHz [19], suggesting Sscin = 4.3 Jy and Sint = 23 Jy. The integrated spectrum that the main contribution to the 102-MHz scintilla- remains steep to the lowest frequencies. There is a tions should be the southern compact feature. 1.6-GHz modest flattening in the spectrum of the northern com- observations with resolution ~0.3″ show that this fea- ponent (similar to the spectrum in Fig. 6). ture is compact (0.2″ × 0.2″) and is submerged in a halo 1.2″ × 0.7″ in size. 3C 213.1 We observed 3C 213.1 over four days at elongations near 40°. No scintillations were detected. We estimate This source is a radio galaxy with redshift z = 0.194. an upper limit for S ≤ 0.8 Jy and estimate the inte- ≤ ″ scin Observations with resolution as good as 0.2 are few grated flux density to be Sint = 12 Jy. We are not able to in number and reveal a complex structure [19, 23, 29]. construct an accurate spectrum using our measure- Observations with resolution ~1″ show two large, dif- ments; however, we suggest that it is very likely that fuse regions, in one of which (the northern region) there there is a low-frequency turnover in the spectrum of the are compact features [29]. The overall extent of the compact emission. source is ~30″. 8.4-GHz observations with resolution ~0.2″ show three compact features in the northern com- ponent, lying along a line northÐsouth [29]. The maxi- 3C 216 mum distance between these features is 7–8″. The This source is a quasar with redshift z = 0.668. In southern compact feature in the northern component is low-resolution (~1″) maps, it has an irregular shape ~8″ the most powerful and is strongly polarized (to 38%) in size extending from southwest to northeast (see, for [29]. The central feature in the northern component is example, [28, 29, 42]). In higher resolution (~0.2″) the weakest, is unpolarized (~0.2%), and is probably images, three components can be distinguished, whose the core [29]. This feature was not detected in 5- and relative flux densities vary with frequency [18, 22, 28, 29].

ASTRONOMY REPORTS Vol. 44 No. 5 2000 292 TYUL’BASHEV, CHERNIKOV

S, Jy We estimate Sscin = 6 Jy and Sint = 36 Jy. It is difficult to 3ë 216 derive the spectrum of the jet due to uncertainty about separating the core and jet emission in our observa- tions. It is likely that the spectrum of the jet remains steep to 102 MHz. Note also that the spectrum of 3C 216 is the most complex among those considered in 10 this study.

3C 237

Jet This source is a radio galaxy with redshift z = 0.877. 1 Core The numerous observations of this source at various frequencies with various resolutions indicate that it has an eastÐwest double structure with separation ~1″ [17Ð19, 22, 28, 29, 51]. Both components have sizes ~0.1″–0.2″, Radio lobe? so that they should scintillate at 102 MHz. 0.1 We observed 3C 237 over three days at elongations 0.01 0.1 1 10 30–35°. Powerful scintillations were recorded (∆S = ν, GHz 6.8 Jy). We estimate Sscin = 7.4 Jy for each of the com- ponents, and Sint = 28 Jy. The spectra for the two com- Fig. 7. Integrated spectrum of 3C 216 (filled triangles) based pact features are very similar. They appear to have max- on points at 14.9 GHz (1.27 Jy) [48], 10.695 GHz (1.41 Jy) [49], 8.085 GHz (1.26 Jy) [50], 5 GHz (1.8 Jy) [15], 2.7 GHz ima at 102 MHz and to turn over at lower frequencies. (2.45 Jy) [15], 0.968 GHz (6.2 Jy) [15], 0.408 GHz (11.9 Jy) The integrated spectrum shows a clear flattening at [37], 0.178 GHz (21.54 Jy) [15], 0.102 GHz (36 Jy) [this 102 MHz (similar to the spectrum in Fig. 8 below). paper], 0.038 GHz (51 Jy) [38], and 0.026 GHz (89 Jy) [15]. Spectrum of the radio lobe (?) (hollow triangles) based on points at 15 GHz (0.125 Jy) [28], 8.4 GHz (0.204 Jy) [29], 3C 241 5 GHz (0.45 Jy) [22], 0.6 GHz (0.505 Jy) [18]. Spectrum of the core and jet (?) (hollow squares) based on points at 5 GHz This source is a radio galaxy with redshift z = 1.617. (0.966 Jy) [43], 5 GHz (0.7 Jy) [45], 5 GHz (0.5 Jy) [44], It is an eastÐwest double with separation ~0.9″ [3, 17Ð19, 2.3 GHz (0.5 Jy) [46], 1.6 GHz (0.32 Jy) [47], and 0.6 GHz 22, 23, 28, 29]. Maps with resolution ~0.1″ show that (0.389 Jy) [18]. the western component is made up of two compact (<0.1″) features [3, 17, 18, 23, 28], which should both The southwest component is extended (>1″) and weak, scintillate at 102 MHz. so that it should contribute to the total flux density at We observed 3C 241 over four days at elongations 102 MHz, but should not scintillate. The northeast near 35°. The mean value of the flux variations as ∆S = component is compact (≈0.19″ × 0.035″ [18]) and has a 3 Jy. We estimate Sscin = 2.5 Jy for each of the two com- steep spectrum above 5 GHz (α ≈ 1, see the fluxes in pact components, and Sint = 14.5 Jy. The spectra of the [28, 29, 42]). The spectrum of this component (Fig. 7) compact features clearly have maxima near 400 MHz has a maximum near 2 GHz and turns over at lower fre- and turn over at lower frequencies (Fig. 8). A modest quencies. The central component, in turn, consists of a flattening is also apparent in the integrated spectrum. variable, very compact (~0.1 milliarcsecond [43]) core and ≈0.09″ × 0.05″ jet; superluminal expansion has been observed on milliarcsecond scales [18, 43Ð47]. It 4C 14.41 is difficult to measure the spectrum of the core due to Few data are available for this source; its redshift is its variability. Using observations separated by roughly unknown, but it is expected to be a quasar [3]. The a and assuming that the variability will be modest source was unresolved in 1.6-GHz observations with on such time scales (≤10% of the total core flux den- resolution ~1″ [53]. Higher resolution (~0.01″) obser- sity), it appears that the core spectrum has a maximum vations at 1.66 and 0.6 GHz [3, 54] show that the source at 10 GHz and turns over at lower frequencies, while is double. The components are both compact (~0.01″) the jet spectrum remains steep. The flux densities from and are separated by ~0.08″ in the northwestÐsoutheast the core and jet are approximately equal at 5 GHz direction. We expect both these features to scintillate at [22, 44]. The spectrum constructed (Fig. 7) suggests that 102 MHz. the flux density from the core should fall toward lower We observed 4C 14.41 over two days at elongation frequencies, while the expected flux density from the jet 21°. In spite of the small fluctuations in the flux density rises. Thus, we expect the jet to make the dominant con- (∆S = 0.65 Jy), scintillations were clearly detected. We tribution to 102-MHz scintillation observations. estimate Sscin = 0.4 Jy for each of the two components, We observed 3C 216 over two days at elongations and Sint = 4.5 Jy. The spectra of the scintillating compo- ~40°. The quality of the observations was high. Large nents have obvious low-frequency turnovers (similar to fluctuations of the flux density ∆S = 3 Jy were recorded. the spectrum in Fig. 8).

ASTRONOMY REPORTS Vol. 44 No. 5 2000 PHYSICAL CONDITIONS IN STEEP-SPECTRUM RADIO SOURCES 293

3C 258 S, Jy This is a radio galaxy with redshift z = 0.165. Low- resolution (~2″) 5-GHz maps show a bright, unresolved 3ë 241 core with jets ~30″ in length extending toward the north 10 and south [55]. The core emits a large fraction of the energy. Higher-resolution (0.08″) 5-GHz observations indicate that the core has two components with similar flux densities separated by ~0.1″ in the southwestÐnorth- east direction [23]. 1.6-GHz observations with resolution 1 0.007″ show that these components are ~0.01″ in size [3]. Thus, we expect both components to scintillate at 102 MHz. We observed 3C 258 at the end of September 1997 over two days, in the presence of sharply worsening 0.1 Radio lobes? interference. The observations were of low quality, though scintillations were detected. We estimate Sscin = 0.55 Jy and Sint = 3.5 Jy. There is no low-frequency turnover in the spectra. 0.01 3C 266 0.01 0.1 1 10 ν This is a radio galaxy with redshift z = 1.275. Obser- , GHz vations at 8.4, 5, and 1.6 GHz show that it has a double structure with separation 4″ [19, 29, 56]. The northern Fig. 8. Integrated spectrum of 3C 241 (filled triangles) based ″ × ″ on points at 4.85 GHz (0.388 Jy) [13], 2.7 GHz (0.8 Jy) [52], and southern components have sizes 0.5 0.1 and 1.41 GHz (1.7 Jy) [52], 0.635 GHz (4.8 Jy) [52], 0.178 GHz 0.7″ × 0.3″, respectively [19]. The flux densities of the (11.1 Jy) [52], 0.102 GHz (14 Jy) [this paper], and 0.08 GHz components are roughly equal at all frequencies. We (17 Jy) [52]. Spectrum for one compact (scintillating) com- therefore expect both components to scintillate at ponent (hollow triangles) based on points at 15 GHz (0.046 Jy) 102 MHz. [28], 8.4 GHz (0.124 Jy) [29], 5 GHz (0.22 Jy) [3], 1.6 GHz (0.65 Jy) [3], 0.6 GHz (2 Jy) [18], and 0.102 GHz (2.4 Jy) We observed 3C 266 over six days. Scintillations [this paper]. Spectrum for the other compact (scintillating) were recorded reliably. Given the source elongation component (hollow squares) based on points at 15 GHz (44°), the component separation (4″), and the direction (0.011 Jy) [28], 8.4 GHz (0.043 Jy) [29], 5 GHz (0.095 Jy) [3], of solar-wind propagation, we estimated the flux den- 1.6 GHz (0.57 Jy) [3], 0.6 GHz (2.2 Jy) [18], and 0.102 GHz sity in the compact features and the integrated flux den- (2.4 Jy) [this paper]. sity to be Sscin = 4.2 Jy and Sint = 19 Jy, respectively. There is no low-frequency turnover in the spectra of the (5) spectral indices in the optically thin region of the two compact features. spectrum α, (6) magnetic-field strength, (7) energy of the magnetic field per cm3, (8) total magnetic-field 3. PHYSICAL CONDITIONS IN THE CORES energy in the volume L3, (9) density of relativistic elec- OF THE SOURCES STUDIED trons, (10) energy of relativistic electrons per cm3, and 3 Our observations indicate the presence of low-fre- (11) total relativistic-electron energy in the volume L . quency (102 MHz) turnovers in the spectra of the com- The accuracy of this method is approximately an order pact components of nine of the 21 radio source investi- of magnitude, or two orders of magnitude if the source gated. For three sources, there is a turnover at higher angular size is not accurately known [57]. frequencies. As discussed in [4], the most probable rea- The redshifts of 4C 34.07, OI 407, and 4C 14.41 are son for a low-frequency turnover is synchrotron self- unknown, and we have adopted z = 2. In the formula absorption. Artyukh [57] has presented a method for used to derive the physical parameters considered, the estimating the magnetic field and relativistic plasma redshift enters as (1 + z)–1, so that the value of z only density without resorting to the usual assumption of weakly influences the final result. In the case of 3C 147 equipartition. and 3C 241, the estimates of the physical parameters Assuming that the radio emission of the sources are different for different components; the components studied is synchrotron radiation and that the inferred are marked by letters A and B in Table 1. low-frequency turnovers are the result of synchrotron We can see from Table 1 that all estimates corre- self-absorption, we have estimated the corresponding spond to linear scales L ~ 100 pc, except in the case of physical parameters for all 12 sources with low-fre- 4C 31.04, 4C 34.07, OE 131, and OI 407, where L ~ 10 pc. Ð1 Ð1 quency turnovers. We adopted H0= 75 km s Mpc and Our results indicate equipartition of energy only for q0 = 1/2 for the calculations. The results are presented 4C 31.04. The magnetic field varies over a wide range in Table 1. The columns give (1) the source name, from 10–11 to 105 G. It is obvious that the existence of (2) redshift z, (3) angular size Ω, (4) mean linear size L, fields of 105 or 104 G on scales of 100 pc is improbable.

ASTRONOMY REPORTS Vol. 44 No. 5 2000 294 TYUL’BASHEV, CHERNIKOV

Table 1. Estimates of the physical conditions in compact radio sources with steep spectra

E , Ω, α H⊥ E , erg Ð3 Ee, Name z arcsec L, pc H⊥, G H⊥ ne, cm Ð3 Ee, erg erg cmÐ3 erg cm 4C 31.04 0.059 0.03 × 0.02 20 0.4 10Ð3 10Ð7 4 × 1052 10Ð3 4 × 10Ð7 1053 +0.043 × 0.024 4C 34.07 2 0.006 × 0.003 20 0.4 10Ð7 6 × 10Ð16 1044 103 5 × 101 1061 OE 131 2.67 0.0012 × 0.0002 3 0.5 10Ð10 6 × 10Ð22 3 × 1035 109 109 4 × 1065 OF 247 0.219 0.02 × 0.01 50 1.1 6 × 10Ð2 10Ð4 4 × 1056 4 × 10Ð3 7 × 10Ð8 1053 3C 147 A 0.545 0.17 × 0.05 400 1.1 10Ð1 10Ð3 6 × 1060 10Ð6 10Ð11 3 × 1052 3C 147 B 0.545 0.5 × 0.5 2500 1.3 10Ð3 10Ð7 1059 7 × 10Ð7 6 × 10Ð11 3 × 1055 OI 407 2 0.008 × 0.004 30 0.8 6 × 10Ð5 10Ð10 1050 1 10Ð3 1057 +0.008 × 0.004 3C 186 1.063 0.224 × 0.083 500 1.6 104 4 × 106 7 × 1070 10Ð12 7 × 10Ð20 1045 3C 190 1.197 0.03 × 0.017 100 0.8 105 3 × 108 1070 5 × 10Ð12 10Ð19 4 × 1042 3C 213.1 0.194 0.2 × 0.2 600 0.8 5 1 5 × 1063 6 × 10Ð10 10Ð15 1049 3C 237 0.877 0.2 × 0.1 800 1.4 4 × 10Ð3 7 × 10Ð7 1058 3 × 10Ð5 10Ð9 1055 3C 241 A 1.617 0.1 × 0.1 600 1.1 10Ð2 4 × 10Ð6 1058 3 × 10Ð6 10Ð10 7 × 1053 3C 241 B 1.617 0.1 × 0.1 600 1.6 6 × 10Ð3 10Ð6 7 × 1057 2 × 10Ð5 10Ð9 5 × 1054 4C 14.41 2 0.01 × 0.01 60 0.4 10Ð4 7 × 10Ð10 3 × 1051 6 × 10Ð2 7 × 10Ð5 3 × 1056

It is likely that such results are severely overestimated in Fig. 9a that the distribution of points is not random. due to our inadequate knowledge of the source struc- We have fit linear dependences to both distributions on ture. As noted above, we took the angular sizes for logarithmic scales. The points corresponding to results components from VLBI observations and assumed that obtained in the present paper show the largest scatter these angular sizes did not depend on frequency. How- about these lines. This probably reflects larger errors in ever, in contrast to the situation in [4], where there were the derived physical parameters resulting from the rel- numerous VLBI observations for most of the sources atively sparse input data available for these sources. studied, many of the sources considered in the present The parameters for the two linear dependences fit are paper have only sparse data on the angular sizes of their components. It is likely that some of the angular sizes H = 10–11.7L3.6 G, we have adopted are only upper limits. The method 9.5 –5 –3 used to estimate the physical parameters is very sensi- ne = 10 L cm . tive to the source angular size. Note that, if the density behaves in a fairly natural Let us consider, for example, the quasar 3C 190. fashion (the smaller the scale for a feature, the higher According to [41], the angular size of the core (for which the density in that feature), the behavior of the magnetic we determined the physical parameters) was ≈0.03″ × field is difficult to explain. The field is virtually absent 0.017″. However, the highest resolution measurements on small linear scales and begins to grow as the linear for 3C 190 have resolution 0.034″ × 0.034″ [41]. Thus, size scale increases. the real angular sizes of components could be smaller It is also possible that the turnovers in the spectra are than indicated in radio measurements. For example, if due to thermal absorption. In [4], the possible presence the angular size of the core of 3C 190 is decreased by of thermal plasma in the region emitting synchrotron an order of magnitude, the magnetic field will decrease radiation was considered; the temperature of the non- by four orders of magnitude and become 10 G for a lin- relativistic plasma was ~104 K, and X-ray emission ear scale of 10 pc. For comparison, a magnetic field should come from a very small region Lp < 1 pc. In this >0.1 G was obtained for a scale of 40 pc for the giant case, the emission coefficient is determined by the syn- radio galaxy DA 240 [58]. It is likely that such errors in chrotron radiation, while the absorption coefficient is the angular sizes of components led to errors in the cal- determined by the thermal plasma. It was demonstrated culated physical conditions for 3C 186, 3C 190, and that the inferred density of thermal electrons was com- 3C 213.1. parable to the density of thermal electrons in the cores Figure 9 presents the dependences of the (a) mag- of ultraluminous galaxies. In [4], we consid- netic field and (b) relativistic-electron density on the ered this to be improbable. However, it seems appropri- linear scale of the source. We can see especially clearly ate to estimate in the same way the thermal electron

ASTRONOMY REPORTS Vol. 44 No. 5 2000 PHYSICAL CONDITIONS IN STEEP-SPECTRUM RADIO SOURCES 295

–3 H, G ne, cm 1 (‡) (b) 108

10– 2 106

104 10– 4 102

10– 6 1 10– 2 10– 8 10– 4

10– 6 10– 10 1 10 100 1000 1 10 100 1000 L, pc L, pc

Fig. 9. Dependence of the (a) magnetic field strength H and (b) density of relativistic electrons ne on the linear size of the radio source L. The squares present results from [4] and the triangles results from the present paper. densities for the radio sources studied here, since the Absorption due to the RazinÐTsytovich effect [61, 62] previous estimates for the thermal electron density was also considered in [4], where it was shown to be exceed the expected densities on the size scales consid- improbable. ered by only an order of magnitude. The resulting esti- mates are presented in Table 2. The columns present (1) the source name, (2) linear size, (3) peak frequency in 4. CONCLUSION the spectrum, and (4) inferred thermal electron density. (1) 102-MHz interplanetary-scintillation observa- We can see from this table that the estimates for ne are tions of 21 compact steep-spectrum radio sources car- higher than the densities of thermal electrons in our ried out on the LPA of the Lebedev Physical Institute Galaxy on the same scales [59] and are close to the indicate that nine of the sources have low-frequency thermal plasma densities in ultraluminous IR galaxies turnovers in their radio spectra. Three of the 21 sources [60]. Consequently, as asserted in [4], this mechanism had been known earlier to have high-frequency turn- for the low-frequency turnovers is improbable. overs. (2) The most probable origin for such turnovers is Table 2. Estimates of the nonrelativistic-electron density at synchrotron self-absorption. Based on this assumption, temperature T = 104 K we estimated the magnetic-field strengths, relativistic- electron densities, and the energies of the magnetic ν Ð3 Source name L, pc p, GHz ne, cm field and relativistic electronÐpositron plasma in the compact components of the sources studied. 4C 31.04 20 0.5 140 In all the quasars with low-frequency turnovers in 4C 34.07 20 0.4 110 their spectra, the relativistic-electron energy is much OE 131 3 0.4 290 larger than the magnetic-field energy, with the excep- OF 247 50 1 90 tion of 3C 147, in which the opposite is true. At the 3C 147 A 400 1.6 100 same time, in all the radio galaxies analyzed, the mag- netic-field energy is much larger than the relativistic- 3C 147 B 2500 0.1 3 particle energy, with the exception of 4C 31.04, in OI 407 30 0.6 140 which the two energies are comparable. 3C 186 500 1 60 (3) There may be a correlation between the size of 3C 190 100 3 380 compact features and the physical conditions in them. 3C 213.1 600 0.3 16 The smaller the feature, the higher the density of rela- tivistic electrons and the weaker the magnetic field. At 3C 237 800 0.2 9 the same time, the number of steep-spectrum sources 3C 241 600 0.3 16 for which we have been able to estimate these physical 4C 14.41 60 0.4 66 conditions is insufficient for a statistical analysis of the

ASTRONOMY REPORTS Vol. 44 No. 5 2000 296 TYUL’BASHEV, CHERNIKOV properties of these sources and of possible differences 24. R. B. Phillips and R. L. Mutel, Astrophys. J. 244, 19 between the conditions in quasars and radio galaxies. (1981). We are currently planning a final paper that will present 25. F. N. Owen, R. W. Porcas, and S. G. Neff, Astron. J. 83, our results for the remaining (~20) sources from the 1009 (1978). sample of [3], together with an analysis of our results 26. W. J. M. van Breugel, G. Miley, and T. Heckman, Astron. J. for the sample as a whole. 89, 5 (1984). 27. C. E. Akujor, R. E. Spencer, and P. N. Wilkinson, Mon. Not. R. Astron. Soc. 244, 362 (1990). ACKNOWLEDGMENTS 28. W. J. M. van Breugel, C. Fanti, R. Fanti, et al., Astron. We wish to thank V.S. Artyukh for posing the initial Astrophys. 256, 56 (1992). problem and for his attention to our work, D. Dallacasa 29. C. E. Akujor and S. T. Garrington, Astron. Astrophys. for presenting us with data for some of the sources, and Suppl. Ser. 112, 235 (1995). L.B. Potapova for help in preparing the manuscript. 30. P. N. Wilkinson, A. C. S. Readhead, G. H. Purcell, et al., This work was supported by the State Science and Nature 269, 764 (1977). Technology Project “Astronomy” (project 1.2.5.7). 31. A. C. S. Readhead, P. F. Napier, and R. C. Bignell, Astrophys. J. Lett. 237, 55 (1980). REFERENCES 32. R. S. Simon, A. C. S. Readhead, A. T. Moffet, et al., Astrophys. J. 236, 707 (1980). 1. J. A. Peacock and J. V. Wall, Mon. Not. R. Astron. Soc. 198, 843 (1982). 33. R. S. Simon, A. C. S. Readhead, A. T. Moffet, et al., Astrophys. J. 354, 140 (1990). 2. V. K. Kapahi, Astron. Astrophys. Suppl. Ser. 43, 381 (1981). 34. F. J. Zhang, C. E. Akujor, H. S. Chu, et al., Mon. Not. R. 3. H. S. Sanghera, D. J. Saikia, E. Lüdke, et al., Astron. Astron. Soc. 250, 650 (1991). Astrophys. 295, 629 (1995). 35. W. Alef, E. Preuss, K. I. Kellermann, et al., IAU Symp. 4. V. S. Artyukh, S. A. Tyul’bashev, and P. A. Chernikov, 129: The Impact of VLBI on Astrophysics and Geophys- Astron. Zh. 76, 3 (1999). ics, Ed. by M. J. Reid and J. M. Moran (Reidel, Dor- drecht, 1988), p. 95. 5. K. J. Kellerman, Astrophys. J. 140, 969 (1964). 36. R. L. White and R. H. Becker, Astrophys. J., Suppl. Ser. 6. V. S. Artyukh, Astron. Zh. 58, 208 (1981). 79, 331 (1992). 7. V. S. Artyukh and S. A. Tyul’bashev, Astron. Zh. 73, 669 (1996). 37. A. Ficarra, G. Grueff, and G. Tomassetti, Astron. Astro- phys. Suppl. 59, 255 (1985). 8. http:/www.prao.psn.ru/papers/tyulbashev 38. P. J. S. Williams, S. Kenderdine, and J. E. Baldwin, 9. J. M. Wrobel and R. S. Simon, Astrophys. J. 309, 593 Mem. R. Astron. Soc. 70, 53 (1966). (1986). 39. T. A. T. Spoelstra, A. R. Patnaik, and Gopal-Krishna, 10. R. A. Perley, Astron. J. 87, 859 (1982). Astron. Astrophys. 152, 38 (1985). 11. B. J. Geldzahler and A. Witzel, Astron. J. 86, 1306 (1981). 40. T. V. Cawthorne, P. A. G. Scheuer, I. Morison, et al., Mon. Not. R. Astron. Soc. 219, 883 (1986). 12. I. I. K. Pauliny-Toth, K. I. Kellermann, M. M. Davis, et al., Astron. J. 77, 265 (1972). 41. R. E. Spencer, R. T. Schilizzi, C. Fanti, et al., Mon. Not. R. Astron. Soc. 250, 225 (1991). 13. P. C. Gregory and J. J. Condon, Astrophys. J., Suppl. Ser. 75, 1011 (1991). 42. I. Fejes, R. W. Porcas, and C. E. Akujor, Astron. Astro- phys. 257, 459 (1992). 14. A. Witzel, P. Veron, and M. P. Veron, Astron. Astrophys. 11, 171 (1971). 43. T. J. Pearson and A. C. S. Readhead, Astrophys. J. 328, 114 (1988). 15. H. Kuehr, A. Witzel, I. I. K. Pauliny-Toth, et al., Astron. Astrophys. Suppl. Ser. 45, 367 (1981). 44. P. D. Barthel, T. J. Pearson, and A. C. S. Readhead, 16. J. D. H. Pilkington and P. F. Scott, Mem. R. Astron. Soc. Astrophys. J. 329, L51 (1988). 69, 183 (1965). 45. T. J. Pearson and A. C. S. Readhead, Astrophys. J. 248, 17. C. Fanti, R. Fanti, P. Parma, et al., Astron. Astrophys. 61 (1981). 143, 292 (1985). 46. R. A. Preston, D. D. Morabito, J. G. Williams, et al., 18. Rendong Nan, R. T. Schilizzi, C. Fanti, et al., Astron. Astron. J. 90, 1599 (1985). Astrophys. 252, 513 (1991). 47. A. G. Polatidis, P. N. Wilkinson, W. Xu, et al., Astro- 19. R. E. Spencer, J. C. McDowell, M. Charlesworth, et al., phys. J., Suppl. Ser. 98, 1 (1995). Mon. Not. R. Astron. Soc. 240, 657 (1989). 48. R. Genzel, I. I. K. Pauliny-Toth, E. Preuss, et al., Astron. 20. D. Dallacasa, PhD Thesis, Bologna (1993). J. 81, 1084 (1976). 21. D. Dallacasa, C. Fanti, R. Fanti, et al., Astron. Astro- 49. I. I. K. Pauliny-Toth, A. Witzel, E. Preuss, et al., Astron. J. phys. 295, 27 (1995). 83, 451 (1978). 22. T. J. Pearson, R. A. Perley, and A. C. S. Readhead, 50. F. N. Owen, S. R. Spangler, and W. D. Cotton, Astron. J. Astron. J. 90, 738 (1985). 85, 351 (1980). 23. C. E. Akujor, R. E. Spencer, F. J. Zhang, et al., Mon. Not. 51. B. Anderson and W. Donaldson, Mon. Not. R. Astron. R. Astron. Soc. 250, 215 (1991). Soc. 137, 81 (1967).

ASTRONOMY REPORTS Vol. 44 No. 5 2000 PHYSICAL CONDITIONS IN STEEP-SPECTRUM RADIO SOURCES 297

52. A. Wright and R. E. Otrupcek, Bull. Inf. Centre Donnees 58. V. S. Artyukh and M. A. Ogannisyan, Pis’ma Astron. Zh. Stellaires 41, 47 (1992). 14, 706 (1988). 53. D. W. Murphy, I. W. A. Browne, and R. A. Perley, Mon. 59. D. Downes and A. Maxwell, Astrophys. J. 146, 653 Not. R. Astron. Soc. 264, 298 (1993). (1966). 54. L. Padrielli, W. Eastman, L. Gregorini, et al., Astron. 60. V. S. Artyukh, A. V. Zasov, and S.A. Tyul’bashev, Pis’ma Astrophys. 249, 351 (1991). Astron. Zh. 21, 723 (1995). 55. R. G. Strom, J. M. Riley, H. Spinrad, et al., Astron. 61. V. A. Razin, Izv. Vyssh. Uchebn. Zaved., Radiofiz. 3, Astrophys. 227, 19 (1990). 921 (1960). 56. R. Liu and G. Pooley, Mon. Not. R. Astron. Soc. 249, 62. V. N. Tsytovich, Nonlinear Effects in Plasma (Nauka, 343 (1991). Moscow, 1967), p. 287. 57. V. S. Artyukh, Tr. Fiz. Inst. im. P. N. Lebedeva 189, 223 (1988). Translated by D. Gabuzda

ASTRONOMY REPORTS Vol. 44 No. 5 2000

Astronomy Reports, Vol. 44, No. 5, 2000, pp. 298Ð308. Translated from Astronomicheskiœ Zhurnal, Vol. 77, No. 5, 2000, pp. 345Ð356. Original Russian Text Copyright © 2000 by Danilov.

Equilibrium Phase-Space Density Distribution in Numerical Dynamical Models of Open Clusters V. M. Danilov Astronomical Observatory, Ural State University, pr. Lenina 51, Yekaterinburg, 620083 Russia Received November 12, 1998

Abstract—In dynamical models for open clusters, virial equilibrium is not achieved over the violent relaxation τ ε ε ε time scale vr. The stars form an equilibrium distribution in ( , ζ, l) space, where and l are the energy and angular momentum per unit in the combined field of the Galaxy and cluster and εζ is the energy of motion perpendicular to the Galactic plane per unit mass of cluster stars in the gravitational field of the Galaxy. > τ This distribution of stars changes little when t vr. The stellar phase-space distribution corresponding to this type of equilibrium and the regular cluster potential vary periodically (or quasi-periodically) with time. This phase- space equilibrium is probably possible due to an approximate balance in the stellar transitions between phase- space cells over times equal to the oscillation period for the regular cluster field. © 2000 MAIK “Nauka/Inter- periodica”.

1. INTRODUCTION Kandrup et al. [2] considered the energyÐspace domain occupied by the particles of an isolated system In [1], we considered models of open clusters that that is nonstationary in the regular field, subdividing were nonstationary with respect to the regular cluster this domain into 20-particle intervals at each moment in field. In the course of their dynamical evolution, such time. The mean particle energies in these intervals models develop an equilibrium distribution of stars in remain nearly constant once the system attains a certain the space of certain parameters of the stellar motion equilibrium state. According to [2], this indicates a con- τ over the first violent-relaxation time scale vr. This dis- straint on the coarse-grained phase-space density, > τ tribution varies little with time when t vr. The stellar which restricts the possible outcome of violent relax- phase-space distribution corresponding to this equilib- ation in the system. We believe that this constancy of rium and the regular potential of the cluster vary peri- the average particle energies is due to a balance in the odically (or quasi-periodically) with time. In the open- particle transitions between the energy intervals con- cluster models of [1], virial equilibrium was not sidered in [2], which is typical of systems that have > τ attained, and at t vr, the virial coefficients of the clus- attained equilibrium in the particle energy space. ter models continued oscillating with nearly constant Kandrup [3] performed a theoretical analysis of the amplitude and period. We will call the phase-space den- evolution of collisionless self-gravitating systems. sity of the cluster corresponding to the equilibrium > τ According to [3], in a “coarse-grained” description, such attained at t vr the equilibrium phase-space density. systems can approach a time-independent equilibrium In this equilibrium, there is probably an approximate during their evolution, which corresponds to time- balance in the stellar transitions between phase-space dependent distribution functions f0 for certain parame- cells on time scales equal to the oscillation period of the ters that are “energy extremal” with respect to all per- regular cluster field. turbations δf in the systems of gravitating points. Due to the instability of the phase-space density to Depending on the type of extremum, the function f0 can small perturbations, open-cluster models with low be linearly stable or unstable, and the system’s evolu- phase-space densities develop different equilibrium tion can proceed with linear or nonlinear phase-space Ӎ τ oscillations about f0. At the same time, Kandrup [3] phase-space densities over the time t vr. In each of the intervals of distance from the cluster center consid- points out that it is not clear what sort of coarse-grained ered in [1], the relative differences between the phase- description should be implemented in order to recog- space densities of such models averaged over the stellar nize the approach of the system to equilibrium. < τ > τ velocities increased at t vr and stabilized at t vr. Theoretical and experimental analyses of the prop- Here, tr is the local violent relaxation time a distance r erties and parameters of the equilibrium phase-space from the cluster center. The fact that the mean-velocity density that develops as a result of the violent relax- relative phase-space density differences stabilize and ation of a nonstationary open cluster are of interest remain constant in time indicates the formation of an here. In the case of open clusters moving in circular equilibrium phase-space density in the cluster models. orbits in the Galactic plane, the phase-space density

1063-7729/00/4405-0298 $20.00 © 2000 MAIK “Nauka/Interperiodica”

EQUILIBRIUM PHASE-SPACE DENSITY DISTRIBUTION 299 that forms as a result of violent relaxation is con- plane in the cluster model (here, r = (ξ, η, ζ) is the strained by the conditions of conservation of cluster radius-vector of a cluster star). mass M and energy E, as well as by the symmetries In the collisionless approximation, the right-hand present in the density and stellar velocity distributions sides of the equations of stellar motion (formulas in the cluster. (5.517Ð5.519) in [4]) should be replaced by partial According to formulas (5.517Ð5.519) in [4], the derivatives (multiplied by mi) of the regular potential U equations of motion for cluster stars in (ξ, η, ζ) coordi- with respect to the corresponding coordinates ξ, η, ζ. In nates imply, under a number of assumptions, that the this case, after computing the stellar integrals of motion cluster “energy” is an integral of motion; i.e., E = const in the usual way, it can easily be shown that the energy ε (formula (5.522) in [4]). In addition, if the cluster is of the motion per unit stellar mass i is an integral of the symmetric with respect to the ξζ and/or ηζ planes, the motion: angular momentum of the cluster is another integral of 2 ξ2 ξ2 motion; i.e., Aζ = const (formula (5.530) in [4]). (Chan- v i i i ε = ----- Ð U + α ---- + α ---- = const, (2) drasekhar [4] uses the rotating coordinate system (ξ, η, ζ) i 2 1 2 3 2 fixed to the center of mass of the cluster, where the ξ, η ζ α 2 ξú 2 η2 ζú 2 , and axes are directed from the center of mass of the where 1 = const, v i = i + ú i + i , and the angular cluster away from the Galactic center, in the direction of ε momentum per unit mass li and energy ζ, i of motion Galactic rotation, and perpendicular to the Galactic plane, along the ζ coordinate are not integrals of the motion: ζ d i ζú 2 2 respectively.) After multiplying by i = ------and summing l = ξú η Ð ξ ηú Ð ω (ξ + η ) ≠ const, dt i i i i i c i i (3) over all cluster stars, the stellar equation of motion along εζ, ≠ const, the ζ coordinate (formula (5.519) in [4]) yields i ω N N where c = const is the angular velocity of the cluster d ú ∂Ω with respect to the Galactic center. ---- ∑ m εζ, = Ð∑ ζi------= Q, (1) i i ∂ζ N dt i i = 1 i = 1 The quantities M, E, Aζ = ∑i = 1 mili, and Eζ are con- Ω served during violent relaxation of the cluster (see [4] where mi is the mass of the ith cluster star; and N are the potential energy and the number of stars in the clus- and text above). Following [5], we can obtain the most likely (equilibrium) stellar distribution in (ε, εζ, l) space 1 2 2 ter, respectively; m εζ = -- m (ζú + α ζ ) is the energy from the condition of maximum entropy of the system i , i 2 i i 3 i combined with the above constraints on M, E, Aζ, and of motion of the ith cluster star along the ζ axis in the Eζ. If the possible states are filled by only a few stars Galactic gravitational field; and α = const. 3 (i.e., if the stars occupy only a small fraction of the In the cluster models considered here, Q and Eζ = phase-space microcells belonging to a predefined macro- N ∑ m εζ exhibit oscillations that are damped with cell—a “coarse-grained” cell in the phase space [6]), this i = 1 i , i leads to the MaxwellÐBoltzmann distribution time t. Q oscillates about Q = 0. The largest deviations (ε, ε , ) ( ν µε β γ ε ) of Q from zero are small and correspond to relative Eζ f ζ l = f 0 exp Ð Ð Ð l Ð ζ , (4) variations of several per cent (6.4 and 4.5%) for cluster ν µ β γ models 1 and 2 below, respectively. The oscillations in where f0, , , , are constants. ε If a star cluster develops an equilibrium distribution Q and Eζ are due to oscillations in ζ, i for the relatively ε of the form (4) as a result of various relaxation mecha- small number of stars with the highest ζ, i values (see ε nisms (including those due to stellar encounters), the Fig. 2 below). The relative deviations of ζ, i for these ε ε < τ mean values of , ζ, and l for stars located in pre- stars are, on average, 20% when t vr and decrease by ε ε > τ ε defined cells in ( , ζ, l) space (see below) should be a factor of 2Ð4 when t vr. The oscillations in ζ, i aver- > τ ε conserved at times t vr. Kandrup et al. [2] obtained aged over several stars within specified ζ, i intervals a similar result for ε in models of isolated star clusters (see Fig. 2 below) indicate that the establishment of an for systems that have attained some sort of equilibrium ε equilibrium stellar distribution function in ζ space pro- by the end of the violent-relaxation stage. ceeds more slowly at high εζ than at low and intermedi- ε ε An experimental analysis of the form, parameters, ate ζ. Some of the stars with the highest ζ, i values and properties of the equilibrium phase-space distribution gradually escape from the cluster in the course of its ε ε ε ε of stars F0(r, v) ~ f( , ζ, l) = f( (r, v), ζ(r, v), l(r, v)) for evolution. nonstationary open-cluster models is of interest. Here, Using the condition Q = 0 as an approximation to ξú ηú ζú describe the evolution of the open cluster, we find, in v = ( , , ) is the star’s velocity vector in the cluster. The results of such analyses can be used to construct accordance with (1), that Eζ = const. This condition fol- lows from the symmetry of the potential U(r) and the analytical cluster models that are nonstationary in the regular cluster field, investigate the stellar velocity velocity distribution g(ζú ) with respect to the ζ = 0, fields in such clusters, and estimate the total masses and

ASTRONOMY REPORTS Vol. 44 No. 5 2000 300 DANILOV other cluster parameters from the coordinate and veloc- of θ and θv over the interval [0, π] in model 1, the fac- ity data for cluster stars. θ ξ2 η2 θ ξú 2 η2 The aim of this paper is to analyze the equilibrium tors sin = r/ + and sin v = v/ + ú appear stellar distribution functions in (ε, εζ, l) space and phase in the expressions for the initial density and velocity space that develop during violent relaxation in open- distributions for this model. Here, r2 = ξ2 + η2 + ζ2, v2 = 2 2 cluster models that are nonstationary in the regular ξú + ηú 2 + ζú and, for the sake of brevity, we do not cluster field. present the formulas for the above distributions. According to [7], cluster model 5 develops a core that 2. DESCRIPTION OF MODELS is extended along the ζ axis. Our computations show that model 2 develops a spherical core. The extension We now consider a cluster consisting of N = 500 of the cores of the cluster models analyzed in [7] along stars moving in the Galactic plane about the Galactic the ζ axis was, thus, due to the particular choice of ini- center in a circular orbit of radius 8200 pc. At the initial tial conditions. time t, the cluster is modeled as a system of two con- centric gravitating spheres, imitating the halo and core Models 1 and 2 are defined such that, at t = 0, they of the cluster. We considered open-cluster models with do not rotate with respect to external galaxies, and the R N cluster and its subsystems meet the conditions of virial initial parameters -----1 = 0.24 and ------1 = 0.25 and strong equilibrium for isolated clusters. We used the units 1 pc, R N 2 2 1 Ma, and 1 M᭪ in our computations. For each model manifestations of non-stationarity in the regular cluster with fixed Ri, Ni (i = 1, 2) at t = 0, we computed the field. Here, R1 and R2 are the radii of the cluster core dynamical evolution several times with slightly (and and halo, respectively, and N1 and N2 are the number of randomly) different initial stellar phase-space coordi- stars in the core and halo, respectively. These parame- nates. See [7] for a description of the technique used to ters correspond to model 5 in [7]. We assumed that all generate the magnitudes of the vectors v for the initial cluster stars have mass 1 M᭪. We used a random-num- stellar velocities and small perturbations of the stellar ber generator to specify the initial positions and veloc- phase-space coordinates in the cluster. We smoothed θ ϕ θ ϕ ities of the stars in the (r, , ) and (vr, v, v) spherical the force functions in the right-hand sides of the stellar ξ η ζ ξú η ζú equations of motion (see [7] for a description of the coordinate systems fixed to the ( , , ) and ( , ú , ) technique employed and the smoothing parameter). v Cartesian reference frames. Here, r is a star’s radial ξ η ζ velocity component and r = |r|. In our computations, we used the ( , , ) coordi- nate system (see above) and corresponding equations We analyzed two cluster models. In model 1, as in ϕ ϕ of stellar motion (formulas (5.517)Ð(5.519) in [4]) with the models considered in [7], the angles and v spec- the components of the Galactic field given up to linear ifying the direction of the vectors r and v in the (ξ, η) terms in ξ, η, and ζ. We used the formula for Φ presented and (ξú , ηú ) planes, and also the angles θ and θ speci- by Kutuzov and Osipkov [8] to compute the constant v α α fying the projections of the vectors r and v onto the ζ coefficients 1 and 3 in the expansion for the regular Φ ξ η ζ ζú θ Galactic potential as a power series in , , and . and axes, are randomly distributed in the intervals , In contrast to [4], we have (in the notation adopted here): θ ∈ π ϕ ϕ ∈ π v [0, ] and , v [0, 2 ] using a random-number 2 2 ϕ ϕ ∂Φ  1 ∂Φ ∂ Φ ∂ Φ generator. In model 2, the azimuthal angles and v Φ ≥ 0, ------≤ 0, α = ------Ð ------, and α = – ------, ∂R 1 R ∂R 2 3  ∂ 2  are distributed as in model 1, and the angles θ and θv ∂R 0 Z 0 are distributed in the interval [0, π] in accordance with where R and Z are cylindrical Galactocentric coordi- 1 nates; the subscript 0 indicates that the derivatives of Φ the probability density functions p(θ) = -- sinθ and 2 are taken at R = 8200 pc and Z = 0 for the adopted cir- cular cluster orbit in the Galactic plane. 1 p(θ ) = -- sinθ , respectively. In both models, the dis- v 2 v We monitored the integration errors by checking the tance r is distributed over the intervals [0, R ] with prob- constancy of the integral of the energy E of peculiar i motions of the cluster stars (formula (5.522) in [4]). We 2 3 ability density p(r) = 3r /Ri (i = 1, 2) to ensure con- also used the statistical method [9], which is more sen- stancy of the stellar number density in subsystems at sitive to computational errors, to monitor the accuracy different radii. Model 1 corresponds to model 5 in [7]. of integration of the equations of motion; this allowed The initial stellar number densities in model 2 are us to check the accuracy of the computed phase-space approximately the same at all points of the subsystems density in the dynamical cluster models. considered. In model 1, there is some excess of stars To integrate the equations of motions of the cluster with large |ζ| and |ζú | at t = 0, resulting in a two-peaked stars, we used an improved (optimized) version of the initial velocity distribution, with local maxima at large code [9] (we reduced the number of operations required to solve the equations, eliminated certain operations and small ζú . Because of the uniform initial distribution involving numbers with different sizes in the computer

ASTRONOMY REPORTS Vol. 44 No. 5 2000 EQUILIBRIUM PHASE-SPACE DENSITY DISTRIBUTION 301 memory, etc.). As a result, we increased the time inter- 〈ε(t)〉, (pc/Myr)2 val t for the dynamical evolution during which the 0.3 0 6 accuracy of the computed cluster phase-space density τ can be considered sufficient from t0 = 1.7 vr (the maxi- τ 0.1 mum attained in [9]) to t0 = 1.9 vr for model 1 and t0 = τ 2.2 vr for model 2. As in [7, 9], we adopted an initial 5 τ × 7 –0.1 4 violent relaxation time of vr = 2.6t cr ~5 10 yr; t cr is the average cluster crossing time. 3 –0.3 We integrated the equations of motion using sixth 2 and seventh order RungeÐKutta methods with the grid functions (2) from [9], with an accuracy of 15Ð16 dec- –0.5 imal places. The maximum relative error of the com- 1 × –8 × –9 puted energy E did not exceed 2.8 10 and 2.1 10 for –0.7 0 0.5 1.0 1.5 2.0 2.5 cluster models 1 and 2, respectively. The computations τ also met the statistical criterion for the accuracy of the t/ vr ≤ ≤ computed phase-space density (0.9 Pj 1.0) for all distance intervals ∆r (j = 1, …, 10) considered (see Fig. 1. 〈ε〉 values averaged over five stars as a function of j ∈∆ time. The stars are numbered in order of increasing ε. below). Here, Pj = P(r rj) are the probabilities that Curves 1, 2, 3, 4, 5, and 6 are based on the results for stars the samples of stellar phase-space coordinates obtained 1Ð5, 101Ð105, 201Ð205, 301Ð305, 401Ð405, and 496Ð500 by solving the equations of motions with seventh- and in ε space, respectively. sixth-order integration methods are drawn from the same stellar phase-space coordinate population. According to Ω [9], the differences between the stellar phase-space coor- regular field (here, Ec = T + and T is the total kinetic dinate distributions obtained using the seventh- and energy of the cluster-star peculiar motions). sixth-order methods are random, and the accuracy of ψ The fractional phase-space density differences j the computed stellar phase-space coordinates can be due to instability of the phase-space densities to small considered sufficiently high to use them to draw con- initial perturbations in the stellar phase-space coordi- clusions about the physical properties of the phase- nates increase rapidly during a time tr, then remain space density of a cluster model. approximately constant, and an equilibrium phase- We compared the results obtained with cluster mod- space density is established in cluster models 1 and 2 els 1 and 2 with slightly different initial stellar phase- (see, e.g., Fig. 1 in [1]). The values of tr can be consid- space coordinates. To this end, we subdivided all the ered estimates of the local relaxation time. tr in the cen- Ӎ × τ stars of cluster model 1 at time t into ten groups of 50 tral regions of models 1 and 2, whereas tr 0.5 vr stars in order of increasing distance r from the cluster and t Ӎ 1.1 × τ and t Ӎ 1.5 × τ at the peripheries of ∆ r vr r vr center of mass. Let rj be the distance interval rj corre- models 1 (see also [1]) and 2, respectively. Thus, model 2, sponding to the jth group (j = 1, …, 10). See [1] for the ú technique used to partition the velocity space occupied which has a single-peaked initial distribution of ζ by the cluster stars into k equal cells. When computing coordinates that is closer to the equilibrium distribu- the cluster evolution, we followed the time variations of tion, is slower to develop the equilibrium phase-space ψ the relative differences j of the phase-space densi- density at its periphery. In the central regions, the equi- ties of cluster models 1 and 2 in the distance intervals librium phase-space density develops over essentially ∆ × τ rj averaged over all k velocity-space cells with k = the same time for models 1 and 2 (0.5 vr). 1000. To analyze the properties and parameters of the equilibrium distributions that develop in models 1 and 2, we, like Kandrup et al. [2], subdivide all cluster stars at 3. RESULTS OF COMPUTATIONS each time t into groups consisting of five stars (10 and AND DISCUSSION 20), with each in order of ascending specific energy ε per unit stellar mass [see formula (2)]. We then com- In the cluster models considered, there is initially a 〈ε〉 ε 〈ε〉 small contraction (mainly perpendicular to the Galac- pute the mean and the dispersions of — for each group and perform analogous computations for l and εζ tic plane) that ends by a time t Ӎ (0.30–0.35) × τv . At > × τ r [see formula (3) and the discussion of formula (1)]. Fig- t 0.6 vr, steady-state oscillations in the regular field ures 1Ð3 show 〈ε〉, 〈εζ〉, and 〈l〉 plotted as functions of are established, with periods of P = 0.7 × τv and P = × τ r r r 0.6 vr in models 1 and 2, respectively. The mean time for a number of the stellar groups in cluster model 2 ratios δα of the amplitudes of the virial coefficient α = (the corresponding relations for model 1 are similar to 2E /Ω and the value α = α averaged over the period P those in Figs. 1Ð3). 〈ε〉, 〈εζ〉, and 〈l〉 remain virtually c v r > τ ε ε are equal to 0.55Ð0.59 for models 1 and 2, indicating constant at t vr throughout most of the , ζ, and l that the models are appreciably nonstationary in the domain occupied by the cluster stars. Deviations of 〈ε〉,

ASTRONOMY REPORTS Vol. 44 No. 5 2000 302 DANILOV

2 > τ 〈ε(t)〉, (pc/Myr) t vr. These oscillations are due to the small number 0.7 of stars outside the cluster core, at the cluster periphery. Since 〈ε〉, 〈εζ〉, and 〈l〉 are primarily constant in time < 6 only for stars at distances r Rt (i.e., inside the cluster), 0.5 the constancy of 〈ε〉, 〈εζ〉, and 〈l〉 must be due to stellar encounters within the cluster. (The periods of temporal 〈ε〉 〈ε 〉 〈 〉 oscillations of , ζ , and l are mostly near Pr = τ 〈ε〉 〈ε 〉 〈 〉 0.3 0.6 vr, so that temporal oscillations of , ζ , and l are most likely due to oscillations of the regular cluster 5 field.) 0.1 4 3 The dispersions σε, σε ζ, and σ of the deviations of 2 , l 1 ε, εζ, and l from 〈ε〉, 〈εζ〉, and 〈l〉 in the intervals of these –0.1 parameters considered are also characterized by the 0 0.5 1.0 1.5 2.0 2.5 sizes of these intervals and, at t > τ , should exhibit τ vr t/ vr small, random oscillations about certain constant val- ues over most of the domain S = (ε, εζ, l) occupied by Fig. 2. Same as Fig. 1 for 〈εζ〉. ε ε σ the cluster stars in the ( , ζ, and l) coordinate space. l for stars with small l increases when these stars move to > σ distances r 2Rt . The dispersions ε for the stars with the 〈εζ〉, and 〈l〉 from their constant values become appre- ε < τ highest and lowest change appreciably at t vr (the ciable with time only near the boundaries of the distri- stellar ε distributions develop their “wings” during this ε ε butions of , ζ, and l, where the number of stars is period). The constancy of 〈ε〉, 〈εζ〉, and 〈l〉 and of σε, 〈ε〉 ε > τ σ σ > τ small. The increase of with time at high at t vr ε, ζ, and l for most of the cluster stars at t vr indi- is due to stars at the cluster periphery (near or outside cates that the stellar distribution in (ε, εζ, and l) space the cluster tidal radius r = Rt [10]). The fractional has attained equilibrium. amplitudes of oscillations of 〈ε〉 are small for small ε, < τ > τ Figure 4 shows the distribution of stars in model 2 amounting to ~11 and 6% at t vr and t vr, respec- ε ε in ( , ζ, l) space for seven equally spaced times ti (sep- tively, and are determined by stars located near the ∆ τ τ ∈ 〈 〉 arated by steps t = 0.1 vr for ti/ vr [1.0, 1.6], i = cluster center. The decrease with time of l at small l is 1, …, 7) spanning the period P = t – t for the oscilla- due to stars escaping from the cluster, located at dis- r 7 1 > tions of the regular cluster field. The dashed curves tances r 2Rt from the cluster center. The highest frac- show the distributions averaged over Pr. The central tional amplitudes of oscillations of 〈εζ〉 at large εζ are 1 7 13Ð25% at t < τ and decrease by a factor of 2Ð4 at distributions correspond to t = --∑ t = 1.3 × τv , vr 7 i = 1 i r

〈l(t)〉, pc2/Myr 6

4 5

2 4 0 3 2 –2

–4 1

–6 0 0.5 1.0 1.5 2.0 2.5 τ t/ vr

Fig. 3. Same as Fig. 1 for 〈l〉. Curves 1, 2, 3, 4, and 5 are based on results for stars 101Ð105, 201Ð205, 301Ð305, 401Ð405, and 496Ð500 in l space, respectively.

ASTRONOMY REPORTS Vol. 44 No. 5 2000 EQUILIBRIUM PHASE-SPACE DENSITY DISTRIBUTION 303 and have been constructed as follows. During the inte- n(ε) gration of the equations of stellar motions, the sets of 100 stellar phase-space coordinates for the seven times indi- (a) cated are written to a single file, which is then used to 80 compute the overall stellar distributions in (ε, εζ, l) space. The resulting distributions are then normalized 60 to the number of stars in the cluster (N = 500). Although the cluster is strongly nonstationary, and in spite of the 40 oscillations of the regular field in cluster model 2 over ε ε 20 the period Pr, the “instantaneous” distributions of , ζ, and l at time t differ little from the corresponding distri- 0 –1.2 – 0.9 – 0.6 – 0.3 0 0.3 0.6 ε butions averaged over the period for oscillations of the ε regular field. Figure 5 shows the distributions of the n( ζ) magnitude of the stellar velocities v |v| for the times 160 = (b) indicated, which differ more from the mean distribution than do the distributions in Fig. 4. It is interesting that, 120 during the violent relaxation of the cluster model con- sidered, i.e., for t /τv ∈ [0, 0.6], the differences i r 80 between the instantaneous distributions of εζ and l and the corresponding mean distributions are as small as those in Fig. 4. The differences between the instantaneous 40 ε τ ∈ and mean distributions of for ti/ vr [0, 0.6] are some- τ ∈ what more substantial than for ti/ vr [1.0, 1.6]. Although cluster model 2 is strongly nonstationary 0 0.2 0.4 0.6 0.8ε in the regular field, the instantaneous distribution of the n(l) ζ ε ε > τ 120 stars in ( , ζ, l) space attains equilibrium at t vr; the (c) resulting equilibrium distribution varies only slightly over the period P , does not differ statistically from the ε ε r distribution f( , ζ, l) averaged over Pr, and exhibits only 80 small random deviations from (ε, εζ, l) (see below).

40 4. EQUILIBRIUM DISTRIBUTION OF CLUSTER STARS IN (ε, εζ, l) SPACE We estimated the parameters of the stellar distribu- 0 tion in (ε, εζ, l) space for cluster model 2 for several –5 –3 –1 1 3 5 l times t . To this end, we used the distribution (4) written Fig. 4. The ε, εζ, and l distributions for stars in model 2 for in the more convenient form × τ times ti = [1 + (i – 1) 0.1] vr, i = 1, …, 7. The dashed curves show the n(ε), n(εζ), and n(l) distributions of the f (ε, εζ, l) = f exp(е ε Ð ε Ð β l Ð l Ð γ εζ), (5) 0 τ cluster stars averaged over Pr at time t = 1.3 vr. ε ε where and l are and l averaged over Pr, derived from the set of stellar phase-space coordinates for seven and derived the parameters of distribution (5) using the equally spaced times corresponding to the given t . The method of Marquardt [11] to minimize the sum of the ν ν ± µε ± β quantity from (4) can be written = l , squares of the deviations of f1 from f computed using (5). where the signs in front of the µ and β coefficients We repeated these computations for L = 20 and 40. ε ε depend on those of – and l – l . The table summarizes the resulting coefficient esti- To estimate the parameters of distribution (5), we mates for L = 60 (for L = 20 and 40, the function f is ε subdivided the domain occupied by the cluster stars more coarsely represented by f1 near the maximum f during the given period Pr into L = 60 equal intervals values). We can see from the table that the parameters ∆ε. To determine this domain, we used the combined ε ε of the stellar distribution f( , ζ, l) averaged over Pr for file containing the stellar phase-space coordinates for cluster model 2 vary comparatively little over the time seven equally spaced times spanning the period P . We r interval considered (from × τ to × τ ). also partitioned the ∆εζ and ∆l domains occupied by the t = 1.3 vr t = 1.9 vr ε Such variations of the stellar distribution in (ε, εζ, l) cluster stars during the period Pr into intervals ζ and l in the same way. We then determined the stellar number space could be due, among other things, to approxima- ε ε densities f(ε, εζ, l) in the resulting (ε, εζ, l)-space cells tion errors when fitting the function f1( , ζ, l) using (5).

ASTRONOMY REPORTS Vol. 44 No. 5 2000 304 DANILOV

n(v) 100

80

60

40

20

0 0.4 0.8 1.2 1.6 v, km s–1

× τ Fig. 5. Distributions of stellar velocity magnitudes in model 2 for times ti = [1 + (i – 1) 0.1] vr, i = 1, …, 7. The dashed curve τ shows the n(v) distribution of the cluster stars averaged over Pr at time t = 1.3 vr.

The function (5) used as f(ε, εζ, l) only approximately for each time t from the table. The subscripts 1 and 2 in ε ε ε ε ε ε describes the properties of f1( , ζ, l). A more accurate relations 1, 2 = 1, 2( ζ) refer to the lowest and highest description of the averaged ε distribution of stars could values for stars whose εζ fall in the interval [εζ, εζ + ∆εζ]. probably be obtained by using two exponential func- The subscripts 1 and 2 in relations εζ = εζ (ε) have ε , 1, 2 , 1, 2 tions in (5) to fit the distributions of in the cluster core an analogous meaning. Using the adopted notation, f0 and halo (see Fig. 4a). The small temporal variations of in (5) can be determined from the normalization condi- distribution (5) in the table could also be due to system- tion for (5): atic changes (evolution) of the averaged cluster model. (ε, ε , ) ε ε f0 systematically decreases over the time considered in N = ∫∫∫ f ζ l d d ζdl the table due to the expansion of the cluster in phase S space. ε (6) Ψ(β) max ε ε ε ε ( µ ε ε (γ, ε)) ε When comparing f1( , ζ, l) and f( , ζ, l), we should = f 0------∫ exp Ð Ð C d , ε ε βγ bear in mind that and ζ are not independent. To deter- ε min mine the boundaries of the (ε, εζ) domain occupied by where the stars in the cluster model averaged over Pr, we used the combined file containing the stellar phase-space coor- Ψ(β) ( β( )) ( β( )) = 2 Ð exp Ð l Ð lmin Ð exp Ð lmax Ð l , dinates for seven equally spaced times spanning period Pr . ε ε ε ε ε ε (γ, ε) ( γ(ε (ε))) ( γ (ε (ε))) We derived the relations 1, 2 = 1, 2( ζ) and ζ, 1, 2 = ζ, 1, 2( ) C = exp Ð ζ, 1 Ð exp Ð ζ, 2 .

ε ε Parameters of the stellar distribution in ( , ζ, l) space averaged over Pr for model 2 τ t / vr Parameter 1.3 1.4 1.5 1.6 1.7 1.8 1.9 µ (Myr/pc)2 4.155 ± .089 4.252 ± .087 4.326 ± .082 4.069 ± .079 3.944 ± .076 4.415 ± .083 3.692 ± .066 β (Myr/pc)2 0.984 ± .020 0.977 ± .019 1.022 ± .019 0.995 ± .019 0.992 ± .018 1.050 ± .018 1.040 ± .018 γ (Myr/pc)2 11.090 ± .227 11.143 ± .224 11.810 ± .220 11.177 ± .216 11.656 ± .227 13.049 ± .232 12.331 ± .234

ASTRONOMY REPORTS Vol. 44 No. 5 2000 EQUILIBRIUM PHASE-SPACE DENSITY DISTRIBUTION 305

The subscripts “min” and “max” in (6) indicate the n(ε) minimum and maximum values in the domain S for ε 100 and l. According to (5), the formulas for the stellar dis- (a) tribution in (ε, εζ, l) space can be written in the form 80 χ(ε) (ε, ε , ) ε 60 = ∫∫ f ζ l dld ζ

(l, εζ) 40 Ψ(β) (γ, ε) ( µ ε ε ) = f 0----β---γ-----C exp Ð Ð , 20 0 –1.0 – 0.7 – 0.4 – 0.1 0.2 0.5 0.8 ψ(l) = f (ε, εζ, l)dεdεζ ε ∫∫ n(εζ) (ε, εζ) 150 ε (7) (b) max exp(Ðβ l Ð l ) = f ------exp(е ε Ð ε )C(γ, ε)dε, 0 γ ∫ 100 ε min

ω(ε ) (ε, ε , ) ε ζ = ∫∫ f ζ l d dl 50 (ε, l) Ψ(β) = f ------K(µ, εζ)exp(Ðγ εζ), 0 βµ 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 ε n(l) ζ µ ε µ|ε ε ε | µ|ε ε where K( , ζ) = 2 – exp(– – 1( ζ) ) – exp(– 2( ζ) – 120 (c) ε |) and (l, εζ), (ε, εζ), and (ε, l) refer to the domains for the quantities in parentheses occupied by the stars in the combined file containing stellar phase-space coor- 80 dinates for seven equidistant times in the period Pr. On the whole, formulas (7) for ω(εζ) and ψ(l) with the parameters of the distribution (5) given in the table 40 describe the εζ and l distributions of the cluster stars averaged over P fairly well (see Figs. 6b, 6c). χ(ε) fits r 0 the ε distribution somewhat worse (see Fig. 6a), due to –5 – 3 – 1 1 3 5 l the fact that the cluster includes two subsystems (core and halo) with different ε values (see above for a pos- Fig. 6. The n(ε), n(εζ), and n(l) distributions of the cluster χ ε τ sible way to refine ( )). stars averaged over Pr at time t = 1.3 vr for model 2 (solid To study the properties of the phase-space density curves). The dashed curves show the χ(ε), ω(εζ), and ψ(l) τ function that forms in cluster model 2 as a result of its distributions for time t = 1.3 vr computed in accordance violent relaxation, we compared the (ξú , ηú , ζú ) stellar with formulas (7) and the data from the table. velocity distribution averaged over Pr with the corre- sponding instantaneous distributions for each of the the same population. We also computed the probabili- seven equally spaced times t in each of the 10 distance i ∈ ∆ σ ∈ ∆ ∆ ≤ ≤ ties P (r rj) and dispersions (r rj) for the intervals rj and for each of the intervals 0 r rj, P j ∈ ∆ ∈ ∆ ∆ deviations of P(r rj) from P (r rj) averaged over where rj = ∑k = 1 rk, j = 1, …, 10. We also performed the seven times t ∈ P . We performed analogous com- an analogous comparison of the averaged (ε εζ l) dis- i r , , ≤ tribution of the cluster stars and the corresponding putations for stars at distances r rj, j = 1, …, 10. We instantaneous distributions. We used the Kolmogorov will denote the computed probabilities and their disper- ≤ σ ≤ criterion in the comparisons; see [9] for a description of sions P(r rj) and (r rj). ∈ P the technique used. We computed for each time ti Pr ∈ ∆ Let us first consider the comparison of the (ε, εζ, l) the probabilities P(r rj) that the differences in the stellar distribution averaged over Pr and the corre- stellar parameters considered [(ξú , ηú , ζú ) or (ε, εζ, l)] sponding instantaneous distributions. The probabilities for the distance interval r ∈ ∆r (i.e., r ∈ [r , r ], j j – 1 j ≤ ≤ Ð P(r rj) in the cluster (i.e., at r R , see below) increase where r0 = 0) were purely random and that the corre- t ≤ Ӎ sponding samples of stellar parameters are drawn from with j and distance, from P (r r1) 0.55–0.65 to

ASTRONOMY REPORTS Vol. 44 No. 5 2000 306 DANILOV

≤ Ð Ӎ Ð P (r Rt ) 0.85–0.95, for all times in the table. In this can be considered statistically equivalent is P (r ≤ R ) Ӎ σ ≤ t case, the characteristic P (r rj) are equal to 0.05Ð0.06. 0.65–0.75 [which is lower than the corresponding prob- ε ε Here, RÐ and R+ denote the tidal radii of the cluster for ability for the ( , ζ, l) distribution], while the differ- t t ences between them in velocity space can be consid- stars with “retrograde” and “prograde” motions, respec- ered to be random with the same probability for all the tively, when the star’s angular-velocity vector with respect times t in the period P . Indeed, the instantaneous dis- to the ζ axis is opposite or parallel to the angular-velocity i r vector of the cluster with respect to the Z axis of the tributions of the velocity magnitudes for cluster model 2 in Fig. 5 differ more from the average distribution than Galactic rotation. The technique we used to estimate ε ε Ð + the instantaneous ( , ζ, l) distributions differ from the Rt and Rt from the distance dependences of the corresponding average distributions in Fig. 4. squares of the velocities and velocity dispersions for τ ∈ ∆ stars with prograde and retrograde cluster orbits is At times t = (1.3–1.5) vr, the probabilities P (r rj) ∈ ∆ Ӎ described in [7]. In our models, RÐ Ӎ 2R+ . decrease with distance, from P (r r1) 0.75–0.85 at t t the cluster center to Ӎ0.38–0.40 at the cluster periph- ε ε Thus, the ( , ζ, l) distributions of the cluster stars ery. At t ≥ 1.8 × τ , two regions with low probability averaged over P are statistically indistinguishable from vr r ∈ ∆ Ӎ Ӎ + the corresponding instantaneous distributions, and the P (r rj) 0.51–0.53 form in the cluster, near r Rt Ӎ Ð ∈ ∆ probability that the differences between these two dis- and r Rt . P (r r1) remains high and is equal to tributions are purely random for all times ti considered Ӎ σ ∈ ∆ 0.85. The dispersions P (r rj) inside the cluster ≤ Ð Ӎ is P (r Rt ) 0.85–0.95. are 0.04Ð0.10 in this case. ∈ ∆ τ ∈ ∆ The probabilities P (r rj) for times t = (1.3–1.5) vr Note that, near the cluster center, P (r r1) and ≤ increase with distance, from P (r ∈ ∆r ) Ӎ 0.55–0.65 to P (r r1) are substantially lower for the distribution of 1 ε ε P (r ∈ ∆r ) Ӎ 0.90; further, they decrease to 0.65Ð0.75 stellar coordinates ( , ζ, l) than for the distribution of 3 ú ú Ӎ + Ӎ Ð stellar velocities (ξ , ηú , ζ ). This is due to the following near r Rt , increase to 0.85Ð0.94 near r Rt , and specific features of the stellar distributions in (ε, εζ, l) finally decrease to P (r ∈ ∆r ) Ӎ 0.5. P (r ∈ ∆r ) Ӎ 10 j ξú ηú ζú ε 0.60–0.65 at t > 1.5 × τ . They remain approximately and ( , , ) space. The distribution of near the clus- vr ter center differs appreciably from any of the ε distribu- < + + < constant at r Rt , increase with r in the interval Rt tions for the cluster stars in Fig. 4a, since it has a max- r < RÐ , reaching 0.85 at r Ӎ RÐ , and then decrease to imum near ε Ӎ –0.56 and contains a large fraction of t t stars with small ε < Ð0.5. According to Fig. 1, the mean Ӎ0.5 in the interval r ∈ ∆r . The σ (r ∈ ∆r ) values 10 P j 〈ε〉 for stars in this ε domain exhibit appreciable peri- are equal to 0.04Ð0.05 inside the cluster and increase to odic oscillations, which result in substantial variations Ӎ > Ð ε 0.1 at r Rt . of the distribution for stars near the cluster center dur- ∆ ing the period Pr. The distribution of the stellar velocity It follows that, for all rj intervals inside the cluster, ε ε magnitudes near the cluster center has a maximum near the ( , ζ, l) distributions of the cluster stars averaged | | Ӎ Ð1 over P are statistically equivalent to the corresponding v 0.38–0.40 km s and has only a comparatively r small number of stars with |v| > 0.7 km sÐ1, where peri- ∈ ∆ Ӎ instantaneous distributions with probability P (r rj) odic oscillations of 〈|v|〉 become appreciable. There- 0.6–0.9; i.e., this is the probability that the differences fore, the stellar velocity distribution near the cluster between the distributions can be attributed to random center undergoes much weaker variations over the ε fluctuations for all times ti in the period Pr. period Pr than does the distribution of the cluster stars; this is reflected in the comparisons of the (ε, εζ, l) Let us consider the comparison of the averaged (ξú , and (ξú , ηú , ζú ) distributions of the cluster stars averaged ηú , ζú ) stellar velocity distribution and the correspond- over Pr and the corresponding instantaneous distribu- ing instantaneous distributions. For all times t in the tions (see above). ≤ table, the probabilities P (r rj) primarily decrease The largest changes of the stellar velocity distribu- ≤ Ӎ ≤ Ð Ӎ τ with distance, from P (r r1) 0.85 to P (r Rt ) tion over the period Pr at t = (1.3–1.5) vr take place at + Ð ≤ Ӎ the cluster periphery, in the distance interval R < r < R 0.65, or decrease from P (r r1) 0.75–0.85 to 0.61Ð t t ≤ Ð Ӎ ∈ ∆ Ӎ 0.65 and then increase to P (r Rt ) 0.75. The typical [where P (r rj) 0.38–0.40]; this separates in the σ ≤ course of the cluster’s evolution at > 1.6 × τ into two dispersions P (r rj) are equal to 0.05Ð0.07. It follows t vr ∈ ∆ Ӎ that, considering the entire cluster, the probability that regions with low P (r rj) 0.5 and appreciable tem- the (ξú , ηú , ζú ) stellar velocity distribution averaged over poral variations of the velocity distribution near the Ӎ + Ӎ Ð Pr and the corresponding instantaneous distributions tidal boundaries of the cluster at r Rt and r Rt .

ASTRONOMY REPORTS Vol. 44 No. 5 2000 EQUILIBRIUM PHASE-SPACE DENSITY DISTRIBUTION 307 > τ 〈ε〉 〈ε 〉 〈 〉 The deviations of the instantaneous phase-space densi- (3) At t vr, the , ζ , and l values averaged ties from the phase-space density of the cluster model over several stars in intervals of ε, εζ, and l remain averaged over Pr corresponding to these temporal vari- approximately constant over most of the domain S of ations of the stellar velocity distributions are due to the (ε, εζ, l) space occupied by cluster stars at distances departure of the cluster model from virial equilibrium and r ≤ RÐ . Near the boundaries of and outside S, both the nonstationarity of the cluster in the regular field. The most t important temporal variations of the stellar distributions in number of stars and the stellar density are small, and ε ε 〈ε〉, 〈εζ〉, and 〈l〉 are not constant in time. In the cluster ( , ζ, l) space over the period Pr occur near the cluster cen- models considered, the dispersions σε, σε ζ, and σ of ter and near r Ӎ R+ . However, even here, the probabilities , l t the deviations of ε, εζ, and l from the corresponding ∈ ∆ 〈ε〉 〈ε 〉 〈 〉 P (r rj) are equal to 0.55Ð0.75; i.e., they are fairly high. average values , ζ , and l for stars occupying the Thus, this is the probability that the differences between (ε, εζ, l) intervals determined above have small (and the instantaneous distributions of the cluster stars in probably random) oscillations about certain constant ε ε > τ σ σ ( , ζ, l) space and the corresponding distributions aver- values throughout most of S at t vr. ε and l can aged over Pr are purely random. vary appreciably with time near the boundaries of and 〈ε〉 〈ε 〉 According to Fig. 4 and the results of our statistical outside S. The approximate constancy of the ( , ζ , 〈 〉 σ σ σ comparison of the stellar distribution in (ε, εζ, l) space l ) and ( ε, ε, ζ, l) values for stars in the domain S of (ε, εζ, l) space at t > τ indicates the formation of an averaged over Pr and the corresponding instantaneous vr distributions, the distribution (5) with the parameters equilibrium stellar distribution function f(ε, εζ, l) in the given in the table are statistically indistinguishable cluster. from the equilibrium distribution that develops in clus- (4) At t > τ , the probability that the instantaneous ter model 2, which is nonstationary in the regular field. i vr distribution function for the cluster stars in (ε, εζ, l) The linear form space is statistically indistinguishable from the corre- sponding distribution function averaged over the period Θ = Θ(r, v) = е ε Ð ε Ð β l Ð l Ð γ εζ (8) Pr for oscillations of the regular cluster field at all times ∆ ∈ ∆ Ӎ for (ε, εζ, l) in (5) is also constant over the period P and ti and in all distance intervals rj is P(r rj) 0.6–0.9. τ r varies only slightly over a time interval ~2 vr. We now fix In other words, this is the probability that the differ- the coordinates of the vectors r and v. If Θ = const, peri- ences between the stellar distributions in (ε, εζ, l) space odic variations of the cluster potential U(r) at the point averaged over Pr and the corresponding instantaneous r should produce oscillations (with period Pr) in the stellar distributions are random in all distance intervals stellar distribution function, which are detected via ∆r considered (r ≤ RÐ ). On the whole, the stellar dis- appreciable changes in the magnitudes of the stellar j t tribution function in (ε, εζ, l) space for the entire cluster velocities over the period Pr (Fig. 5). averaged over Pr is statistically indistinguishable from the corresponding instantaneous distributions with proba- 5. CONCLUSIONS ≤ Ð Ӎ bility P(r Rt ) 0.85–0.95, while the differences (1) The open-cluster models considered in this between them are random with the same probability for paper develop an equilibrium stellar distribution in all the times t considered in the period P . Our tech- ε ε Ӎ τ i r ( , ζ, l) space over a time t vr, which varies slowly at nique for computing a cluster-model distribution func- > τ ε ε ε ε t vr. The equilibrium in ( , ζ, l) space develops as a tion in ( , ζ, l) space averaged over Pr can be used to result of relaxation in the absence of virial equilibrium. determine the equilibrium stellar distribution function ε ε ε ε The equilibrium distribution function f( , ζ, l) for the f( , ζ, l) at the central time of the period Pr. stars in the cluster models considered can be written f(ε, εζ, l) ~ exp(Θ), where Θ is a linear form for the vari- (5) The instantaneous and averaged stellar distribu- tions in velocity space are statistically indistinguishable ables (ε, εζ, l), whose coefficients vary only slightly ≤ Ð Ӎ over a time interval ~2τ . The equilibrium stellar with probability P(r Rt ) 0.65–0.75 [i.e., somewhat vr ε ε phase-space distribution function F0(r, v) corresponding lower than for the ( , ζ, l)-space distribution]. In the dis- + Ð to the equilibrium in (ε, εζ, l) space and the regular cluster ≤ ≤ × τ tance interval Rt r Rt at t = (1.3–1.5) vr and near potential vary periodically (or quasi-periodically) with ± > × τ time. This equilibrium is probably possible due to an the tidal boundaries of the cluster r = Rt at t 1.6 vr, approximate balance of the stellar transitions between the probability that the instantaneous stellar velocity phase-space cells over time intervals equal to the oscil- distributions and the corresponding averaged distribu- lation period for the regular cluster field. tions are statistically indistinguishable (i.e., differ only ∈ ∆ Ӎ (2) The local time for cluster relaxation toward the randomly) is only P(r rj) 0.4–0.5 for all times ti in Ӎ × τ the period P . The temporal variations of the instanta- equilibrium state is tr 0.5 vr in the central regions r Ӎ × τ of both cluster models considered and tr 1.1 vr and neous phase-space density of the cluster stars over Pr Ӎ × τ tr 1.5 vr at the peripheries of models 1 and 2, are due primarily to deviations from the stellar phase- respectively. space density functions averaged over Pr at the cluster

ASTRONOMY REPORTS Vol. 44 No. 5 2000 308 DANILOV periphery and near the tidal boundaries of the cluster. 3. H. E. Kandrup, Astrophys. J. 500, 120 (1998). The oscillations of the phase-space density for cluster 4. S. Chandrasekhar, Principles of Stellar Dynamics (Univ. stars about the averaged phase-space density are due to Chicago Press, Chicago, 1942; Inostrannaya Literatura, the absence of virial equilibrium in the cluster and its Moscow, 1948). nonstationarity in the regular field. 5. D. Lynden-Bell, Mon. Not. R. Astron. Soc. 136, 101 (1967). 6. W. C. Saslaw, Gravitational Physics of Stellar and ACKNOWLEDGMENTS Galactic Systems (Cambridge Univ. Press, Cambridge, This work was supported by the Ministry of Science 1985; Mir, Moscow, 1989). of the Russian Federation (the “Astronomy” State Sci- 7. V. M. Danilov, Astron. Zh. 74, 188 (1997). ence and Technology program). 8. S. A. Kutuzov and L. P. Osipkov, Astron. Zh. 57, 28 (1980). 9. V. M. Danilov, Pis’ma Astron. Zh. 23, 365 (1997). 10. I. R. King, Astrophys. J. 67, 471 (1962). REFERENCES 11. D. W. Marquardt, J. Soc. Indust. Appl. Math. 11, 431 1. V. M. Danilov, Astron. Zh. 76, 93 (1999). (1963). 2. H. E. Kandrup, M. E. Mahon, and H. Smith, Jr., Astron. Astrophys. 271, 440 (1993). Translated by A. Dambis

ASTRONOMY REPORTS Vol. 44 No. 5 2000

Astronomy Reports, Vol. 44, No. 5, 2000, pp. 309Ð322. Translated from Astronomicheskiœ Zhurnal, Vol. 77, No. 5, 2000, pp. 357Ð372. Original Russian Text Copyright © 2000 by Fedorova, Bisikalo, Boyarchuk, Kuznetsov, Tutukov, Yungel’son.

Nonconservative Evolution of Cataclysmic Binaries A. V. Fedorova1, D. V. Bisikalo1, A. A. Boyarchuk1, O. A. Kuznetsov2, A. V. Tutukov1, and L. R. Yungel’son1 1Institute of Astronomy, Russian Academy of Sciences, ul. Pyatnitskaya 48, Moscow, 109017 Russia 2Keldysh Institute of Applied Mathematics, Russian Academy of Sciences, Miusskaya pl. 4, Moscow, 125047 Russia Received May 5, 1999

Abstract—We present a mechanism to take into account angular-momentum loss in binary systems with non- conservative mass transfer. In a number of cases, mass loss in the system can increase the orbital angular momentum of the stars. Including this mechanism in evolutionary models substantially expands the domain of stable mass transfer in binary systems. All observed cataclysmic binaries with known component masses fall within the calculated area for stable mass transfer. © 2000 MAIK “Nauka/Interperiodica”.

1. INTRODUCTION model for the evolution of a cataclysmic binary, we take Cataclysmic variable stars are close binary systems into account the loss of mass and angular momentum consisting of a low-mass main-sequence star that fills by the system during mass transfer. its Roche lobe and a white dwarf. The We pay special attention to the stability of the mass star (donor) loses matter through the region surround- transfer for various values of the donor mass M2 and ing the inner Lagrangian point L1. The white dwarf (accretor) accretes at least some of this matter via an donor-to-accretor mass ratio q = M2/M1. It is known accretion disk or accretion columns in polar areas (pro- that mass loss from a star results in violation of its vided the white dwarf possesses a strong magnetic hydrostatic and thermal equilibrium. Hydrostatic equi- field). librium is reestablished adiabatically over the dynami- cal time scale, while thermal equilibrium is reestab- The physics and evolution of cataclysmic binaries (and of low-mass X-ray systems, whose evolution is lished over the Kelvin time scale. The reestablishment of similar) have been actively studied since the end of the the star’s equilibrium is accompanied by variations in its 1960s (see, for example, [1Ð10] and references therein). radius, which depend on the convective or radial stability Շ Theoretical studies of cataclysmic binaries have been of the outer envelope. For stars with mass M M᭪ and a driven by the fact that their evolution is determined by deep convective zone, and also for white dwarfs, mass orbital angular momentum loss via gravitational-wave loss is accompanied by an increase of the star’s radius, radiation [11, 12] and magnetic stellar wind [13Ð16]. while, for stars with radiative envelopes, mass loss A number of studies (see, for example, [17Ð19]) have results in contraction. The mass transfer in a close considered the effect of angular-momentum loss due to binary system is unstable if, in the course of evolution, matter outflow from the system over the course of a cat- the donor consistently tends to expand beyond its aclysmic binary’s evolution, as well as redistribution of Roche lobe. This occurs when the radius of the donor momentum between the components and accretion R2 increases more rapidly (or decreases more slowly) disk. However, in the absence of gas dynamical calcu- than does the average radius of the Roche lobe.1 It is lations of mass transfer in binary systems, it was neces- also possible for the radius of the donor to increase sary to study these processes in parametric approxima- while that of the Roche lobe decreases. Thus, the situa- tions. tion is determined by the ratio of the derivatives of the Recent three-dimensional gas dynamical calcula- ∂ ∂ donor radius R2/ M2 and the average Roche-lobe tions of gas flows in cataclysmic binaries (see, for radius ∂R /∂M , which admits the possibility of unsta- instance, [20, 21]) have shown that, in the course of RL 2 mass transfer, an intercomponent envelope is formed ble mass transfer even for stars that contract as they lose around the binary, and a considerable fraction of the mass. matter lost by the donor leaves the system. In the Refining our understanding of the conditions for sta- present study, we perform a numerical analysis of the ble mass transfer has two interconnected objectives: evolution of cataclysmic binaries taking into account the results of these three-dimensional gas dynamical 1 The average radius of the Roche lobe is defined to be the radius of calculations. In accordance with these results, in our the sphere whose volume is equal to that of the Roche lobe.

1063-7729/00/4405-0309 $20.00 © 2000 MAIK “Nauka/Interperiodica”

310 FEDOROVA et al. (a) To understand the 10% of cataclysmic binaries The derivative of the Roche lobe radius can be expressed with known component masses for which any combi- ∂ ∂ ∂ ( ) nation of M and q should make stable mass transfer lnRRL ln A ln RRL/A 2 -∂------= -∂------+ ------∂------. (3) impossible under the standard assumptions about vari- lnM2 lnM2 lnM2 ations of the system’s angular momentum. We can see from (2) and (3) that the variation of the (b) To understand which detached systems consisting average Roche-lobe radius as a function of the decrease of a white dwarf and low-mass secondary can become cat- of the donor mass is determined by the variations of the aclysmic binaries at the beginning of their evolution (i.e., masses of the donor and accretor and the distance between for which of these systems stable mass transfer is possi- them, i.e., on the mass and angular momentum lost from ble), which is essential for modeling of the cataclysmic- the system. As a rule, only the orbital angular momen- binary population. A similar problem also arises, for tum of the system has been considered; i.e., the angular example, in studies of low-mass X-ray binaries. momentum is taken to be the sum of the angular In Section 2, we describe the main factors determin- momenta of the two stars, which are treated as material ing the evolution of cataclysmic binaries and the condi- points. Consequently, the angular momenta of the axial tions for stable mass transfer in binary systems. Rea- rotation of the components, of the accretion disk (if it sons to abandon conservative approximations are con- exists), and of matter streams in the system are usually sidered in Section 3. In Sections 4 and 5, we present a not included in the total angular momentum. The devi- model for the angular-momentum loss due to the non- ation of the donor’s angular momentum from that of a conservative character of the mass transfer, based on point, which can be substantial, is not taken into con- three-dimensional gas dynamical calculations of matter sideration [24]. This approach has been necessitated by flows in semidetached binary systems. In Section 6, we the extreme difficulty of taking all these factors into compare the results of calculations made for conserva- account in calculations. In the simplified case we con- tive and nonconservative evolutionary models. The sider here, the orbital angular momentum of a binary conditions for stable mass transfer in an evolutionary system with a circular orbit is model that is nonconservative with respect to mass and M M angular momentum are determined in Section 7. J = G1/2 A1/2------1------2------, (4) ( )1/2 M1 + M2 2. MAIN FACTORS DETERMINING THE EVOLUTION OF CATACLYSMIC BINARIES where G is the gravitational constant. In the present study, we consider dynamically stable It is currently widely believed that the evolution of mass transfer; i.e., we assume that, during mass trans- cataclysmic binaries is determined by the loss of fer, the state of the star can differ appreciably from ther- momentum from the system via gravitational-wave radi- mal equilibrium and the mass-transfer rate can exceed ation (GWR) and/or a magnetic stellar wind (MSW) appreciably a rate corresponding to the Kelvin time from the donor, as well as via the transfer of mass between the components. The standard model of close- –7 –1 scale ( Mú Ӎ 3 × 10 RL/M M᭪ yr , with the radius R, binary evolution supposes that the mass transfer itself luminosity L, and mass M of the star in solar units). We does not affect the momentum of the system and that its take the mass transfer to be unstable when the rate of influence is only indirect, via possible mass loss from the variation of the donor radius exceeds that for the effec- system and carrying away of angular momentum with tive radius of the Roche lobe: Rú > Rú . Using the the outflowing matter. Therefore, the variation of the sys- 2 RL tem’s orbital angular momentum can be written derivatives of radius with respect to the donor mass M2, we can formulate the mass-transfer stability condition dJ ∂J ∂J ∂J ----- = ----- + ----- + ----- . (5) ∂lnR ∂ln R dt  ∂t  ∂t  ∂t ζ ------2- > ------RL---- ζ GWR MSW loss * = ∂ ∂ = RL. (1) lnM2 lnM2 Let us consider the various components of (5). The average radius of the donor Roche lobe can be esti- (1) Angular Momentum Loss from the System via GWR mated, for example, using the interpolation formula of The variations of the system’s orbital angular momen- Eggleton [22]: tum due to the radiation of gravitational waves can be written [25] q2/3 ≈ ------RRL 0.49A 2/3 1/3 , 3 ( ) ( ) ∂ln J 32G M1M2 M1 + M2 0.6q + ln 1 + q ------= Ð------, (6)  ∂t  5 5 4 where A is the semimajor axis of the orbit. For q Շ 1, GWR c A the approximation of Paczynski [23] is more conve- where c is the speed of light. One characteristic feature nient: of GWR is the very strong dependence of its intensity 1/3 on the orbital semimajor axis and, accordingly, on the ≈ 2  q  RRL ------A------ . (2) orbital period. GWR is substantial for short-period sys- 4/3 1 + q Շ h 3 tems with orbital periods Porb 10 , since only in this

ASTRONOMY REPORTS Vol. 44 No. 5 2000 NONCONSERVATIVE EVOLUTION OF CATACLYSMIC BINARIES 311 case does the characteristic time for momentum loss of the coefficient C in (7). In accordance with the become shorter than the age of the Universe. results of [7], we adopted C = 3.0. (2) Angular Momentum Loss from the System via a (3) Angular-Momentum Loss by the System During MSW from the Donor Matter Outflow A mechanism for angular momentum loss via a As a rule, the possibility of mass loss from the sys- MSW from the donor was suggested in [13, 14]. If the tem has been treated as a parameter in theoretical stud- donor possesses a convective envelope and, accord- ies of the evolution of cataclysmic binaries and related ingly, a surface magnetic field, its intrinsic axial rota- systems; for example, this parameter could be varied in tion is inhibited by the magnetic stellar wind, and the order to reach consistency with observational estimates angular-momentum loss rate can be appreciable even of the cataclysmic-binary minimum period [1]. More when the mass loss rate is small. The subsequent syn- specific assumptions about the evolution of the system chronization of the donor axial rotation and the orbital were made only in the following cases. rotation, due to tidal interactions between the compo- (a) The rate of capture of matter by the accretor nents, results in a loss of orbital angular momentum (a white dwarf) is limited by the rate of hydrogen burn- –7 –6 from the system and the decrease of A. A semiempirical ing (~10 –10 M᭪/yr), while the donor mass-loss rate formula from [16], which is based on extrapolation of appreciably exceeds this limit, not exceeding, however, the dependence of rotation velocity on age for G main the Eddington limit for the dwarf, which is close to × –5 sequence stars found by Skumanich [26] to the K and M 1.5 10 M᭪/yr (see, for example, [28, 29]). The excess components of cataclysmic binaries is most frequently matter can be lost via stellar wind [30]. used to take this effect into account. The decrease of (b) Matter is lost in flares on the white dwarf (see, orbital angular momentum as a function of time given for example, [17, 31Ð33]).3 In both cases, it is usually by this relation is assumed that the specific momentum of the matter flowing from the system is equal to the specific ( )2 4 ∂ln J Ð28 2 M + M R momentum of the accretor. ------= Ð0.5 × 10 k CG------1------2------2 . (7)  ∂  5 To describe mass and angular-momentum loss from t MSW M1 A the system, it is convenient to use parameters character- Here, k is the gyroradius of the donor and C is a coeffi- izing the degree of nonconservation with respect to cient determined by comparing theoretical calculations mass [34] with observations. Based on the observed magnetic- ∂ Mú 1 M1 field decay for stars with masses smaller than ~0.3M᭪, β = Ð------= Ð∂------Spruit and Ritter [27] proposed that the total mixing Mú 2 M2 that occurs in these stars when their masses decrease to 4 this limit results in an abrupt “switch off” of the and angular momentum: dynamo mechanism responsible for the generation of ú ∂ ψ J J ln J the stellar magnetic field. At that time, the time scale for = ----/---- = ∂------, (8) angular-momentum loss determined by equation (7) is Mú M lnM shorter than the thermal time scale for a star with mass where M = M1 + M2. ~0.3M᭪, and the star is no longer in thermal equilib- rium. In this case, the star’s radius exceeds the radius of In this case, the loss of angular momentum carried a star with the same mass in thermal equilibrium.2 When from the system by outflowing matter assumes the form the action of the MSW carrying away angular momentum ∂J J from the system terminates, the rate of decrease of the ----- = (1 Ð β)Mú 2ψ----.  ∂t M orbital semimajor axis slows, the donor-star radius loss decreases to its equilibrium value, and the star shrinks In cataclysmic-binary evolution calculations, an within its Roche-lobe surface [27]. Since the binary con- equation for the orbital semimajor axis as a function of tinues to lose angular momentum via GWR, its compo- time is used rather than the time dependence of the nents continue to approach each other slowly and, after angular momentum of the system; the former equation some time, the main-sequence star again fills its Roche can be derived from the latter. Let us differentiate lobe. The further evolution of the system is determined expression (4) for the orbital angular momentum of the by the loss of angular momentum via GWR. 3 In the two last studies, the angular momentum loss due to interac- It is currently thought that termination of the MSW tions between the donor star and the envelope ejected by a nova after total mixing of the donor-star material can explain was also considered. the so-called gap in the observed orbital periods of cat- 4 The specific ψ angular momentum carried from the system in aclysmic binaries. The theoretical width of this gap ΩA2 units, where Ω is the angular velocity of the orbital rotation, compared to its observed value determines the choice is sometimes used instead: Jú 2 Jú J q M1M2 2 α = ----/ΩA = ----/--- = ------ψ, µ = ------. For a discussion of the reaction of stars to mass loss on various ú ú µ 2 M time scales see, for example, [5, 9]. M M (1 + q)

ASTRONOMY REPORTS Vol. 44 No. 5 2000 312 FEDOROVA et al. ú system and substitute the result into (5), making use of RRL. Since the variation of R2 depends on M2 and M2 , the “mass ratio q donor mass M ” diagram is useful the relation Mú 1 = –β Mú 2 . Then, the variation of the — 2 orbital semimajor axis can be written for determining the boundaries of the domain of stable mass transfer. dA ∂A ∂A ∂A ∂A In a number of cases, expression (3) can be simpli- ------ = --∂---- + --∂---- + --∂---- + --∂---- . (9) dt t trans t loss t GWR t MSW fied and analytical formulae for the derivative of the Roche lobe radius can be obtained. In particular, in the Here, the subscript “trans” reflects variation of A as a absence of angular-momentum loss due to GWR and result of mass transfer between the components. Note MSW, we can obtain the dependence of the distance ∂J between the components [the first term in (3)] on their that, if there is no “trans” term -∂---- in (5), the cor- mass ratio and the parameters β and ψ from (10) and (11): t trans ∂ ∂ln A 2ψ(1 Ð β)q + qβ + 2q2β Ð q Ð 2  A ------= ------. responding term --∂---- term is present in equation ∂ (14) t trans lnM2 1 + q (9), as naturally follows from the dependence of the For the second term (3), using (2), as suggested by orbital angular momentum on both the orbital semima- Paczynski [23], we obtain for R jor axis and the mass of the components. For the case RL ∂ ( ) of conservative mass transfer (with respect to mass and ln RRL/A 1 1q(1 Ð β) angular momentum), variation of the orbital semimajor ------∂------= -- Ð ------. (15) lnM2 3 3 1 + q axis is determined by the assumption that the mass transfer does not affect the orbital angular momentum From (14) and (15), we have for totally conservative of the system. During the mass transfer, matter with mass transfer without mass and angular-momentum specific angular momentum from the donor is carried to loss by the system (β = 1) the accretor, and eventually acquires specific momen- ∂ ln RRL 5 tum from the accretor. If the orbital angular momentum -∂------= 2q Ð --. (16) is constant and the donor is less massive than the accre- lnM2 3 tor, the excess angular momentum of the accreted mat- In so-called Jeans mass loss, there is no mass trans- ter should increase the orbital momentum, so that the fer in the system, but the donor loses matter via its stel- mass transfer increases the semimajor axis of the orbit. lar wind, which leaves the system, carrying away spe- If the donor is more massive, the deficit angular cific momentum from the donor (β = 0, ψ = 1/q). We momentum of the accreted matter should be taken from obtain in this case the orbital angular momentum, so that the mass transfer 1 decreases the semimajor axis. -- Ð q ∂lnR In our modeling of the evolution of a cataclysmic RL 3 -∂------= ------. (17) binary, we simultaneously calculated the evolution of lnM2 1 + q the donor and the time variation of the orbital semima- For the case of a stellar wind from the accretor, when jor axis of the binary. We did not take into account the all the matter lost by the donor flows onto the accretor but evolution of the accretor, except for variations in its some later leaves the system, carrying away specific angu- mass. Let us consider the parameters determining the lar momentum from the accretor (here, we must formally time evolution of the semimajor axis. Using the param- adopt β = 0 and ψ = q), we have eters β and ψ, we derive formulas for the components of equation (9), where mass is given in M᭪, distance in 2 5 ∂ 2q Ð q Ð -- R᭪, and time in : lnRRL 3 -∂------= ------. (18) ∂ lnM2 1 + q  A M2 Ð M1 β ú --∂---- = 2A------M2. (10) t trans M1M2 At any stage of evolution, a comparison of the rates of variation of the radii of the donor and the Roche lobe (3) ∂A 2M (1 Ð qψ) + M determines the type of mass transfer (i.e., stable or ------= Ð(1 Ð β)A------1------2 M . (11)  ∂t  M (M + M ) 2 instable). There are two methods for determining the loss 2 1 2 boundary of the domain of stable mass transfer. We can ( ) ∂A Ð9 M M M + M directly calculate the derivatives of the donor radius ------= Ð1.65 × 10 ------1------2------1------2---, (12)  ∂  3 with respect to mass for various ú and compare them t GWR A M2 with derivatives of the average radius of the Roche ∂ ( )2 4 lobe, which depend on M , M , and A, as was done, for  A × Ð7 M1 + M2 R2 1 2 --∂---- = Ð6.06 10 C----------- . (13) example, in [35]. In this way, we can delineate the area t MSW M1 A of the q – M2 diagram in which mass transfer is dynam- As we can see from (10)Ð(13), the mass ratio q is ically stable and identify areas in which mass transfer one of the main parameters determining the variation of should occur on various time scales: the Kelvin, nuclear,

ASTRONOMY REPORTS Vol. 44 No. 5 2000 NONCONSERVATIVE EVOLUTION OF CATACLYSMIC BINARIES 313 or momentum-loss time scale. Alternatively, we can M2/M᭪ M1 = 1.4M᭪ adopt as a condition for stable mass transfer that the 1.2 1 accretion rate immediately after the donor Roche lobe is 3 filled not to exceed the Eddington limit for dwarfs. This approach is justified by the fact that evolution calcula- tions often yield very high mass-loss rates immediately after the Roche lobe filling, after which Mú decreases 2 rapidly. In essence, this approach limits the dynamical 0.8 stability of the mass transfer but makes it possible for the star to lose mass on the thermal time scale.

3. DOMAIN OF STABLE MASS TRANSFER M1 = 0.15M᭪ IN A MASS-CONSERVATIVE 0.4 CATACLYSMIC-BINARY MODEL Figure 1 presents the boundaries for the domain of stable mass transfer determined using standard mass- conservative models for cataclysmic-binary evolution. Boundary 1 was found in [35], and boundary 2 was pre- sented in [36]. These are in good agreement in the area 0 0.1 0.2 0.5 1 2 3 4 of low donor masses but differ somewhat at large q = M /M masses. Since the technique used to calculate this 2 1 boundary was not described in [36], we can only sup- pose that these discrepancies result from differences in Fig. 1. Plot of donor-to-accretor mass ratio q versus donor mass M2. Cataclysmic binaries from [37] are marked by the codes used, the ways in which possible attenuation asterisks and circles (the circles denote less reliable data). of the donor’s MSW at masses exceeding ~1M᭪ was For some stars, errors are indicated. Cataclysmic binaries taken into account, and the methods used to calculate from [38] are marked with triangles. The dashed line 1 indi- the derivative of the stellar radius in this area. cates the boundary of the domain of stable mass transfer according to [35], and the dashed line 2 shows the same Figure 1 also plots cataclysmic binaries with known boundary according to [36]. The solid line 3 delineates the component masses from the catalog [37]. Among the domain of stable mass transfer derived in the present study. Շ 80 systems plotted, 11 cataclysmic binaries are in the The dashÐdot lines indicate the upper (M1 1.4 M᭪) and տ lower (M1 0.15 M᭪) limits for the mass of the accretor, domain of unstable mass transfer. For eight of these which is a white dwarf. 11 systems, the errors are known and marked in Fig. 1 with bars. We can clearly see that the location of these stars in the domain of unstable mass transfer cannot be explained solely by the errors in the component masses. mass β and angular momentum ψ. These parameters The cataclysmic binaries whose parameters Smith and can be estimated only through three-dimensional gas Dhillon [38] consider to be most accurately determined dynamical modeling of the mass transfer in cataclys- are also indicated in Fig. 1. For a number of systems, mic binaries. Such studies have been made for binary refined mass values are given in [38], which are some- systems with mass ratio q = 1/5 [20] and q = 1 [21]. what different from those in [37]. Nonetheless, even for In addition, to study the effect of q on the flow pat- these well-studied cataclysmic binaries, the situation tern, we made numerical simulations for a system with remains unchanged: Three of 22 systems fall in the for- q = 5. bidden area. To explain this result, we will consider a model for The calculations indicated that the mass transfer remains nonconservative for all q values, and the cataclysmic-binary evolution in which we assume mass β and momentum loss from the system. This assumption degree of nonconservation is ~ 0.4–0.6. The typical is grounded in the results of our previous three-dimen- flow patterns in three-dimensional numerical simula- sional gas dynamical calculations of matter flows in tions also indicate that matter leaves the system with close binary systems. substantial angular momentum. However, determining the degree to which the mass transfer is nonconserva- tive with respect to angular momentum proves to be 4. CALCULATION more difficult. Let us consider the equations determin- OF THE ANGULAR-MOMENTUM LOSS ing angular-momentum transfer in the system. Since RATE IN THREE-DIMENSIONAL GAS viscosity plays an important role in redistributing angu- DYNAMICAL MODELS lar momentum in the system, we must consider the The above analysis shows that evolution calcula- NavierÐStokes equations. From the stationary gas tions critically depend on two parameters: the degree of dynamical equations for the x and y components of the nonconservation of the mass transfer with respect to velocity in a rotating coordinate system (here, u = (u, v, w)

ASTRONOMY REPORTS Vol. 44 No. 5 2000 314 FEDOROVA et al. is the velocity vector, ρ the density, ᏼ the stress tensor, Φ where Σ is the boundary of the donor, Σ is the bound- Ω 1 Σ 2 the Roche potential, and the orbital angular velocity) ary of the accretor, and 3 is the outer boundary. By analogy with the mass flux from the system ∂u ∂u ∂u 1∂ᏼ ∂ᏼ ∂ᏼ  u----- + v ----- + w----- + ------x---x + ------x---y + ------x---z ∂x ∂y ∂z ρ ∂x ∂y ∂z  Mú = Ð ∫ ρu ⋅ dn, ∂ î Σ = Ð------+ 2Ωv , 3 ∂x we can estimate the angular-momentum flux from the system ∂v ∂v ∂v 1∂ᏼ ∂ᏼ ∂ᏼ  u------+ v ------+ w------+ ------y---x + ------y---y + ------y---z ∂x ∂y ∂z ρ ∂x ∂y ∂z  Jú = Ð F ⋅ dn (20) ∂î ∫ = Ð------Ð 2Ωu, Σ ∂y 3 and use this in our expression for the parameter speci- we obtain the equation for angular-momentum transfer fying the degree to which the evolution is nonconserva- ∂λ ∂λ ∂λ 1 tive with respect to angular momentum (8). u----- + v ----- + w----- + --((r Ð r ) × divᏼ) ∂x ∂y ∂z ρ CM z Note that, in our simulations, we solved the Euler equations for a nonviscous gas rather then the NavierÐ (( ) × ) = Ð r Ð rCM gradî z, Stokes equations. Accordingly, to calculate the integral (20) in the expression for the momentum flux density F where rCM is the radius vector for the center of mass of in the stress tensor ᏼ, only the isotropic term corre- the system, and the angular momentum λ (in the labo- sponding to the gas dynamical pressure was taken into ratory coordinate frame) is determined by the expres- account: ᏼαβ = Pδαβ. This substitution is completely sion justified, since viscosity does not play an appreciable role at the outer boundary of the domain where the inte- λ ( ) Ω(( )2 2) = x Ð xCM v Ð yu + x Ð xCM + y gral (20) is calculated. = (x Ð x )(v + Ω(x Ð x )) Ð y(u Ð Ωy). Applying expression (20) to the simulation of a sys- CM CM 1 ψ tem with q = /5 [20], we obtain the value ~ 6, which Let us write the equation for angular-momentum trans- corresponds to α = 0.83 (see the determination of α fer in divergence form: given above in Note 4), while we obtain ψ ~ 5 for a binary system with q = 1 [21], which corresponds to div(ρλu) + ((r Ð r ) × divᏼ) α = 1.25. This latter value is consistent with the results CM z α ρ(( ) × ) of Sawada et al. [39], who obtained = 1.65 for a binary = Ð r Ð rCM gradî z. system with equal component masses. ψ This yields an integral expression for variations in the However, applying these estimates to in our evo- angular momentum: lution calculations indicated that the angular-momen- tum loss rates in binary systems are so high that, in most binaries, the mass transfer quickly becomes ρλu ⋅ dn + ((r Ð r ) × ᏼdn) unstable and the donor mass-loss rate begins to increase ∫ ∫ CM z without bound. Apparently, it is not entirely correct to Σ Σ use formula (20) to estimate the angular-momentum loss in cataclysmic-binary evolution calculations, since ρ(( ) × ) the gas dynamical and evolutionary models do not fit = ∫ r Ð rCM gradî zdV = è. together well. The gas dynamical model does not take V into account variations of the stellar positions in time The value of Π describes angular-momentum varia- (the distance between the components is assumed to be tion due to the noncentral character of the force field constant), so that variations in the angular momentum determined by the Roche potential. The quantity of the gas due to the noncentral character of the field Π are not compensated by corresponding variations in the ρλ ᏼ ( ( , , )) F = u + r' r' = Ðy x Ð xCM 0 (19) angular momenta of the stars. λ A more general gas dynamical model taking into represents the angular-momentum flux density . As a account variations of the positions of the stars would result, we obtain for the stationary case enable correct estimation of the variations of the system angular momentum in the form of a momentum flux- ⋅ ⋅ ⋅ density integral for the flux through the outer boundary Ð∫ F dn = ∫ F dn + ∫ F dn Ð è, of the closed “donor + accretor + gas” system. Due to Σ Σ Σ 1 2 3 difficulties in the gas dynamical calculations, however,

ASTRONOMY REPORTS Vol. 44 No. 5 2000 NONCONSERVATIVE EVOLUTION OF CATACLYSMIC BINARIES 315 such estimations are possible only for a specific stage in These considerations enable us to write the momen- the life of a binary. Moreover, the gas dynamical results tum flux lost by the system in the form cannot be directly used in standard evolutionary models, 2 which do not take into account the presence of intercom- loss 2 1  2  M2 2  Fλ = η(Fλ Ð Fλ) = η Ω∆ Mú 2 Ð Ω ------A βMú 2 , ponent matter in the system. Therefore, in the nearest   M   future, an approach in which a simplified model for angular-momentum transfer in the system is developed where η is a parameter determining the fraction of on the basis of gas dynamical calculations is likely to be intercomponent-envelope momentum carried away by promising; the results of this simplified and parame- the matter leaving the system. Then, 1 Ð η is the frac- trized model can then be used in evolution calculations. tion of the angular momentum of the intercomponent envelope that returns to the system via tidal interac- tions. From this point on, we will assume that η = 1; 5. A SIMPLIFIED MODEL FOR ESTIMATING i.e., all the momentum of the intercomponent envelope THE ANGULAR-MOMENTUM LOSS RATE is carried away with matter leaving the system. IN SEMI-DETACHED SYSTEMS In this case, the specific angular momentum of the matter leaving the system is Gas dynamical modeling of mass transfer in 2 semidetached systems suggests that the outflow from Ω∆2 ú β M2 Ω 2 ú the donor’s surface passes through a quite small area M2 Ð ------ A M2 λ loss ú M near L1, so that the specific angular momentum of the loss = Fλ /Mloss = ------(1 Ð β)Mú 2 outflowing matter can be estimated as λ = Ω∆2, L1 ∆ | | or, in units of the specific angular momentum of the where = xc – x is the distance from L to the center L1 1 system (λ = ΩA2M M /M2), of the mass of the system. Accordingly, the angular- syst 1 2 ψ λ λ momentum flux from the donor surface is = loss/ syst 2 2 2 2  1   q  Fλ = λ Mú 2 = Ω∆ Mú 2. ( ) β ------(21) L1  f q Ð ------ Ð   2 1 + q 1 + q (1 + q) = ------, Since the mass transfer is nonconservative, only 1 Ð β q some fraction β of the matter flowing through L is 1 where f(q) = x /A is the dimensionless distance from accreted, and the specific angular momentum of the L1 accreted matter is equal to the specific angular momen- the donor center of mass to L1. tum of the accretor (neglecting the finite radius of the Note that (21) was derived assuming coaxial syn- accretor and/or the residual momentum of the matter, chronous rotation of the binary components. The angular- which is a reasonable approximation for cataclysmic momentum loss due to matter outflow through the vicinity binaries). In this case, the accreted matter possesses of L1 can make the intrinsic rotation of the donor star and zero angular momentum relative to the accretor in the the orbital rotation of the system become noncoaxial rotating coordinate frame. Consequently, it does not [40]. Although the flow structure in systems with spin up the accretor; i.e., the problem of the efficiency noncoaxial, asynchronous rotation was treated previously of the transfer of momentum from the axial rotation of in [41], we will not consider this effect here, due to the the accretor to the orbital motion of the system need not simplified form of the evolutionary models under study. be considered. As a result, the flux of angular momen- ψ tum onto the accretor is Figure 2 presents a plot of the relation (q) for var- ious degrees of nonconservation β. This figure also pre- ψ ψ 2 sents the values = 0 and = 1, marked with dashed 1 M2 2 lines. In the domain ψ > 1, the angular momentum carried Fλ = λ βMú 2 = Ω ------A βMú 2. accr  M  out per gram of matter leaving the system exceeds the average specific momentum of the system, resulting in a The expressions for the flux of angular momentum decrease of the binary’s specific momentum: from the donor to the accretor correspond to the general ∂  J  formula for the angular-momentum flux density F (19). ------< 0. As at the outer boundary, viscosity does not play an ∂t M important role near the surfaces of the donor or accretor, ψ where the matter has already lost its angular momentum In the domain 0 < < 1, the angular momentum carried completely, and the flow is radial in the rotating coordi- out per gram of matter leaving the system is smaller than nate frame. Consequently, in this case, the stress tensor the average specific momentum of the system, so that reduces to an isotropic pressure, which does not con- ∂  J  > tribute to the integral of the angular-momentum flux -------- 0. density. ∂t M

ASTRONOMY REPORTS Vol. 44 No. 5 2000 316 FEDOROVA et al. ψ β = 0.1 Y/A 15 1.5 β = 0.3 1.0

0.5 10 β = 0.5 α = 1 0 –0.5

–1.0 0 –1.5 1.5 –5 1.0 β = 0.9 β = 0.7 0.5 –10 0.1 1 10 0 q = M2/M1 –0.5 Fig. 2. Dependence of the degree of nonconservation with respect to angular momentum ψ on the mass ratio q for var- ious degrees of nonconservation with reference to mass β –1.0 [see (21)]. The dashed lines correspond to ψ = 0 and ψ = 1. The thick dashed curve shows the dependence ψ = (1 + q)2/q, cor- –1.5 responding to α = 1. –2.0 –1.5 –1.0 –0.5 0 0.5 1.0 1.5 2.0 2.5 X/A

As we can see from Fig. 2, for most cases of non- Fig. 3. Velocity vectors in the equatorial plane for three- conservative mass transfer (with the exception of β ~ 0), dimensional gas dynamical calculations for a binary with ψ M2/M1 = 1 : 5 (q = 1/5, top) and with M2/M1 = 5 : 1 (q = 5, becomes negative for some range of q. In this case, bottom). The position of the accretor is marked by an aster- the redistribution of angular momentum in the closed isk. The shaded area shows the donor, and the dashed curves “donor + accretor + gas” system, due to the outflow of its Roche equipotentials. The vector in the upper right cor- matter to infinity, increases the total momentum of the ner corresponds to a velocity 3AΩ. binary: Jú > 0. It is obvious that such “pumping” of angular momentum into the system can stabilize the the total angular momentum of the system. The typical mass transfer in cataclysmic binaries. For example, a flow patterns for systems with M2/M1 = 1 : 5 and M2/M1 = qualitative examination of variations in the orbital 5 : 1 (Fig. 3) indicate that, in the laboratory coordinate semimajor axis resulting from mass transfer indicates frame, the gas flowing out of the system rotates in dif- that, in conservative models, the mass transfer tends to ferent directions relative to the orbital motion (in both increase the semimajor axis only when q < 1. In “non- Figs. 3a and 3b, the orbital motion is counterclock- conservative” models, the interval of q in which both wise). Accordingly, in the solutions, the variation of the mass transfer and mass loss from the system tend to total angular momentum of the binary system displays increase the semimajor axis is considerably broader: 1 opposite signs: For q = /5, the momentum of the system When β = 0.5, this occurs for 0 < q < 2.8. decreases, while for q = 5, it increases. The dependence ψ(q, β) [see Fig. 2 and (21)] dis- A reasonable physical explanation for the different plays the following interesting peculiarity: The ψ(q) types of gas dynamical solutions obtained is provided curves for different values of β go through a common by the following qualitative picture. point. The position of this point corresponds to the case (1) The flux of matter leaving the system through when the center of mass of the system is exactly the vicinity of the outer Lagrangian point L is made up ψ 2 between L1 and the accretor, since, in this case, does of intercomponent gas whose angular momentum is β not depend on . Fortunately, this does not affect the sufficient for it to overcome the gravitation of the accre- solution. tor. The initial velocity of this matter (in the rotating Three-dimensional simulations of the gas dynamics coordinate frame) coincides with the direction of the of matter flows in cataclysmic binaries confirm that, orbital motion. depending on the component mass ratio q, the matter (2) The velocity of gas leaving the system varies outflow from the system can either decrease or increase depending on the gravitation of both components of the

ASTRONOMY REPORTS Vol. 44 No. 5 2000 NONCONSERVATIVE EVOLUTION OF CATACLYSMIC BINARIES 317

ζ ζ system, the centrifugal force, the Coriolis force, and the RL, * pressure force. Variations in the azimuthal velocity of the gas are determined primarily by the Coriolis and Cons Accr-wind gravitational forces, since the centrifugal force is cen- 1.0 trally symmetric and the deviation of the pressure from ζ = 0.6 central symmetry is also small. The action of the Cori- * β = 0.7 olis force deflects the flux in the direction opposite to β = 0.5 β the orbital motion. In turn, the initial flux of matter ζ = 0 = 0.3 0 * leaving the vicinity of L2 can be divided into two parts: ζ The first maintains the direction of its initial motion, * = –1/3 while the second, being deflected by the Coriolis force, assumes the opposite direction. The predominance of Jeans one or the other flux determines the final direction for –1.0 the gas motion in the stationary solution. (3) In a centrally symmetric gravitational field, the effect of the Coriolis force obviously cannot result in a solution where the direction of gas rotation does not –2.0 coincide with the orbital rotation in the laboratory coor- 0 1.0 2.0 3.0 dinate frame. The situation is fundamentally different q = M2/M1 when the field of the binary is not central. In this case, the gravitation of the donor star can additionally accel- Fig. 4. Dependence of the logarithmic derivative of the ζ erate the matter of the flow, resulting in a solution with Roche-lobe radius with respect to donor mass RL = the opposite direction of gas rotation in the laboratory ∂ ln RRL coordinate frame. --∂------on the mass ratio q (thick curves) for various lnM2 (4) It follows from (21) and Fig. 3 that there exists degrees of mass nonconservation β [(3) and (21)]; depen- an interval of q for which the gas in the system moves dence of the logarithmic derivative of the donor radius with opposite to the orbital rotation. This corresponds to the ∂ln R respect to ζ ------2-- on q (thin curves) for the case of case when the gravitation of the donor star becomes * = ∂ lnM2 sufficient to accelerate the gas to the extent required mass-conservative transfer [“cons,” equation (16)], a Jeans (the left boundary of the interval) but is not so high that mass-loss regime [“Jeans,” equation (17)], and a stellar it leads to inverse accretion back onto the donor (the wind [“accr-wind,” equation (18)]. The dashed curves indi- cate ζ values for totally convective and degenerate stars right boundary of the interval). For all other values of q, * the flux in the direction of the orbital rotation prevails. (ζ = Ð1/3), roughly solar-mass main-sequence stars in ther- * The presence of an additional interval of q where mal equilibrium (ζ = 0.6), and subgiants with degenerate * nonconservative mass transfer is accompanied by an low-mass helium cores (ζ = 0). increase in the momentum of the binary expands the * domain of stable mass transfer. This can be displayed with a q – ζ plot, frequently used to qualitative illustrate cific angular momentum from the donor or accretor. the problem of mass-transfer stability in cataclysmic ∂ ln R The derivative ζ ------2- for totally convective and binaries. Curves corresponding to analytical formulas * = ∂ for the derivative of the Roche-lobe radius with respect lnM2 ζ degenerate stars (ζ = Ð1/3), roughly solar-mass main- to the donor mass RL can be drawn, together with lines * corresponding to the derivative of the star’s radius with sequence stars in thermal equilibrium (ζ = 0.6), and respect to the mass ζ , which is known from studies of * * subgiants with degenerate low-mass helium nuclei their internal composition (see, for example, [35]). The (ζ = 0) are also plotted. We can see that, in accordance mass transfer will be stable for the interval of q where * the ζ curve passes below the line ζ for the donor star with (21), mass loss from the system is able to stabilize RL * the mass transfer from the donor over a substantially of the corresponding type. Such a plot is presented in broader interval of q than in the case of conservative trans- Fig. 4, which contains the derivative of the Roche-lobe fer. Note that, in the case of conservative mass transfer for ∂lnR a star with mass of the order of M᭪, the transfer time scale radius ζ = ------RL---- as a function of the mass ratio q RL ∂lnM is determined by the momentum-loss or nuclear time 2 Շ for the case when the momentum loss is described by scale for q 1.2 [35], but, in the case considered, the (21) for various values of the mass nonconservation corresponding boundary for q is also appreciably parameter β. Figure 4 also shows the dependences higher. ζ RL(q) for mass transfer that is completely conservative We note again that, in the case under study, there is with respect to mass and momentum and for cases a region of unstable mass transfer for small q. The insta- when matter leaves the system and carries away spe- bility—i.e., the failure to satisfy condition (1)—results

ASTRONOMY REPORTS Vol. 44 No. 5 2000 318 FEDOROVA et al.

. We calculated the evolution of the donor using a logM2 (M᭪/yr) modified version of a code designed for studies of low- –7 mass stars, used previously in [7, 10, 44Ð48]. The code uses the opacity tables compiled by Hebner et al. [49], supplemented at low temperatures with the tables of Alexander et al. [50], and the equation of state derived –9 by Fontaine et al. [51], with the corrections of Denisen- M1 = 1.0M᭪ kov [52]. We adopted nuclear reaction rates in accor- M2 = 1.0M᭪ dance with [53, 54]. –11 1 All calculations assumed that the donor fills its 2 Roche lobe immediately after the star reaches the zero- age main sequence. The donor mass was taken to be between 0.1 and 1.2M᭪, with various mass ratios q. We –13 assumed the donor had a chemical composition X = 0 0.70, Y = 0.28, Z = 0.02. In calculations for the convec-

M1 = 0.28M᭪ tive temperature gradient, we took the mixing length M2 = 1.0M᭪ parameter l/Hp to be 1.8. We used the method suggested –5 by Kolb and Ritter [55] to calculate the mass-loss rate 3 of the donor for a given donor radius and average 4 Roche-lobe radius. Based on the earlier results of gas dynamical calculations, when considering nonconser- –10 vative evolution, we assumed that the fraction of matter accreted by the dwarf β was 0.5 and that the momentum carried away by matter leaving the system was deter- mined by (21). We did not take into consideration pos- –15 h sible mass and momentum loss due to the ejection of 1 2 3 4 5 6 7 matter in flares. Porb Let us consider the evolution of cataclysmic bina- Fig. 5. Evolutionary tracks in the “orbital period–logarithm ries with stable mass transfer in a mass-conservative of the donor mass-loss rate” plane for systems with initial model. Figure 5 (a plot of orbital period vs. the loga- component masses M1 = 1.0 M᭪, M2 = 1.0 M᭪ (top) and rithm of the donor mass-loss rate) presents two tracks, M1 = 0.28 M᭪, M2 = 1.0 M᭪ (bottom). Tracks 1 and 3 one calculated in the “conservative” approximation (dashed) are “conservative,” and tracks 2 and 4 (solid) are “nonconservative.” (track 1) and the other in the “nonconservative” approx- imation (track 2). The initial masses of both the donor and accretor are 1.0 M᭪. We can see that there are no from the rapid increase of the ratio of the lost to the aver- fundamental differences between tracks 1 and 2. In the age specific momentum of the system [see (14) and (21)]. initial stage of evolution of the cataclysmic binary, the This can lead to the destruction of a donor with very predominant factor is the loss of momentum from the low mass.5 Thus, it is possible that some cataclysmic system via a MSW from the donor. As a result of this momentum loss, the orbital semimajor axis and period binaries that start their evolution as stable stars end with decrease in the course of the system’s evolution. With the catastrophic destruction of the donor when q ∂ decreases below some lower limit.  A the decrease of the donor mass and radius, --∂---- t MSW the MSW also decreases [see (13)], which, after some 6. EVOLUTION OF CATACLYSMIC BINARIES time, results in a gradual decrease of the donor mass- LOSING MASS AND MOMENTUM loss rate Mú 2 . Mú 2 is appreciably smaller for the non- conservative track than for the conservative track, since In order to take into account the results of three- ψ is negative at this stage of the evolution (Fig. 2), and dimensional gas dynamical calculations of mass trans- the total loss of orbital angular momentum in the non- fer in cataclysmic binaries and corresponding estimates conservative system is smaller than in the conservative of the loss of momentum from the system via matter system. At later stages of evolution, after q decreases outflow, we have studied the evolution of cataclysmic enough that the loss of momentum for the nonconser- binaries with various assumptions about the extent to vative evolution increases sharply, the situation is which it is conservative. reversed.

5 Another case of mass-transfer instability for small q is known, After the star becomes totally convective, the MSW brought about by low efficiency of the tidal interaction between from the donor terminates. The stellar mass for which total the accretion disk and orbital motion (see, for example, [42, 43]). mixing occurs depends on the extent of deviation from

ASTRONOMY REPORTS Vol. 44 No. 5 2000 NONCONSERVATIVE EVOLUTION OF CATACLYSMIC BINARIES 319 thermal equilibrium. Equilibrium models for main- (a) For cataclysmic binaries with donor masses sequence stars become totally convective at masses 0.25–0.4 M᭪, mass transfer starts immediately within ~0.36M᭪; stars in binary systems losing mass become the gap. totally convective at smaller masses, close to 0.25– (b) If the donor fills its Roche lobe at a later stage of 0.30M᭪ (for our assumptions about the chemical com- core hydrogen burning rather than at the zero age main position and opacity of the stellar matter). sequence, total mixing does not occur when its mass For the conservative track, mixing occurs when M2 = decreases to ~0.3 M᭪, due to the presence of a helium h core (see, for example, [46, 47]; note, however, that 0.265M᭪ (at this time, Porb = 3. 27), while, for the non- such systems are very rare). conservative track, mixing occurs when M = 0.249M᭪ 2 The stage of evolution after the period gap is char- h 6 (Porb = 3. 54). Thus, the donor mass and orbital period acterized by substantially lower rates of mass loss by corresponding to the upper boundary of the gap in the the donor, since the main factor determining the evolu- cataclysmic-binary periods differ little for the conser- tion is the radiation of gravitation waves by the system. vative and nonconservative tracks. In this case, the loss of momentum from the system is After the donor MSW ceases, the momentum loss substantially smaller than at earlier stages, when rate in the system decreases dramatically, and the momentum is lost via MSW from the donor. The other approach of the two components decelerates. As a characteristic feature of this stage of evolution is the result, the donor is no longer able to fill its Roche lobe, presence of a minimum period. This is related to the and its mass loss terminates. Since the radius of the increase in the degeneracy of the matter in the donor donor exceeds the radius of a main sequence star in ther- and the corresponding variation of the radiusÐmass mal equilibrium when the MSW ceases, the donor radius dependence of the star [56, 57]: When the donor mass decreases to its equilibrium value [27]. Consequently, decreases to ~0.05 M᭪, its radius begins to increase as the ratio of the radii of the donor and its Roche lobe its mass decreases, and the exponent of the massÐradius decreases even further. The only factor determining the relation tends to Ð1/3. The minimum period corre- evolution of the system during this detached stage of sponds to the exponent +1/3. The minimum periods for the conservative and nonconservative tracks are 75m evolution is GWR, which causes the stars to slowly m approach until the donor again fills its Roche lobe and and 81 , respectively. This discrepancy in Pmin is due to the mass transfer resumes. During the detached stage, differences in the total masses of the systems and in the the system is not manifest as a cataclysmic binary, so extent to which the donor radii deviate from the values that it “disappears” during some interval of periods. corresponding to thermal equilibrium, due to differ- This provides an explanation for the observed gap in ences in their mass-loss rates. After passing the mini- the periods of cataclysmic binaries. mum period, the orbital period begins to increase. At this stage, the rate of mass loss by the donor decreases The observed period gap is between 2 h. 1 and 3 h. 1 rapidly as the period increases. We can see from Fig. 5 [37]. The edges of the theoretical period gap depend on that the difference between the conservative and non- the rate of momentum loss via the MSW and the initial conservative tracks increases at this stage. This is due physical parameters of the evolution code, which deter- to the fact that the orbital angular-momentum loss rate mine the theoretical radii of the stars. The code we used rapidly increases as q decreases in the nonconservative overestimates the lower boundary to some extent. The case. edges of the gap for the conservative track were 2 h. 5–3 h. 3, The mass-transfer rates for the conservative and h nonconservative tracks become considerably different while those for the nonconservative track were 2. 4– after the passage of the minimum orbital period. The 3 h. 5. The discrepancy is due to the different rates of probability of observing a semidetached binary in a momentum loss before the gap and also to the different given interval of orbital periods is masses of the accretors after termination of the first γ semidetached stage of evolution. (ÐMú 2) p(logP) ∝ ------, (22) Note that a pronounced deficiency of cataclysmic Pú/P binaries is observed in the period gap, rather than their where γ is a positive constant. For apparent-magnitude total absence [37]. The presence of cataclysmic bina- γ ries in the gap can be understood as follows: limited samples, Ӎ 1 [58]. After the minimum period is passed, in the period 6 Since we are interested in the details of cataclysmic-binary evolu- h h tion related to possible nonconservation of mass and momentum, interval 2 –2. 5, the average probability for observing a we formally continued our calculations after the accretor reached cataclysmic binary on the nonconservative track is the Chandrasekhar mass, despite the fact that the evolution of real approximately a factor of 1.2 larger than the probability systems is interrupted at this time by thermonuclear explosion of for a binary on the conservative track. Thus, in spite of the white dwarf. In principle, our extension of the calculations for > ≈ the appreciable difference in mass-transfer rates, the M1 MCh 1.4 M᭪ is justified by the fact that, in reality, the accretion may spin up the accretor and that the critical mass for probabilities of observing cataclysmic binaries in the rapidly rotating dwarfs can substantially exceed MCh. two cases only slightly differ, since a nonconservative

ASTRONOMY REPORTS Vol. 44 No. 5 2000 320 FEDOROVA et al. system passes through the corresponding interval of domain of stable mass transfer are in good agreement periods more rapidly. with those based on comparing the derivatives of the Since the characteristic variability of cataclysmic donor and Roche-lobe radii. binaries depends on their mass-transfer rates, the differ- The new boundary for the stable mass-transfer domain (assuming nonconservative evolution) is pre- ences in Mú 2 predicted for conservative and nonconser- vative evolution could be reflected in the distribution of sented in Fig. 1. We can see that this domain includes cataclysmic binaries according to variability type. all observed cataclysmic binaries with estimated com- ponent masses. Thus, in the transition to a nonconser- Figure 5 presents tracks for cataclysmic binaries vative model for the evolution of cataclysmic binaries, that should undergo dynamically unstable mass trans- the problem of observed binaries being outside the fer in a standard mass-conservative evolutionary domain of stable mass transfer disappears. model. The initial masses for the donor and accretor are In the q – M2 diagram, all observed cataclysmic equal to 1.0 M᭪ and 0.28 M᭪, respectively. We have chosen such an extreme mass ratio purely for the pur- binaries are concentrated toward q Շ 1, whereas the pose of demonstration. nonconservative model predicts considerably higher q values. To study the origin for this discrepancy, we cal- Track 3 was calculated under the standard assump- culated a number of tracks for nonconservative systems tion of momentum loss only due to GWR and MSW. It init init represents a typical case of unstable mass transfer with with initial donor masses from M2 = 0.1 M᭪ to M2 unlimited increase of the mass-loss rate of the donor = 1.2 M᭪ in steps of 0.1 M᭪ and q values such that these and rapid approach of the mass-loss time scale to the tracks originate in the immediate vicinity of the new dynamical time scale. Strictly speaking, this can only boundary for the domain of stable mass transfer. These be considered a qualitative illustration of this type of tracks are presented in the q – M2 diagram in Fig. 6. We evolution, since the calculation did not take into used these tracks to derive the isochrones presented in account the formation of a common envelope in the Fig. 6, corresponding to times of 105, 106, 107, 108, 109, system, which should increase the orbital angular- and 1010 yrs, calculated from the onset of the mass momentum loss rate even more. transfer between the components.7 We chose this Track 4, calculated assuming a loss of mass and “extreme” group of tracks since these systems spend momentum from the system in accordance with (21), is the most time between the new and old boundaries for similar to tracks for the stable evolution of cataclysmic the stable mass-transfer domain. It is obvious that sys- binaries, with all their typical features. It differs from tems with smaller initial q, whose tracks originate to track 2 in the higher mass-loss rate of the donor in the the left of the new boundary, will spend a shorter time initial stage of evolution, due to the substantial increase there. Therefore, cataclysmic binaries with initial ∂A parameters close to these “extreme” values have the in the ------for large q [see (13)]. This results in a highest probability to fall in this area.  ∂t  MSW According to the locations of the isochrones in Fig. 6, larger deviation of the donor from equilibrium and, the tracks can be divided into two distinct groups: accordingly, in variations in the upper edge of the h (1) Tracks for systems with massive donors (0.4– period gap: This edge is at 4. 4 for track 4, whereas it is 1.2 M᭪), for which the specifics of evolution before the 3 h. 5 for track 2. The minimum period for track 4 is 76m, period gap are primarily determined by the loss of which is nearly the same as that for track 1, but some- momentum from the system via the donor’s MSW. what smaller than that for track 2. (2) Tracks with low-mass donors (0.1–0.3 M᭪), for which this stage is absent, since their donors are com- 7. BOUNDARIES OF THE DOMAIN pletely convective from the very beginning. OF STABLE MASS TRANSFER The tracks for the first group are characterized by a –6 –9 high (~10 –10 M᭪/yr) donor mass-loss rate before We determined the boundary of the domain of stable the period gap and, consequently, a rapid decrease of mass transfer in the non-mass-conservative model as the donor mass and the component mass ratio. This follows. We calculated a series of tracks for a given results in the rapid evolution of the system in the q – M2 donor mass with different q values in steps of 0.1. We diagram, so that the system passes through the section took the boundary to be the maximum q value for near the new boundary of the stable mass-transfer which the accretion rate onto the white dwarf remained domain over a time not exceeding 107 yrs. This explains below the Eddington limit in the course of evolution, the absence of observed cataclysmic binaries in the vicin- i.e., such that, when q was increased by 0.1, the accre- ity of this boundary. The evolution of tracks in the second tion rate exceeded the Eddington limit. Note that, upon further increase of q, an unlimited increase in the mass- 7 Note that the isochrone corresponding to 109 yrs merges with the loss rate of the donor arises in the calculated tracks 8 init ≥ when q becomes 0.2Ð0.3 larger than the limiting value. 10 -yr isochrone for tracks with initial donor masses of M2 Note that the results of this method for determining the 0.4M᭪, since both times fall in the period gap.

ASTRONOMY REPORTS Vol. 44 No. 5 2000 NONCONSERVATIVE EVOLUTION OF CATACLYSMIC BINARIES 321

M2/M᭪ M = 1.4M᭪ 1.2 1 1.2

1.1 1.0 1.0 105 0.9

0.8 0.8 106 0.7

0.6 0.6

107 0.5 M1 = 0.15M᭪ 0.4 0.4

0.3 108 0.2 0.2 9 10 0.1 1010 0 1 2 3 4 5 q = M2/M1

Fig. 6. Cataclysmic binary evolutionary tracks in the “qM2” plane for a nonconservative model (dotted; the number next to the origin of the track denotes the initial donor mass in M᭪). The thick curves are isochrones; the numbers near them denote time in years. The dashed line indicates the boundary for the domain of stable mass transfer according to [35]. The solid line indicates the correspond- ing boundary derived in the present study. Other notation is the same as in Fig. 1.

group, excluding a short initial stage with a high mass- time, a number of observed cataclysmic binaries have –10 transfer rate, is characterized by a low (~10 M᭪/yr) combinations of donor and accretor masses that are donor mass-loss rate, which remains nearly constant “forbidden” under the standard assumptions for varia- over an extended period of time (until the minimum tions of the angular momentum of the system, since the period is reached). Accordingly, in the course of evolu- mass transfer should be unstable in these cases. This tion, the donor mass and the component mass ratio vary problem is resolved by the model proposed in our very slowly. Consequently, systems starting their evo- study, which takes into account the loss of mass and lution near the new stable mass-transfer boundary angular momentum from the system, in accordance should, in principle, remain near it for a fairly long with the results of the gas dynamical calculations. We time. The absence of cataclysmic binaries in this area is have shown that the observations can be satisfactorily due to the fact that the masses of dwarfs in zero-age cat- explained if we estimate the momentum loss using an aclysmic systems cannot exceed ~0.15 M᭪—the mini- mum mass for the helium core of a star leaving the main approximation in which the specific angular momen- sequence. All the observed cataclysmic binaries in Fig. 6 tum of the matter flowing out of the system is deter- are located in the area delineated by the lines corre- mined by the difference between the specific momen- sponding to the maximum and minimum masses of a tum of the matter of the donor at the Lagrangian point white dwarf—1.4 and 0.15 M᭪. L1 and the specific momentum of the accretor. In this case, the fraction of matter lost by the donor that leaves the system can be of the order of 50%. It is important 8. CONCLUSION that, in our evolution calculations, the transition to a Results of three-dimensional gas dynamical calcu- nonconservative model does not result in appreciable lations indicate that, in the course of a cataclysmic variations of such evolutionary-track parameters as the binary’s evolution, a considerable fraction of the matter boundaries of the period gap or the minimum period of lost by the donor should leave the system. At the same the system.

ASTRONOMY REPORTS Vol. 44 No. 5 2000 322 FEDOROVA et al.

ACKNOWLEDGMENTS 29. L. Yungelson and M. Livio, Astrophys. J. 497, 168 (1998). This work was supported by the Russian Founda- tion for Basic Research (project no. 99-02-17619 and 30. M. Kato and I. Hachisu, Astrophys. J. 437, 802 (1994). 99-02-16037) and a Presidential grant of the Russian 31. L. Yungelson, M. Livio, J. Truran, et al., Astrophys. J. Federation (99-15-96022). 466, 890 (1996). 32. M. Livio, A. Govarie, and H. Ritter, Astron. Astrophys. 246, 84 (1991). REFERENCES 33. K. Schenker, U. Kolb, and H. Ritter, Mon. Not. R. 1. S. Rappaport, P. C. Joss, and R. F. Webbink, Astrophys. J. Astron. Soc. 297, 633 (1998). 254, 616 (1982). 34. A. V. Tutukov and L. R. Yungel’son, Nauchn. Inform. 2. S. Rappaport, F. Verbunt, and P. C. Joss, Astrophys. J. Astron. Soveta 20, 88 (1971). 275, 713 (1983). 35. A. V. Tutukov, A. V. Fedorova, and L. R. Yungel’son, 3. I. Iben, Jr., and A. V. Tutukov, Astrophys. J. 284, 719 Pis’ma Astron. Zh. 8, 365 (1982). (1984). 36. M. de Kool, Astron. Astrophys. 261, 188 (1992). 4. J. Patterson, Astrophys. J., Suppl. Ser. 54, 443 (1984). 37. H. Ritter and U. Kolb, Astron. Astrophys., Suppl. Ser. 5. R. F. Webbink, in Interacting Binary Stars, Ed. by 129, 83 (1998). W. H. G. Lewin, J. van Paradijs, and E. P. J. van den Heuvel 38. D. A. Smith and V. S. Dhillon, Mon. Not. R. Astron. Soc. (Cambridge University Press, Cambridge, 1985), p. 39. 301, 767 (1998). 6. U. Kolb, Astron. Astrophys. 271, 149 (1993). 39. K. Sawada, I. Hachisu, and T. Matsuda, Mon. Not. R. 7. A. V. Fedorova and A. V. Tutukov, Astron. Zh. 71, 431 Astron. Soc. 206, 673 (1984). (1994). 40. J. J. Matese and D. P. Whitmire, Astrophys. J. 266, 776 8. H. Ritter, in Evolutionary Processes in Close Binary (1983). Stars, Ed. by R. A. M. J. Wijers, M. B. Davies, and 41. D. V. Bisikalo, A. A. Boyarchuk, O. A. Kuznetsov, et al., C. A. Tout (Kluwer, Dordrecht, 1992), p. 307. Astron. Zh. 76, 270 (1999). 9. M. Politano, Astrophys. J. 465, 338 (1996). 42. M. A. Ruderman and J. Shaham, Nature 304, 425 (1983). 10. A. V. Fedorova, in Double Stars, Ed. by A. G. Masevich 43. P. Hut and B. Paczynski,« Astrophys. J. 284, 675 (1984). (Kosmoinform, Moscow, 1997), p. 179. 44. A. V. Fedorova and L. R. Yungelson, Astrophys. Space 11. R. P. Kraft, J. Mathews, and J. L. Greenstein, Astrophys. J. Sci. 103, 125 (1984). 136, 312 (1962). « 45. A. V. Tutukov, A. V. Fedorova, É. V. Érgma, et al., Pis’ma 12. B. Paczynski, Acta Astron. 17, 267 (1967). Astron. Zh. 11, 123 (1985). 13. E. Schatzman, Ann. Astrophys. 25, 18 (1962). 46. A. V. Tutukov, A. V. Fedorova, É. V. Érgma, et al., Pis’ma 14. L. Mestel, Mon. Not. R. Astron. Soc. 138, 359 (1968). Astron. Zh. 13, 780 (1987). 15. P. P. Eggleton, in Structure and Evolution of Close 47. A. V. Fedorova and É. V. Érgma, Astrophys. Space Sci. Binary Stars, Ed. by P. P. Eggleton, S. Mitton, and 151, 125 (1989). J. Whelan (Reidel, Dordrecht, 1976), p. 209. 48. A. V. Tutukov and A. V. Fedorova, Astron. Zh. 66, 1172 16. F. Verbunt and C. Zwaan, Astron. Astrophys. 100, L7 (1989). (1981). 49. W. F. Huebner, A. L. Merts, N. H. Magee, et al., Astro- 17. L. R. Yungel’son, Nauchn. Inform. Astron. Soveta 26, 71 physical Opacity Library: Los Alamos Sci. Lab. Rep. (1973). LA-6760-M (Los Alamos, 1977). 18. K. H. Kieboom and F. Verbunt, Astron. Astrophys. 395, 50. D. R. Alexander, H. R. Johnson, and R. L. Rypma, L1 (1981). Astrophys. J. 272, 773 (1983). 19. A. King and U. Kolb, Astrophys. J. 439, 330 (1995). 51. G. Fontaine, H. C. Graboske, and H. M. van Horn, Astro- 20. D. V. Bisikalo, A. A. Boyarchuk, V. M. Chechetkin, et al., phys. J., Suppl. Ser. 35, 293 (1977). Mon. Not. R. Astron. Soc. 300, 39 (1998). 52. P. A. Denisenkov, Nauchn. Inform. Astron. Soveta 67, 21. D. V. Bisikalo, A. A. Boyarchuk, O. A. Kuznetsov, et al., 145 (1989). Astron. Zh. 75, 706 (1998). 53. M. J. Harris, W. A. Fowler, G. R. Caughlan, et al., Ann. 22. P. P. Eggleton, Astrophys. J. 268, 368 (1983). Rev. Astron. Astrophys. 21, 165 (1983). 23. B. Paczynski,« Ann. Rev. Astron. Astrophys. 9, 183 54. G. R. Caughlan, W. A. Fowler, M. J. Harris, et al., At. (1971). Data Nucl. Data Tables 32, 197 (1985). 24. O. A. Kuznetsov, Astron. Zh. 72, 508 (1995). 55. U. Kolb and H. Ritter, Astron. Astrophys. 236, 385 25. L. D. Landau and E. M. Lifshitz, The Classical Theory (1990). of Fields (Pergamon, Oxford, 1975; Fizmatgiz, Moscow, 56. J. Faulkner, Astrophys. J. (Lett.) 170, L99 (1971). 1962). 57. B. Paczynski,« Acta Astron. 31, 1 (1981). 26. A. Skumanich, Astrophys. J. 171, 565 (1972). 27. H. C. Spruit and H. Ritter, Astron. Astrophys. 124, 267 58. U. Kolb, A. R. King, and H. Ritter, Mon. Not. R. Astron. (1983). Soc. 298, L29 (1998). 28. I. Hachisu, M. Kato, and K. Nomoto, Astrophys. J. 470, L97 (1996). Translated by K. Maslennikov

ASTRONOMY REPORTS Vol. 44 No. 5 2000

Astronomy Reports, Vol. 44, No. 5, 2000, pp. 323Ð333. Translated from Astronomicheskiœ Zhurnal, Vol. 77, No. 5, 2000, pp. 373Ð384. Original Russian Text Copyright © 2000 by Lamzin.

Intercombinational Line Profiles in the UV Spectra of T Tauri Stars and Analysis of the Accretion Zone S. A. Lamzin Sternberg Astronomical Institute, Universitetskiœ pr. 13, Moscow, 119899 Russia Received May 5, 1999

Abstract—We have analyzed for the first time profiles of the SiIII 1892 Å and CIII 1909 Å intercombinational lines in HST spectra of the stars RY Tau and RU Lup. The widths of these optically thin lines exceeded 400 km/s, ruling out formation in the stellar chromosphere. Since the intensity of the Si line exceeds that of the C line, it is unlikely that a large fraction of the observed line flux is formed in a stellar wind. The observed profiles can be reproduced in the framework of an accretion shock model if the velocity field in the accretion zone is appre- ciably nonaxisymmetric. In this case, the line profiles should display periodic variations, which can be used to determine the accretion zone geometry and the topology of the magnetic field near the stellar surface; corre- sponding formulas are presented. In addition, periodic variations of the 0.3Ð0.7 keV X-ray flux should be observed. © 2000 MAIK “Nauka/Interperiodica”.

1. INTRODUCTION cal line could be formed not only in the shock, but also in the stellar wind. The activity of classical T Tauri stars is usually interpreted as the result of protoplanetary disk accre- This problem can be substantially simplified if opti- tion onto a magnetized young star. It is thought that cally thin lines, which are obviously formed in the material from the inner regions of the disk is frozen in shock, are used to study the accretion-zone geometry. the star’s magnetic-field lines and then slides along the The SiIII 1892 Å and CIII 1909 Å intercombinational field lines to the stellar surface. When the gas reaches a lines (ICL) observed in the UV spectra of classical velocity of ~300 km/s, it decelerates sharply in a shock T Tauri stars [5] probably provide one example of such and is heated; the radiation of this gas is responsible for lines. We will show that even single observations of the the observed line and short-wavelength continuum profiles for these lines make it possible to draw some emission. important conclusions. In this picture, the profiles and intensities of emis- sion lines should vary as a result of nonstationary accretion and/or of displacements of the accretion zone 2. OBSERVATIONAL DATA relative to the observer due to the star’s rotation. Both Among the large number of UV spectra of T Tauri regular and quasiperiodic variations of emission-line stars in the Hubble Space Telescope Archive Database parameters are observed in classical T Tauri stars (see, (http://archive.stsci.edu/hst/target_descriptions.html), for example, [1Ð3]). only two spectra clearly display SiIII 1892 Å and CIII Information about the shape and extent of the accre- 1909 Å intercombinational lines. These are the spectra of tion zone on the surface of a classical T Tauri star would RU Lup (z10t0109m) and RY Tau (zle70108t) obtained offer a unique possibility to investigate the topology of with the GHRS spectrograph on August 24, 1992, and the magnetic fields of young stars and interactions December 31, 1993, respectively. Each spectrum has a between the accretion disk and magnetosphere, making width of 40 Å, a resolution of Ӎ13 km/s, and consists of studies of the accretion-zone geometry very important. five consecutive exposures with a total duration of Ӎ1500 s. The application of Doppler tomography for this pur- The spectra were processed using the IRAF package pose using the rich observational data available for v2.11 (http://iraf.noao.edu/iraf) and STSDAS/TABLES optical lines [4] is very promising but still premature. code v2.0.2 (http://ra.stsci.edu/STSDAS.html), follow- The main problem is that this technique is based on ing the standard technique described in Section 36 of comparisons of observed profiles with those obtained the “HST Data Handbook” (http://www.stsci.edu/do- using corresponding theoretical models. However, all cuments/data_handbook.html). We adopted the calibra- the optical emission lines in classical T Tauri stars are tion files recommended in the Archive Database for each optically thick, and it is not known how the intensities star. Figure 1a presents the spectrum of RU Lup averaged of these lines depend on the parameters of the accreted over all exposures, and Fig. 1b shows the same spectrum gas or the angular coordinates in shock models. The sit- smoothed with a five-point sliding average. Figure 2 pre- uation is complicated by the fact that almost every opti- sents the same information for RY Tau.

1063-7729/00/4405-0323 $20.00 © 2000 MAIK “Nauka/Interperiodica”

324 LAMZIN

10–13, erg/s/cm2/Å

RU Lup

2

1

0

2

+ + + Fe Fe ] 1 Fe C++ Fe++ Fe+ S0 S0

Si++ 0

1890 1900 1910 λ, Å

Fig. 1. Spectrum of RU Lup obtained with the Hubble Space Telescope. See the text for more details.

For RU Lup, only the SiIII 1892.03 Å line can be 1897.55 Å line; however, this is not observed. Note that identified confidently. One characteristic feature of this emission around 1897.5 Å is also seen in the spectrum line is the presence of extended wings, with the short- of RY Tau (Fig. 2); however, the 1890.24 Å line is not wavelength wing almost twice as broad as the long- observed there. Nonetheless, we cannot exclude the wavelength wing: ∆V Ӎ 400 and 250 km/s, respectively. possibility that the FeII line contributes to the short- We suggest that the profile asymmetry is due to the wavelength wing of the SiIII] 1892 Å line in RU Lup. presence of blending emission lines at wavelengths This could be tested by searching for the 1939.70 Å line 1890–1891 Å. 2 2 0 ( P3/2– D3/2 transition), which, under the same condi- Among spectral lines of the most abundant ele- tions, should be roughly a factor of five more intense ments with a low excitation potential E < 4 eV, we i than the 1897.55 Å line. Unfortunately, the 1892 Å line believe that only the FeII 1890.24 Å line (Fig. 1b)— does not fall in the range of available spectrograms. 2 2 0 which belongs to the b P3/2–w D5/2 multiplet, with Is it possible to relate the SiIII] 1892 Å line to matter Ei = 3.19 eV (see the Atomic Line List v2.01 database outflow from RU Lup? The profiles of the OI 6300 and (http://www.pa.uky.edu/peter/atomic))—is suitable for 5577 Å and the SII 4069 and 6731 Å forbidden lines in the spectrum of this star, indeed, have broad short-wavelength 2 2 0 this role. The FeII 1897.55 Å ( P3/2– D3/2 ) line, which wings extending to Ӎ300 km/s [7]. However, the wings of can be identified with a corresponding emission feature these lines from the red side the extension are a factor of in the spectrum of RU Lup (Fig. 1b), belongs to this five smaller than in the case of the SiIII] 1892 Å line. In same multiplet. Using the values of Aij for FeII [6], we addition, in the spectra of Herbig–Haro objects—i.e., in find that, if the upper levels of the multiplet are popu- regions of stellar wind, where the lines of highly ionized lated in proportion to their statistical weights and the ions are formed—the CIII] 1909 Å line is an order of mag- lines are optically thin, the 1890.24 Å line should be nitude more intense than the SiIII 1892 Å line (see [8] and nearly an order of magnitude more intense than the references therein), in contrast to the situation observed in

ASTRONOMY REPORTS Vol. 44 No. 5 2000

INTERCOMBINATIONAL LINE PROFILES IN THE UV SPECTRA OF T TAURI STARS 325

10–14, erg/s/cm2/Å RY Tau

4

2

0

SiIII CIII 4

2

0

1890 1900 1910 λ, Å

Fig. 2. Spectrum of RY Tau obtained with the Hubble Space Telescope. See the text for more details.

RU Lup. We, therefore, conclude that the bulk of the from a temporary defect in the detector. At the same SiIII] 1892 Å line emission is not formed in the stellar time, these photodiodes were not marked defective in wind. On the other hand, the large width of the line service files for these spectra. Therefore, the nature of rules out the possibility that it is formed in a region sim- the emission featire near 1907 Å remains unclear. ilar to the solar chromosphere. The CIII] 1908.73 Å line can be identified with con- However, these arguments do not exclude the pos- fidence, but due to its low intensity, we can only say sibility that roughly 5% of the observed flux in the that the extent of its long-wavelength wing is roughly SiIII] 1892 Å line forms in material in the stellar wind, the same as that of the SiII] 1892 Å line, while its flux which primarily distorts the shape of the short-wave- is about an order of magnitude lower. This last fact length wing. In this connection, note the narrow peak dis- implies that the density of the accreted gas is much placed Ð280 km/s from the expected position of the larger than derived in [10] on the basis of IUE spectra CIII] 1909 Å line. It is tempting to interpret this peak with resolution 6 Å and is closer to 3 × 1012 cmÐ3 [9]. as a line associated with the stellar wind: The star of RU The relatively small signal-to-noise ratio of the RU Lup Lup is probably viewed nearly pole-on [9], so that a line spectrum also prevents accurate determination of the formed in a jet perpendicular to the disk should be dis- structure of the SiIII] 1892 Å line around its maximum; placed toward the blue. The only problem is that a fea- i.e., we do not know if it really displays two peaks or if ture with similar shape and intensity can be seen in the this is the effect of noise or the contribution of weak z10t0107m spectrogram, which was obtained at lines of other elements. 1622–1657 Å 25 minutes before the spectrogram under In connection with the question of faint lines, based consideration, and this feature falls on the same photo- on the above criteria, we suggest that the remaining diodes of the receiving matrix (342Ð345). Since no emission features in Fig. 1 can be identified as fol- other spectra of the star display any feature of this kind, 2 2 0 it is reasonable to assume that both features resulted lows: FeII 1895.69 Å (a H11/2–w G9/2 ); SI] 1900.29 Å

ASTRONOMY REPORTS Vol. 44 No. 5 2000

326 LAMZIN µ We can see that the lines have roughly the same intensity (a) and total widths exceeding 350 km/s, which, as in the case of RU Lup, makes their formation in the stellar wind or the chromosphere unlikely. However, the qual- ity of the spectrum makes it impossible to judge about the presence of other lines or the similarity of the pro- file shapes for the two ICLs.

3. METHOD FOR CALCULATING ICL PROFILES IN A SHOCK WAVE ω µ If we assume, based on the above discussion, that (b) the ICLs are formed neither in the stellar wind nor in the chromosphere, it makes sense to check whether their profiles can be understood as a consequence of accretion, given that accretion shock models can prob- ably adequately reproduce both the absolute and rela- tive intensities of ICLs in the UV spectra of other T Tauri stars [9]. In this case, the ICL radiation by the Si++ and C++ ions should form essentially completely at the shock front, in the region where the gas is highly ionized by X-rays from behind the shock front. (The ω basis for this and all following statements concerning (c) µ the shock structure is given in [12].) It is essential that these regions be small in extent (∆r Ӷ R∗) and that both the density and velocity of the infalling gas be virtually Ӎ constant (N0 and V0, respectively). Therefore, to model the ICL profiles, we need specify only the veloc- ity field as a function of the coordinates of points on the stellar surface. The region itself can be taken to be plane-parallel. Let us take an arbitrary point A at the surface of the accretion zone and calculate the flux dFλ reaching the from a region around this point with area dS in the Fig. 3. (a) Shape of “quasi-dipole” field lines for n = 1 (inner wavelength interval from λ to λ + dλ. If Iλ(γ) is the curve), 2, 4, and ∞ (straight lines); accretion zone geometry intensity of optically thin ICL radiation in the direction for (b) Model 2 and (c) Model 3. See the text for more 0 details. making an angle γ with the normal, then Iλ(γ) = Iλ cosγ, 0 where Iλ is the intensity of the radiation in the normal 4 3 3 0 3p P2–3p 4s5 S2 ; note that the second component of direction. Therefore, 3 0 λ 0 this doublet P1–5S2 with = 1914.70 Å should have Iλ(γ ) cosγdS Iλ dS dFλ = ------= ------, half the intensity, consistent with the observations); and d2 d2 6 8 0 FeII] 1901.77 Å (a D9/2–z G7/2 ). We especially note where d is the distance to the star. the possible presence of the FeIII 1914.06 Å line, Let us introduce spherical coordinates with their ori- 0 which corresponds to the 7S – 7P transition of the gin at the center of the star and the polar axis coinciding 3 3 with the star’s rotation axis. The position of a point A 3d54s configuration: The upper level of this transi- will be described by the polar angle θ and the azimuth tion could be excited due to absorption of photons ϕ 2 θ θ ϕ from the short-wavelength wing (more exactly, with angle , with dS = R* sin d d , where R∗ is the radius ∆V = Ð265 km/s) of the hydrogen Lα line. Judging of the star. Then, the total flux from the accretion zone from the presence of strong H2 lines in the UV spec- will be trum of RU Lup, which originate due to fluorescent 2 R 0 excitation by the Lα line [11], this is very likely. * θ θ ϕ Fλ = -----2- ∫∫Iλ sin d d . (1) In the studied spectrum of RY Tau, both intercombi- d national lines (SiIII 1892 Å and CIII 1909 Å) can be Due to the comparatively low temperature of the gas identified with confidence (Fig. 2; see also Fig. 8 below). in the line formation region (T < 2 × 104 K), the local

ASTRONOMY REPORTS Vol. 44 No. 5 2000 INTERCOMBINATIONAL LINE PROFILES IN THE UV SPECTRA OF T TAURI STARS 327

Fv

1

0

1

0

1

0 –1 0 1 –1 0 1 –1 0 1 V/V0

θ Fig. 4. ICL profiles for a coaxial quasi-dipole with n = 2 (Model 1). The left row of profiles corresponds to a zone with 1 = 10°, θ = 30° for inclination angles of the rotation axis to the line of sight (from top to bottom) i = 20°, 50°, and 70°; the middle row 2 θ θ θ θ corresponds to a zone with 1 = 30°, 2 = 50° and the right row to a zone with 1 = 50°, 2 = 70° for the same inclination angles. width of the ICL, which is determined by thermal (2) Divide the largest possible interval for variations motions, should not exceed 5 km/s, which is much lower in Vrad in the accretion zone (Vmin–Vmax) into a sufficient than the velocity of the infalling gas, V0 ~ 300 km/s. number of equal subintervals. Therefore, the local line profile can be represented to (3) Divide the zone of interest into cells of equal size δ ∆θ × ∆ϕ good accuracy as a function. In this approximation, and calculate the value of Vrad in each of them each point of the accretion zone emits towards the (see formulae (A.6) and (A.7) in the Appendix). observer monochromatic radiation at the wavelength − λ λ λ (4) For each interval Vi Vi + 1, sum the values for = 0(1 + Vrad/c), where 0 is the laboratory wave- θ θ ϕ sin for cells whose Vrad falls in the length of the ICL and Vrad( , ) is the projection of the given interval and which are situated in the hemisphere velocity of the infalling gas onto the line of sight at this facing the observer (see inequality (A.5) in the Appen- point. dix).

Let us assume that the velocity V0 and the gas den- Then, by dividing the resulting sums by their max- ima, we obtain ICL profiles normalized to unity as a sity N0 are the same at all points of the accretion zone (the accretion is uniform). In this case, the observed function of radial velocity. flux inside a finite wavelength interval λ–λ + ∆λ (or the ∆ corresponding velocity interval V–V + V) is propor- 4. RESULTS FOR ICL PROFILE MODELING tional to the area of the accretion zone in which the radiation in this wavelength interval originates. In order Let us consider several simple accretion zone geom- to determine the ICL profile, we adopt the following etries and the corresponding velocity fields. For T Tauri procedure: stars, the velocity of the infalling gas cannot apprecia- bly exceed 400 km/s, whereas the ICL in the spectrum (1) Specify the velocity field within the accretion of RU Lup is considerably broader. This implies that zone and the zone geometry. the accretion occurs in such an extended area that the

ASTRONOMY REPORTS Vol. 44 No. 5 2000 328 LAMZIN

Fv

1

0

1

0

1

0 –1 0 1 –1 0 1 –1 0 1 V/V0

Fig. 5. Same as Fig. 4 for a quasi-dipole with n = 4. effect of variation of the angle between the line of sight problem. When n = 2, we obtain a purely dipolar field, and the streamlines (due to the sphericity of the star) while for n ∞, the force lines become straight lines plays a substantial role. However, in this case, the (Fig. 3a). accretion flux cannot be in the shape of a quasi-cylin- drical gas stream occupying ~10% of the stellar sur- Let us first consider the case when the “quasi- face, as suggested in [13]. dipole” axis coincides with the rotation axis (Model 1). Due to the axial symmetry, the accretion zone should In the accretion shock, the SiIII] 1892 Å and CIII] θ ≤ θ ≤ θ π − θ ≤ θ ≤ have the shape of two belts 1 2 and 2 1909 Å lines are formed essentially at the stellar sur- π − θ ≤ ϕ π θ ϕ 1, with 0 < 2 , where , are polar coordinates face so that these lines can only be seen from points in related to the rotation axis of the star. It follows from the accretion zone in the hemisphere facing the observer. It follows that, if the gas falls inward along the calculations of [14] that the field of the star is the radius of the star, the radial velocity of visible emit- strongly distorted in the case of disk accretion onto a ting points should be positive, given that the lines are coaxial dipole; however, we are interested only in the optically thin and have very narrow local profiles. Thus, field in the immediate vicinity of the stellar surface, the presence of extended short-wavelength wings in the where the distortions are not so large. In fact, we chose observed ICLs implies that the velocity field of the the “quasi-dipole” approximation specifically to take infalling gas is appreciably nonradial. these distortions into account; for the sake of simplic- ity, we will assume that both the velocity and density of Let us suppose that the gas infall occurs along the the accreted plasma are equal at all points in the accre- star’s magnetic-field lines, whose shape near the stellar tion belts. Another specific feature stemming from the surface bears at least a qualitative resemblance to a calculations of [14] is that the field and velocity vector dipole field. Let the field lines be described in polar coor- Θ Φ Θ Θ n of the infalling material acquire a toroidal component, dinates r, , by the expression r = R∗(sin /sin ∗) , whose value reaches ~1/3 of the meridional compo- where Θ∗ is the polar angle at which a force line nent. In order to qualitatively take this into account, we crosses the stellar surface and n is a parameter of the will assume the presence of a toroidal component of the

ASTRONOMY REPORTS Vol. 44 No. 5 2000 INTERCOMBINATIONAL LINE PROFILES IN THE UV SPECTRA OF T TAURI STARS 329

Fv 0 0.500 1

0 0.125 0.625 1

0 0.250 0.750 1

0 0.375 0.935 1

0 –1 0 1 –1 0 1 V/V0

Fig. 6. Variation of the ICL profile shape in Model 2 as a function of rotation period phase, which is indicated for each curve. The observed profile of the SiIII] 1892 Å line from the spectrum of RU Lup is presented for phase ψ = 0.935. See the text for more details. infall velocity that everywhere constitutes a constant observer), since the other belt is obscured by the disk. ξ fraction of the meridional component V0. The discontinuities of the derivative visible in the cal- culated profiles are due to the assumption that the local In the Appendix, we outline a procedure for deriving ICL profile has a δ-function shape. the radial velocity in the quasi-dipole case. In the situ- We can see that the shape of the calculated profiles ation considered here, we calculated Vrad using (A.6) and (A.7) for θ ≡ θ, ϕ ≡ ϕ, and α = ψ = 0. Figure 4 differs qualitatively from the ICL profiles in the spectra m m of RU Lup and RY Tau. It is obvious that taking into presents the ICL profiles calculated using the technique ξ θ described in the previous section, for the dipole-field account the dependence of V0 and on will not enable geometry under consideration (n = 2). Inspection of the us to substantially improve the agreement between the top-to-bottom sequence of plots indicates that the ICL theory and observations. In other words, the origin for profile depends on the location of the accretion belts on the disagreement must be the initial hypothesis that the θ θ quasi-dipole and rotation axes of the star are coinci- the stellar surface ( 1, 2 = 10°, 30°; 30°, 50°; 50°, 70°); the left-to-right sequence of plots shows the depen- dent. dence of the profile on the angle i between the stellar We now assume that the magnetic axis makes an rotation axis and the line of sight (i = 20°; 50°; 80°). angle α with the rotation axis and that the field near the The solid curves show profiles calculated without tak- stellar surface resembles a quasi-dipole. Let us first ing into account the toroidal component of the velocity consider the situation when α is not so large that the (ξ = 0), and the dashed curves show profiles for the case difference between the parameters of the infalling gas ξ = 0.25. Figure 5 presents the analogous dependences at different points of the accretion zone must be taken of the ICL profiles on the parameters of the problem for into account. In this case, it is reasonable to suppose n = 4. We performed these calculations assuming only that the accretion zone is made up of two belts perpen- one accretion belt is visible (the one closer to the dicular to the quasi-dipole axis (Model 2; Fig. 3b). In a

ASTRONOMY REPORTS Vol. 44 No. 5 2000 330 LAMZIN

Fv 0 0.447 1

0 0.125 0.625 1

0 0.250 0.750 1

0 0.375 0.875 1

0 –1 0 1 –1 0 1 V/V0

Fig. 7. Same as Fig. 6 for Model 3 and the set of free parameters indicated in the text. The observed SiIII] 1892 Å line profile from the spectrum of RU Lup is shown for phase ψ = 0.447. coordinate frame fixed to the quasi-dipole axis, the line profile; however, the corresponding model cannot accretion region is determined by the conditions θ ≤ provide a sufficiently extended short-wavelength wing. θ ≤ θ π θ ≤ θ ≤ π θ ≤ ϕ m1π m m2 and – m2 m – m1, with 0 m < 2 . Figure 6 presents the expected variations of the ICL As before, we will take into account radiation only profile as a function of the phase of the axial rotation α θ θ ξ from a single accretion belt (the one closer to the for i = 18°, = 10°, m1 = 58°, m2 = 71°, = 0.15, and observer). n = 4. For phase ψ = 0.935, we have plotted the observed profile assuming V = 400 km/s. Note that, in The orientation of the accretion belt relative to the 0 both Model 1 and Model 2, the velocity field is sym- observer will periodically vary due to the movement of metrical about the magnetic axis. the magnetic pole on the surface of the star during its rotation. For this reason, the ICL profiles should What will happen if the angle between the magnetic undergo periodic variations, unlike the situation when and rotation axes of the star becomes fairly large, say, α = 0. Mathematically, this implies that the expression appreciably greater than 10°? We can see from Fig. 3b for the radial velocity V (formulas (A.6) and (A.7) in that, in this case, points with the same θ but different rad α ϕ m the Appendix) does not depend only on the angle , but m will be located at considerably different distances also on the rotation phase ψ. We calculated the varia- from the accretion disk. In this situation, it is very tions of the ICL profiles for the given geometry during unlikely that the accretion zone is made up of belts the rotation of the star for various values of i, α, θ , symmetrically placed relative to the magnetic axis, θ ξ α ≤ m1 m2, , and n. For 10°, we were not able to obtain inside of which the velocity and density of the infalling reasonable consistency between the calculated profiles gas are uniform. There do not exist any reliable calcu- and the ICL profiles in the observed spectrum of RY Tau. lations for the magnetic-field topology for the case of For RU Lup, we were able to obtain a better agreement disk accretion onto an inclined rotator that we could use between the calculated profiles and the SiIII] 1892 Å to calculate the ICL profiles for this case. Remaining in

ASTRONOMY REPORTS Vol. 44 No. 5 2000 INTERCOMBINATIONAL LINE PROFILES IN THE UV SPECTRA OF T TAURI STARS 331

Fv 0 0.500 1

0 0.125 0.675 1

0 0.250 0.750 1

0 0.375 0.875 1

0 –1 0 1 –1 0 1 V/V0

Fig. 8. Same as Fig. 7 for another set of free parameters. The observed SiIII] 1892 Å line profile from the spectrum of RY Tau is shown for phase ψ = 0.675. See the text for more details. the framework of the uniform-accretion hypothesis, we We stress that our ability to reproduce the observed can make the model more realistic by assuming that the profiles using Model 3 does not imply that this model, field topology near the stellar surface does not differ too indeed, provides an adequate description of the physi- much from the field of an inclined quasi-dipole with a cal situation. The consistency obtained only demon- toroidal component (Model 2) but that the accretion strates the fundamental possibility of interpreting zones are symmetrically placed relative to the rotation observed ICL profiles using shock models in which axis (Fig. 3c). This Model 3 represents a combination the velocity field in the accretion zone is apprecially of the two preceding models; however, in contrast to nonaxisymmetric. Models 1 and 2, its appropriateness for the situation at hand is far from obvious. 5. CONCLUSIONS Surprisingly, it was this model that was able to Our analysis of the SiIII] 1892 Å and CIII] 1909 Å reproduce the observed SiIII] 1892 Å line profiles. For line profiles in the spectra of RY Tau and RU Lup indi- RU Lup, we were able to reach a good agreement for cates that these lines cannot form in regions similar to α θ θ ξ i = 12°, = 18°, 1 = 56°, 2 = 72°, = 0.15, V0 = the solar chromosphere. It is also unlikely that a large 400 km/s, and n = 4. Figure 7 presents the observed fraction of the line flux is formed in shock waves due to profile for phase ψ = 0.447. Figure 8 presents a series interactions between the stellar wind and the surround- of calculated profiles for i = 36°, α = 18°, θ = 60°, ing medium (i.e., in HerbigÐHaro objects). θ ξ 1 2 = 80°, = 0.25, V0 = 280 km/s, and n = 4.25. We can If we assume that the SiIII] 1892 Å and CIII] 1909 Å see that, for ψ = 0.675, the theoretical profile approxi- lines are formed in an accretion shock, the observed mates the SiIII] 1892 Å line profile in the spectrum of profile shapes can be used to impose some restrictions RY Tau with reasonable accuracy, provided the two- on the character of accretion in low-mass young stars: humped profile shape is real. (1) The velocity field of the infalling gas is substantially

ASTRONOMY REPORTS Vol. 44 No. 5 2000 332 LAMZIN

Z ACKNOWLEDGMENTS

Zm ψ The author is grateful to the Hubble Space Tele- scope Science Institute for permission to access the P Space Telescope Archive. He also thanks A.I. Gomez α de Castro for useful discussions and comments and ϕ – ψ θ A.Yu. Zaœtsev for assistance with the IRAF-STSDAS M package, as well as with the Atomic Line List v.2 data- ϕ E m base. θ m

X O A APPENDIX Figure 9 presents a star whose center is marked by the O; the line OE indicates the direction towards the Earth. Let us introduce a rectangular (right-handed) T coordinate frame with its origin at O, its Z axis coincid- ing with the rotation axis of the star and its X axis in the plane determined by the lines OE and OZ. We will name EOP the inclination angle i. Let us relate the spherical coordinate system to this rectangular system Xm in the standard way, so that, for some point A on the stellar surface, the arc PA will serve as the polar angle Fig. 9. A scheme clarifying the derivation of formulas θ and the dihedral angle between the arcs EP and PA (A.1)Ð(A.7). See text for more detail. will be the azimuthal angle ϕ. The θ, ϕ coordinates of the point A obviously do not depend on time. nonradial, (2) the accretion zone must be appreciably α extended along the stellar disk, (i.e., accretion in the If the magnetic quasi-dipole axis forms an angle > 0 form of an isolated quasi-cylindrical stream is excluded), with the rotation axis of the star, the magnetic pole M will move on the surface of the star. In so doing, its and (3) the structure of the magnetic field directing the θ α infalling gas must be considerably nonaxisymmetric. polar angle M = does not vary, unlike the azimuthal ϕ ≡ ψ coordinate M of the dihedral angle between the This last circumstance suggests there should be arcs EP and PM (Fig. 9). Let us introduce another right- periodic variations of the profiles of line formed in the sided rectangular coordinate frame with its origin at O, shock, due to variations in the orientation of the accre- its Z axis directed along the quasi-dipole axis, and its tion zone relative to the observer. For the same reason, m Xm axis directed as shown in the diagram (Fig. 9), so the 0.3Ð0.7 keV X-ray flux from T Tauri stars should that ∠MOT = π/2. As in the previous case, let us relate also vary periodically. In this case, nonstationary accre- the spherical coordinate system with this rectangular tion will distort the periodicity of the X-ray flux varia- system. As we can see from Fig. 9, the large circular arc tions much less than it will the line-profile variations, θ MA is the polar angle m of the point A in the new coor- since the X-ray flux depends only weakly on the den- dinate system, while the dihedral angle between arcs sity of the infalling gas [15]. ϕ MA and MT is its azimuthal angle m. The coordinates θ ϕ Since only one spectrum is available for each star m, m vary periodically due to the rotation of the star and it is not of very high quality, it does not make sense about its axis. to try to reconstruct the velocity field and geometry of the accretion zone. We wished only to demonstrate that Let us derive relations between the coordinates θ, ϕ θ ϕ the observed ICL profiles could be reproduced in the and m, m of the point A. Applying spherical trigonom- framework of an accretion shock theory, which is not at etry to the triangles PMA, MTA, and PTA [16], we all obvious a priori. As a result, it is now clear that it is obtain the relations a reasonable goal to investigate ICL profile variations θ α θ α θ ϕ in the UV spectra of T Tauri stars, then apply Doppler cos = cos cos m Ð sin sin m cos m, (A.1) tomography to reconstruct the accretion zone geometry and velocity field. In the theoretical model, expression cosθ = cosαcosθ + sinαsinθcos(ϕ Ð ψ), (A.2) (1) should be used for the dependence of the intensity m on the angular coordinates, together with the relation and also between the intensity and N0 and V0 for the infalling gas from [9]. Naturally, the local ICL profile should be θ (ϕ ψ) θ ϕ described as a Gaussian rather than a δ function, whose sin sin Ð = sin m sin m  width is determined by the thermal motion of gas with cosαsinθcos(ϕ Ð ψ) = sinθ cosϕ + sinαcosθ. temperature Ӎ1.5 × 104 K. m m

ASTRONOMY REPORTS Vol. 44 No. 5 2000 INTERCOMBINATIONAL LINE PROFILES IN THE UV SPECTRA OF T TAURI STARS 333

After corresponding transformations, we obtain Assume now that, in addition to the meridional “quasi-dipole” component, the stellar magnetic field θ ϕ θ ϕ ψ ψ ϕ  sin sin = sin m sin m cos + f 0 sin also possesses a toroidal m component. In this case,  (A.3) θ ϕ ψ θ ϕ ψ the total velocity of the infalling gas can be represented sin cos = f 0 cos Ð sin m sin m sin , χ as a sum of meridional V0eM and toroidal V0eT compo- e ϕ α θ α θ ϕ nents, where T is the unit m vector for the point A and where f0 = sin cos m + cos sin mcos m, and ξ is the ratio of the amplitudes of the toroidal and meridional velocities. The projections of the vector e θ ϕ θ (ϕ ψ) T sin m sin m = sin sin Ð onto the axes of the rotating coordinate system are  (A.4) − ϕ ϕ sinθ cosϕ = sinθcos(ϕ Ð ψ)cosα Ð sinαcosθ. sin m, cos m, and 0. As above, we find that the com- m m ponent of the radial velocity relative to the gas motion Relations (A.1) and (A.3) make it possible to make in the toroidal direction and the total radial velocity are a unique transformation from coordinates in the rotat- equal to ing system to coordinates in the fixed system, while for- T mulas (A.2) and (A.4) provide the inverse transforma- V rad ξ( ϕ ϕ ) ------= f 2 cos m Ð f 1 sin m , tion. V 0 (A.7) The observer will see the point A if the angle M T between the vectors OE and OA does not exceed π/2. V rad = V rad + V rad . Since these unit vectors have coordinates (sini; 0; cosi) θ ϕ θ ϕ θ and (sin cos ; sin sin ; cos ), respectively, in the REFERENCES rectangular system OXYZ the condition that point A be visible can be written 1. E. Gullbring, P. P. Petrov, I. Ilyin, et al., Astron. Astro- phys. 314, 835 (1996). cos(∠EOA) = cosicosθ + sinisinθcosϕ ≥ 0. (A.5) 2. A. I. Gómez de Castro and M. Franqueira, Astrophys. J. 482, 465 (1997). Let the shape of the magnetic field lines of the star Θ 3. K. W. Smith, G. F. Lewis, I. A. Bonnell, et al., Mon. Not. r = r( ) be described in the “magnetic” coordinate sys- R. Astron. Soc. 304, 367 (1999). tem by the equation r = R(sinΘ /sinθ )n, where R is the θ m m 4. Y. C. Unruh, A. Collier Cameron, and E. Guenther, Mon. radius of the star and m is the polar angle of the point Not. R. Astron. Soc. 295, 781 (1998). A, where a line of force crosses the stellar surface. This 5. A. I. Gómez de Castro and M. Franqueira, ULDA Access field possesses only a meridional component. Let eM be Guide to T Tauri Stars Observed with IUE (ESA, Noord- a unit vector in the meridional plane, tangent to the wijk, 1997b). force line at the point A and directed towards the star. 6. S. N. Hahar, Astron. Astrophys. 293, 967 (1995). It is easy to show that the projections of this vector 7. F. Hamann, Astrophys. J., Suppl. Ser. 93, 485 (1994). onto the coordinate axes of the OX Y Z system are ϕ ϕ m m m 8. C. L. Imhoff and I. Appenzeller, in Exploring the Uni- equal to w1 cos m, w1 sin m, and w2, where w1 = verse with IUE Satellite, Ed. by Y. Kondo (Reidel, Dor- − θ 2θ 2θ (n + 1)sin2 m/w0, w2 = 2(sin m – ncos m)/w0, and drecht, 1989), p. 295. 2 2θ 2θ 1/2 w0 = 2[n cos m + sin m] . 9. S. A. Lamzin, G. S. Bisnovatyœ-Kogan, L. Érrico, et al., If gas moves near the surface of the star along mag- Astron. Astrophys. 306, 877 (1996). netic field lines, with the velocity of infall V0 equal at 10. A. I. Gómez de Castro and S. A. Lamzin, Mon. Not. R. Astron. Soc. 304, L41 (1999). all points of the accretion zone, then V = V0eM. At an M 11. S. A. Lamzin, Pis’ma Astron. Zh., 2000 (in press). arbitrary point A, the component of the velocity V rad 12. S. A. Lamzin, Astron. Zh. 75, 367 (1998). along the line of sight is EO á V. Expressing the coordi- 13. P. F. C. Blondel, A. Talavera, A. Tjin, et al., Astron. nates of the vector EO in the rotating frame using rela- Astrophys. 268, 624 (1993). tions (A.2) and (A.4), we obtain 14. K. A. Miller and J. M. Stone, Astrophys. J. 489, 890 M (1997). V rad ( ϕ ϕ ) 15. S. A. Lamzin, Pis’ma Astron. Zh. 25, 505 (1999). ------= w1 f 1 cos m + f 2 sin m V 0 (A.6) 16. K. A. Kulikov, A Course in Spherical Astronomy (Nauka, ( α α ψ) Moscow, 1978). Ð w2 cosicos + sin sinicos , α α ψ ψ where f1 = sin cosi – sinicos cos and f2 = sinisin . Translated by K. Maslennikov

ASTRONOMY REPORTS Vol. 44 No. 5 2000

Astronomy Reports, Vol. 44, No. 5, 2000, pp. 334Ð337. Translated from Astronomicheskiœ Zhurnal, Vol. 77, No. 5, 2000, pp. 385Ð389. Original Russian Text Copyright © 2000 by Knyazeva, Kharitonov.

G0ÐG5V Stars and the Sun in the uvby System L. N. Knyazeva and A. V. Kharitonov Fesenkov Astrophysical Institute, Kazakh Academy of Sciences, Kamenskoe Plato, Almaty, 480068 Kazakhstan Received April 30, 1999

Abstract—The color indices of the Sun in the uvby system are calculated using the spectral energy distribution of Lockwood, Tug, and White. This allows errors in the absolute calibration to be excluded from the calculated color indices. The normal position of the Sun on the (v – b)–(b – y) and (v – y)–(b – y) colorÐcolor diagrams for early G stars testifies to the absence of any significant peculiarities in the Sun compared to other stars of similar spectral type. These diagrams can provide a useful tool in searches for candidate solar analogues among faint stars. © 2000 MAIK “Nauka/Interperiodica”.

1. INTRODUCTION based on the observed solar spectral energy distribution and the response curves of this photometric system. The uvby photometric system [1] is very promising for searches for candidate solar analogues [2]. An extensive uvby catalog containing a large number of 2. SYNTHETIC COLOR INDICES stars with spectral types close to that of the Sun has The color indices of a star can be calculated in any been compiled [3]. This system was originally created photometric system if a reliable spectral energy distri- for AÐF stars, and proved to be a useful tool for studies bution (SED) and the response curves of the system are of stars of these spectral types. The b – y is available. A formula modeling photometric observa- similar to the B – V index of the UBV system (b Ð y is a tions in the system (either in the absence of or taking temperature parameter), m = (v – b) – (b – y) reflects into account atmospheric extinction) is used: λ1 the blanketing effect near ~ 4100 Å, and Ò1 = (u – v) – mi Ð m j (v – b) is a measure of the Balmer discontinuity. Later, (1) this system was extended to early G stars (G0ÐG5). ()()ϕλ ()λλ ⁄ ()ϕλ ()λλ = Ð2.5log ∫E* i d ∫E* j d + Cij. We are interested in the question of which indices or color indices in the uvby system could be useful in searches for candidate solar analogues. The Ò1 index is of little interest for this purpose, since it displays no corre- m1 lation with b – y [4] or other color indices for early G stars. 0.24 A comparison of the m1–(b – y) and (v – y)–(b – y) colorÐ color diagrams (Figs. 1, 2) for F0ÐG5 main-sequence 0.22 stars based on the data of [3] indicates that using v – b instead of m1 affects only the behavior of early F stars. 0.20 For late F and early G stars, v – b and m1 show identical b – y dependences. However, since the accuracy of a color index is higher than that of a quantity constructed 0.18 from the difference of two color indices, we prefer to use v – b. The v – y color index for early G stars also displays a good dependence on b – y (Fig. 3). These dia- 0.16 grams can be used in searches for candidate solar ana- logues once we determine the color indices for the Sun in this photometric system. 0.14 There are a number of indirect determinations of b – y for the Sun in the literature. Some of these are 0.12 based on the relationship between b – y and B – V; because of uncertainties in UBV for Sun, these have a 0.10 relatively large scatter, from 0.395m to 0.425m [5Ð9]. 0.10 0.15 0.20 0.25 0.30 0.35 0.40 0.45 Other indirect determinations are based on the depen- b – y dence of b – y on Teff [10Ð12]; these studies yield b – y = m m m Fig. 1. (b – y)–m1 colorÐcolor diagram for F0ÐG5 main- 0.407 , 0.402 , and 0.404 , respectively. We have cal- sequence stars. Crosses: F0ÐF4 stars; pluses: F5ÐF9 stars; culated the color indices of the Sun in the uvby system, triangles: G0ÐG5 stars.

1063-7729/00/4405-0334 $20.00 © 2000 MAIK “Nauka/Interperiodica” G0ÐG5V STARS AND THE SUN IN THE uvby SYSTEM 335

v – b v – y 1.15 0.65

0.60 1.10 51Peg 0.55 HD76151 16CygB 1.05 HD1835 HD44594 16CygA 0.50 1.00 0.45 HD53705 0.95 0.40

0.35 0.90

0.30 0.85 0.10 0.15 0.20 0.25 0.30 0.35 0.40 0.45 b – y 0.80 0.34 0.36 0.38 0.40 0.42 0.44 0.46 b – y

Fig. 2. (b – y)–(v – b) colorÐcolor diagram for F0ÐG5 main- Fig. 3. (b – y)–(v – y) colorÐcolor diagram for G0ÐG5 main- sequence stars. Crosses: F0ÐF4 stars; pluses: F5ÐF9 stars; sequence stars. Crosses: G0ÐG5 stars, open circle: the Sun; triangles: G0ÐG5 stars. triangles: stars from the list of solar analogues [12].

Here, E∗(λ) is the SED of the star in the interval cov- agreement with the observed indices (within 1Ð1.5%) in ϕ λ ϕ λ ered by the system response curve, i( ) and j( ) are all three systems. Using the data of [18] to fix the zero the response curves of the photometric system in the i point worsens this agreement. We conclude that the and j filters, and Cij = (mi – mj)H – (mi – mj)B is a con- SED for Vega presented in [15] better accounts for stant defining the zero point of the calculated color blanketing, which plays an important role when calcu- indices. This constant is determined using a star with a lating color indices, and precisely these data should be known SED, based on the difference between the used to establish the zero point. observed color index and the index for the photometric The table lists a comparison of the observed and cal- system considered calculated using the known SED culated color indices in the three photometric systems; and the first term of (1). n is the number of stars used, ∆ is the mean difference We estimated the accuracy of our calculated color between the observed and calculated color indices, and indices by comparing them with the observed indices σ Σ ∆ ∆ 2 = ( mean – i) /n. This table shows that the accu- for stars in three photometric systems: uvby [3], WBVR racy of the calculated color indices can be comparable [13], and the Vilnius system [14]. For this purpose, we to the accuracy of the observed indices if (1) the selected groups of single, nonvariable, main-sequence response curves for the photometric systems are well stars of spectral types F0ÐG5 from the catalog [15] for known, (2) the SEDs of the stars are known to no worse which observations in these systems are available. The than 2%, and (3) we have reliable data on the SED of the three lists of stars substantially overlap. The accuracy star determining the zero point of the calculated color of the SEDs of these stars is 1.5Ð2% from 4000 to 6000 Å, indices. The good agreement between the observed and falling to 2Ð3% at the violet and red edges. The calculated color indices of the stars in the three photo- response curves were taken from [16] (uvby), [13] metric systems confirms the trustworthiness of the (WBVR), and [17] (Vilnius). response curves used. All these considerations are also We fixed the zero point using two sources of data on relevant for calculation of the solar color indices. the energy distribution in the integrated spectrum of Vega [15, 18]. The absolute calibration at continuum points free of absorption was identical in both cases, 3. COLOR INDICES OF THE SUN but blanketing due to the Balmer lines was taken into IN THE uvby SYSTEM account differently. The resulting differences of the The solar SED can be derived from comparisons constants defining the zero point in the three photometric with standard sources, which are calibrated against a systems was 0.02m. The constants determined using the primary standard—a model blackbody. The procedure Vega data from [15] yield color indices that are in better for calibrating data for the Sun is rather laborous, and,

ASTRONOMY REPORTS Vol. 44 No. 5 2000 336 KNYAZEVA, KHARITONOV

Comparison of calculated and observed color indices in three Below, we present the calculated solar color indices photometric systems based on the data of [21] and the response curves [16]. UPXYZVS (n = 35) u – v = 0.994, ∆ b – y = 0.404, (UÐP)mean 0.015 σ{∆(UÐP)} 0.031 m1 = 0.217, ∆ c = 0.373, (PÐX)mean 0.003 1 σ{∆(PÐX)} 0.030 v – b = 0.621, ∆ (XÐY)mean 0.000 v – y = 1.025. σ{∆(XÐY)} 0.018 The constants were obtained for Vega using the SED ∆(YÐZ) 0.011 from [15], reduced to the absolute calibration of Tug et al. mean [23]. The calculated value of b – y is consistent with σ{∆(YÐZ)} 0.018 that derived from the dependence of Teff on b – y from ∆ (ZÐV)mean Ð0.015 [12], with Teff for the Sun taken to be 5777 K. If we can σ{∆(ZÐV)} 0.041 be confident of the reliability of the Teff – (b – y) rela- tionship derived in [12], this suggests that the relative uvby (n = 47) behavior for the solar SED obtained by Lockwood et al. ∆ [21] in the spectral region we have considered is close to (bÐy)mean Ð0.006 the true distribution. However, the error of b – y given in σ{∆(bÐy)} 0.013 [12] (0.005m) seems to be underestimated, and the good ∆ coincidence may be accidental. (vÐb)mean 0.001 σ{∆(vÐb)} 0.012 The Sun’s calculated color indices are shown in the colorÐcolor diagrams in Figs. 3 and 4, constructed for ∆ m1mean 0.007 G-star data together with data for candidate solar ana- σ{∆m1} 0.019 logues approved by the working group [2]. There is only one star near the Sun, HD 44594 (BS 2290), WBVR (n = 64) whose is equal to the solar value ∆ [12]. This suggests that the solar color indices we have (WÐB)mean 0.014 obtained are quite trustworthy. The Sun lies on both σ{∆(WÐB)} 0.034 sequences in a natural way, testifying to the absence of ∆(BÐV) Ð0.009 any significant peculiarities in the Sun compared to mean other early G stars. σ ∆ { (BÐV)} 0.015 The anomalous position of the solar color indices on the (U – B)–(B – V) colorÐcolor diagram is most likely associated with errors in the observed solar color indi- in many respects, the quality of the result depends on ces; these were derived from direct comparisons of the how reliably the energy scale has been transferred from Sun and other stars, require a large number of reduc- the primary standard to the Sun. When calibrating data tions, and depend on the choice of comparison stars. In for the Sun, different authors have used different stan- the WBVR system [24], we plotted solar indices calcu- dard sources calibrated in different laboratories, and no lated from the mean energy distribution of [25] on a mutual comparison has been performed. There are consid- colorÐcolor diagram constructed using the observa- erable discrepancies among the best known determina- tional data for other stars. The error of the zero-point shifts the Sun in one or the other direction in both coor- tions of the solar SED [19Ð22], which increase in the ultra- dinates. We discussed this problem in [26]. We calcu- violet (to 10%) and red (up to 6%), possibly due to the lated new solar color indices in the WBVR system using effect noted above. We decided to use the data of Lock- the procedure described above. We obtained the values wood et al. [21], since these were the only data for W – B = –0.025m and B – V = 0.656m, which shift the which it was possible to eliminate errors associated former position of the Sun on the (W – B)–(B – V) colorÐ with the absolute calibration. Lockwood et al. [21] cal- color diagram blueward. Note that the normal color indi- ibrated the solar spectrum against Vega, using the abso- ces of G2V stars calculated using the normal energy dis- lute energy distribution they had obtained earlier [23]. tribution for stars of this spectral subtype [27] are per- The advantage of these data is that by fixing the zero- fectly consistent with these values (W – B = –0.025m, point using data for Vega presented in the same energy B – V = 0.663m). scale as the Sun, we can exclude the error introduced by Based on the b – y color index for the Sun, we have the absolute calibration, since these errors will appear selected stars from catalog [3] that have b – y values in the two terms of formula (1) with opposite signs. close to the solar value (±0.005). This yields a rather

ASTRONOMY REPORTS Vol. 44 No. 5 2000 G0ÐG5V STARS AND THE SUN IN THE uvby SYSTEM 337

v – b ACKNOWLEDGMENTS 0.70 We express our sincere gratitude to V.M. Tere- shchenko for useful comments made during discus- sions of this work. 0.65 HD76151 51Peg HD1835 16CygB REFERENCES HD44594 16CygA 1. D. L. Crawford and J. V. Barnes, Astron. J. 75, 978 (1970). 0.60 2. Proc. Second Annual Lowell Fall Workshop, Ed. by J. C. Hall (Flagstaff, 1998). HD53705 3. B. Hauck and M. Mermilliod, Astron. Astrophys. Suppl. Ser. 86, 107 (1990). 4. V. Straœzhis, Multicolor Stellar Photometry (Mokslas, 0.55 Vilnius, 1977). 5. E. H. Olsen, Astron. Astrophys. 50, 117 (1976). 6. Y. Chmielevski, Astron. Astrophys. 93, 334 (1981). 7. T. Gehren, Astron. Astrophys. 100, 97 (1981). 0.50 8. A. Ardeberg, H. Lindgren, and P. E. Nissen, Astron. Astrophys. 128, 194 (1983). 9. D. F. Gray, Publ. Astron. Soc. Pacif. 104, 1035 (1992). 0.34 0.36 0.38 0.40 0.42 0.44 0.46 10. M. Saxner and G. Hammarback, Astron. Astrophys. 151, b – y 372 (1985). 11. P. Magain, Astron. Astrophys. 181, 323 (1987). Fig. 4. (b – y)–(v – b) colorÐcolor diagram for G0ÐG5 main- 12. G. Cayrel de Strobel, Ann. Rev. Astron. Astrophys. 7, sequence stars. Crosses: G0ÐG5 stars; open circle: the Sun; 243 (1996). triangles: stars from the list of solar analogues [12]. 13. V. G. Kornilov, I. M. Volkov, A. I. Zakharov, et al., Tr. Gos. Astron. Inst. im. P. K. Shternberga 63, 3 (1991). 14. V. Straœzys and A. Kazlauskas, Balt. Astron. 2, 1 (1993). long list of candidates, which we will analyze in a 15. A. V. Kharitonov, M. M. Tereshchenko, and L. N. Knya- future paper. We believe that, in combination with b – y, zeva, Spectrophotometric (Nauka, Alma- the v – b and v – y color indices can provide a good tool Ata, 1988). for searches for candidate solar analogues among faint 16. K. Kodaœra, in Problems in Stellar Atmospheres and G stars, as well as for constructing an effective-temper- Envelopes, Ed. by Baschek et al. (1975), p. 155. ature scale for stars of this type. 17. V. Straœzhis and K. Zdanavichus, Byull. Vil’n. Obs. 29, 15 (1970). 18. D. S. Hayes, in IAU Symp. 111: Calibration of Funda- 4. CONCLUSION mental Stellar Quantities, Ed. by D. S. Hayes et al. (Reidel, Dordrecht, 1985), p. 225. The main results of the work are the following: 19. H. Neckel and D. Labs, Solar Phys. 90, 245 (1984). 20. J. C. Arvesen, R. N. Griffin, and B. D. Pearson, Appl. (1) It is possible to obtain accuracies in calculated Opt. 8, 2215 (1969). color indices that are comparable to those of the 21. G. M. Lockwood, H. Tug, and N. M. White, Astrophys. J. observed magnitudes if we have (1) accurate response 390, 668 (1992). curves for the photometric system under consideration, 22. K. A. Burlov-Vasilijev, E. A. Gurtovenko, and Yu. B. Mat- (2) trustworthy SEDs for the stars studied, and (3) accu- vejev, in Proc. Workshop on Solar Electromagnetic Radia- rate data on the SED of the star used to determine the tion Study for Solar Cycle, Ed. by R. F. Donnelly (NOOAA zero point of the calculated color indices. This last ERL, 1992), p. 49. point, in particular, is a very important factor. 23. H. Tug, N. M. White, and G. M. Lockwood, Astron. Astrophys. 61, 679 (1977). (2) We have calculated color indices and the indices 24. L. N. Knyazeva and A. V. Kharitonov, Izv. Ross. Akad. Nauk, Ser. Fiz. 59, 176 (1995). m1 and c1 for the Sun in the uvby system by convolving its SED [21] with the response curves for this system. 25. E. A. Makarova, A. V. Kharitonov, and T. V. Kazachev- Using this solar SED allows us to eliminate errors in the skaya, The Solar Radiation Flux (Nauka, Moscow, absolute calibration from the calculated color indices. 1991). 26. L. N. Knyazeva and A. V. Kharitonov, in Proc. Second (3) The position of the Sun on the (v – b)–(b – y) and Annual Lowell Fall Workshop (Flagstaff, Arizona, (v – y)–(b – y) colorÐcolor diagrams for early G stars is 1998), p. 113. quite normal, testifying to the absence of any signifi- 27. L. N. Knyazeva and A. V. Kharitonov, Astron. Zh. 71, 458 cant peculiarities in the Sun compared to these stars. (1994). These diagrams can provide a useful tool for searching for candidate solar analogues among faint stars. Translated by G. Rudnitskiœ

ASTRONOMY REPORTS Vol. 44 No. 5 2000

Astronomy Reports, Vol. 44, No. 5, 2000, pp. 338Ð347. Translated from Astronomicheskiœ Zhurnal, Vol. 77, No. 5, 2000, pp. 390Ð400. Original Russian Text Copyright © 2000 by Lebedev, Presnyakov, Sobel’man.

+ Photodissociative Absorption by H2 in the Solar Photosphere V. S. Lebedev, L. P. Presnyakov, and I. I. Sobel’man Lebedev Physical Institute, Russian Academy of Sciences, Leninskiœ pr. 53, Moscow, 117924 Russia Received April 30, 1999

Abstract—Photoabsorption by systems of hydrogen atoms and protons in the solar photosphere is studied. + Analytical formulas for the partial cross sections for photodissociation of the H2 molecular ion are derived for the cases of fixed vibrationalÐrotational energy levels and averaging over a Boltzmann distribution for a given temperature. The photoabsorption coefficients for bound–free and free–free transitions of H–H+ in the solar photosphere are calculated. These are compared with the absorption coefficients for photo-ejection of an elec- tron from a negative hydrogen ion H– and free–free transitions of an electron in the field of a hydrogen atom H. Results can be applied to the Sun and hotter stars. © 2000 MAIK “Nauka/Interperiodica”.

2+Σ 1. INTRODUCTION transition between the even (bound) g and odd The largest contribution to the Sun’s radiation is 2+Σ + (repulsive) terms of the H2 ion, which are split by made by the negative hydrogen ion H–, whose density u –8 the exchange interaction between the hydrogen atom in the photosphere is of the order of 10 of the density and proton (see figure). We emphasize that precisely of neutral hydrogen atoms H [1]. Computations of the + spectrum and total intensity of the solar emission usu- H2 is of interest here, although the concentration of H2 ally assume that the optical depth of the photosphere at molecules is a factor of ~104–105 higher. The ground visual and near-infrared wavelengths is determined by 1+Σ 3+Σ photoabsorption via boundÐfree transitions, with the term g and first excited (repulsive) term u for these ejection of an electron from an H– ion, molecules have different multiplicities. Therefore, the intercombination transition between them is strongly H–(1s2 1S) + បω H(1s) + e (1) suppressed, since a spin must be exchanged. and via freeÐfree transitions Process (3), which describes photoabsorption via + H(1s) + e + បω H(1s) + e. (2) boundÐfree transitions of the H–H system, must be supplemented by freeÐfree transitions of the same sys- The effective cross sections for the photoabsorption tem: processes (1) and (2) have been calculated in a number + + of studies. The calculations in [2, 3] are probably the H1()Hs ++បω ()HH+ 2+Σ +បω most trustworthy. The accuracy of the results of [2] g (4) should be at least 1%. The cross section for photoab- + + ()HH+ 2+Σ H1()Hs + . sorption by an H– ion (1) is of the order of 10–17 cm2 in u the energy range 1 < បω < 5 eV and reaches a maximum As for photodissociation (3), the process (4)—absorp- σ ≈ × –17 2 + of max 4 10 cm when the photon energy is tion of a photon during a collision between H and H 1.4 eV. 2+Σ 2+Σ particles—takes place as a result of a g u tran- In connection with the accurate calculations [2, 3], sition, during which there is a change in the symmetry it is interesting to analyze in detail the role of other pos- of the electron wave function. sible photoabsorption processes. The contribution of such processes to the optical depth of the photosphere The possible role of boundÐfree and freeÐfree tran- at visual and near-infrared wavelengths could be as sitions (3) and (4) in the ç–H+ system in some stellar large as 1Ð10% of the contribution from (1) and (2). + atmospheres was first noted in [4]. H2 is the simplest + One such process is absorption by the H2 molecular diatomic molecule and has been the subject of many ion during its photodissociation: theoretical studies. Perhaps the largest contribution to + 2+ 2+ + 2+ +2+ + the theory of the H ion and Σ Σ radiative H ()Σ ++បω H()Σ H1()Hs (3) 2 g u 2 g 2 u transitions has been made by Bates and his coworkers with the formation of a proton and a neutral hydrogen ([5] and references therein). These studies were con- atom. A photon is absorbed as a result of an electron cerned with calculations of energy-level splitting for

1063-7729/00/4405-0338 $20.00 © 2000 MAIK “Nauka/Interperiodica”

PHOTODISSOCIATIVE ABSORPTION 339 the H+ system and of oscillator strengths for 2Σ+ U(R), eV 2 g 14 2Σ+ u radiative electron transitions, which were later also + 12 2Σ + σ ç2 calculated in [6]. u (2p u) The first quantitative treatment of processes (3) and 10 (4) was performed by Bates in [7], where he calculated 8 the integrated absorption coefficient K (ω) for a wide T 6 range of transition frequencies ω and temperatures T = 2Σ + σ g (1s g) 2500Ð12000 K. He derived a simple analytic formula 4 ω for KT( ) [7], based on quantum calculations of the oscillator strengths assuming a fixed internuclear dis- 2 + + ω ç(1s) + ç tance R and treating the H and H nuclei as classical 0 h particles, with the coordinates R distributed according ∝ π 2 to the probability WT(R) 4 R exp[–U(R)/kT] in the –2 bound Ug(R) and repulsive Uu(R) terms. By using a + –4 model with fixed H and H centers and a classical dis- 0 1 2 3 Rω 4 5 6 7 8 9 tribution function for the internuclear distance R in a R[a0] diatomic system with a given interaction potential U(R) (see, for example, [8], p. 167), Bates completely A diagram of light absorption via photodissociation of the avoided calculating the photodissociative absorption + cross sections for particular vibrationalÐrotational H2 molecular ion during an electron transition between its states vK and summing over these states. 2Σ+ 2Σ+ ground (even) g and first excited (odd) u terms. Later, the influence of the quantum character of the + nuclear motions during the photodissociation (3) of H2 បω (3) and (4) together with the integrated photoabsorption ions was studied in [9] for short wavelengths (2.7 < < coefficient, taking into account a realistic Boltzmann 14.3 eV) and relatively low temperatures (T = 1000 and distribution over the vibrationalÐrotational levels and the 2500 K), and the results for the absorption coefficients + were compared with the data of [7]. Further, the calcu- contribution of the continuum of the ç–H system. We lations of [7] were used to compile tables and derive also transform our results to simple analytic expressions approximation formulas for boundÐfree and freeÐfree corresponding to the semiclassical limit of Bates [7]. Spe- cial attention is paid to comparing the roles of these 2Σ+ 2Σ+ + transitions g u in the H2 ion [10]. processes and of photoabsorption by boundÐfree (1) Quantum-mechanical calculations of cross sections and freeÐfree (2) transitions in the ç–e system (i.e., the for photodissociation (3) from fixed vibrational levels v H– ion) in the photospheres of the Sun and hotter stars. were conducted in [11, 12] for the case of small rota- tional quantum numbers K = 0 and 1. The data of Bates cited above were used for the electron matrix elements 2. PHOTODISSOCIATION OF THE H+ ION 2Σ+ 2Σ+ 2 of the g u transition. In [12], the resulting AND FREEÐFREE TRANSITIONS cross sections were summed over v using a FranckÐ IN THE ç–H+ SYSTEM Condon distribution, which is characteristic of the for- + + 2.1. Equilibrium Density of H2 Ions mation of H2 ions during the photoionization of neu- + tral hydrogen molecules H2 or the ionization of H2 by The number density of H2 in the solar photosphere electron bombardment (the temperature of the H2 mol- is of the same order of magnitude as that of H– ions. The ecules was assumed to be low, equal to room tempera- law of mass action specifying the equilibrium condi- ture). The conditions assumed in calculations such as tions (see [8]) leads to the relations those of [11, 12] are obviously far from those in the solar photosphere. 3/2 N + g + πប2 H2 H2 2  (g) D0 The aim of the present paper is to analyze absorp------= ------------ Zv Ð r exp------ N N + g g + µkT kT + H H H H tion by molecular (quasi-molecular) H2 ions at visual, (5) 3/2 near-infrared, and near-ultraviolet wavelengths at fairly 8I  π  D  ≈ ------kT exp ------0 , high temperatures (T = 4500Ð10000 K). Under these Ω M  kT conditions, a large number of vibrational and rotational e p 2Σ+ levels of the ground electron state g are simulta- 2 3/2 N Ð g Ð E Ð H H  2πប   H  neously excited. We adopted a self-consistent approach ------= ------------ exp------ , (6) to calculate the cross sections of the radiative processes NH Ne gHge mekT kT

ASTRONOMY REPORTS Vol. 44 No. 5 2000 340 LEBEDEV et al.

g + = 2, g Ð = 1, gH = 2, [8], pp. 157Ð167). The equilibrium values of the rela- H2 H (7) tive number densities N + /NH N + and N Ð /NHNe cal- H2 H H g = 2, g + = 1. e H culated using (5) and (6) and the ratio N + /N Ð are pre- H2 H Here, g + and g Ð are the statistical weights of the H2 H sented in Table 1. We took the number densities of the + 2Σ+ – 2 1 electrons Ne and protons N + to be equal. The number electron states for the H2 ( g ) and H (1s S) ions; gH H is the statistical weight of a hydrogen atom in its 1s densities N are expressed in cm–3. ground state; m , N , g and M , N + , g + are the e e e p H H masses, number densities, and statistical weights of the 2.2. Cross-Section for Photodissociation + electrons and protons H , respectively; E Ð = 0.754 eV from an Excited VibrationalÐRotational State H – is the binding energy of the electron in an H ion; D0 = The figure shows potential-energy curves for the + ground 2Σ+ (1sσ ) and first excited (repulsive) 2.65 eV is the energy of dissociation of an H2 ion from g g the ground vibrationalÐrotational state v = 0, K = 0 2Σ+ σ + u (2p u) electron terms of H2 . In the case of large µ µ 2 ( = Mp/2 is its reduced mass); I = Re is the moment internuclear distances R ∞, these terms are corre- + of inertia; Re = 2.0a0 is the equilibrium internuclear lated with the H(1s) + H state of the isolated hydrogen × –8 atom and proton; i.e., they tend to the same dissociation distance (a0 = 0.529 10 cm is the Bohr radius); and Ω = 4.37 × 1014 s–1 is the oscillation frequency for the limit—the electron energy of the 1s level of the hydro- e gen atom. The minimum U (R ) = –D = Ð2.79 eV in the lowest level v = 0, K = 0 (see [13]). Expression (5) con- g e e 2Σ+ tains the vibrationalÐrotational statistical sum for the potential energy curve Ug(R) of the g state corre- 2 + + even ground term Σ (1sσ ) of H , which is equal to sponds to the internuclear distance Re = 2.0a0. The cor- g g 2 ∆ responding electron-term splitting energy is Uug(Re) = (g) 1 Z = --Z 11.84 eV. v Ð r 2 v Ð r 2Σ+ 2Σ+ ± (8) Therefore, g , v, K u , E, K 1 phototrans- ⑀ ( )2 1 ( )  vK ≈ kT itions from low vibrationalÐrotational levels vK (v = 0, = --∑ 2K + 1 expÐ------ ------ប----Ω-----, ≈ 2 kT 2Be e 1, 2) near the bottom of the potential well Rω 1.6–2.6 a0 vK of the Ug(R) term correspond to very large energies where Zv – r is the vibrationalÐrotational statistical sum (បω ≈ 8–16 eV) and small wavelengths (λ ≈ 800– for a diatomic molecule with arbitrary nuclei (which 1600 Å). In the visual (and the adjacent near-ultraviolet can, in general, be different from each other, as distinct and near-infrared, with photon energies 0.25 < បω < 6 eV; + ⑀ i.e., 0.2 < λ < 5 µm), the process (3) of photodissocia- from the H2 ion), vK is the energy of the vibrationalÐ + rotational level vK measured from the zero level v = 0, tive absorption by H2 is primarily realized from ⑀ ប2 K = 0 (so that vK > 0), and Be = /2I is the rotational excited levels vK. Phototransitions at these wave- constant in energy units. Relation (8) specifies the quan- lengths take place in the attraction region (Rω > Re) of បΩ Ӷ Ӷ tity Zv – r in the temperature range e/3 kT D0 (see the potential-energy curve Ug(R) of the lowest electron state at internuclear distances 3 < Rω < 7 a0. Table 2 pre- σ sents the potential energies Ug(R) in the 1s g ground + ∆ Table 1. Dependence of relative number densities of H2 state, the splitting energies Uug(R) for even (g) and – odd (u) terms, and the oscillator strengths fug(R) of the and H on temperature (T = 4500Ð10000 K) σ σ 1s g 2p u electron transition as functions of R for ≥ N + /NH N + , N Ð /NHNe, the above range of distances, R 3a0. H2 H H N + / N Ð T [K] H H 10Ð21cm3 10Ð21 cm3 2 The cross section for photodissociation from a given vibrationalÐrotational energy level vK of the ground 4500 3.20 2.34 1.37 2 + electron term Σ to the continuum of the repulsive 5000 1.70 1.68 1.01 g 2Σ+ ប2 2 µ 5500 1.02 1.24 0.82 term u with energy E = q /2 is determined by the 5800 0.79 1.06 0.74 expression [14, 15] 6000 0.67 0.96 0.70 π2ω ( , ) σph.d ω 4 max K' K (u, g) 2 6500 0.47 0.76 0.62 vK ( ) = ------∑ ------dEK', vK , 3c (2K + 1) (9) 8000 0.21 0.43 0.50 K' = K ± 1 បω 10000 0.11 0.25 0.45 E = Ð EvK .

ASTRONOMY REPORTS Vol. 44 No. 5 2000 PHOTODISSOCIATIVE ABSORPTION 341 | | ∆ Here, Ò is the speed of light, EvK is the bonding energy Table 2. Splitting of terms Uug(R), potential energy Ug(R) of + 2Σ+ 2Σ+ + of the H2 ( g ) molecular ion at the level vK, and q is the lowest state g of the H2 ion (measured from the dis- + ∞ ∞ the wave number of H and H particles separated by sociation edge Ug( ) = Uu( ) = 0), and oscillator strengths fug(R) ∞ R . We will measure the potential energies Ug(R) 2Σ+ 2Σ+ ∞ of the g u electron transition according to [5, 6] and Uu(R) from the ion dissociation edge; i.e., Ug( ) = ∞ Uu( ) = 0 and EvK < 0, whereas E > 0 (the vibrationalÐ ∆ R [a0] Ug, eV Uug, eV fug(R) rotational energy EvK of level vK is related to the ⑀ 3.0 Ð2.110 5.701 0.289 energy vK > 0 in (8), measured from the level v = 0, | | ⑀ K = 0, by the expression EvK = D0 – vK). 3.2 Ð1.925 4.925 0.281 (u, g) 3.4 Ð1.744 4.256 0.271 The matrix element d , of the dipole moment of EK' vK 3.6 Ð1.570 3.676 0.261 2Σ+ 2Σ+ the g , v, K u , E, K' transition in the radial 3.8 Ð1.407 3.173 0.250 χ wave functions (R) for the relative motion of the H 4.0 Ð1.254 2.736 0.238 + and H in the discrete spectrum of the initial bound 4.2 Ð1.113 2.356 0.226 2Σ+ + electron state g of the H2 ion and the continuum of 4.4 Ð0.984 2.027 0.213 2Σ+ 4.6 Ð0.866 1.744 0.201 the final repulsive state u takes the form: 4.8 Ð0.760 1.497 0.188 (u, g) χ(u) χ(g) dEK', vK = EK'(R) dug(R) vK(R) 5.0 Ð0.664 1.282 0.175 ∞ (10) 5.5 Ð0.469 0.865 0.144 (χ(u) )∗ ( )χ(g) 2 = ∫ EK'(R) dug R vK(R)R dR. 6.0 Ð0.326 0.580 0.116 0 6.5 Ð0.226 0.385 0.090 7.0 Ð0.152 0.253 0.070 Here, dug(R) is the electron matrix element for the tran- 2 + 2 + 7.5 Ð0.103 0.166 0.052 sition between the Σ Σ terms: g u 8.0 Ð0.070 0.109 0039 〈φ , φ , 〉 dug(R) = u(r R) er g(r R) 8.5 Ð0.047 0.070 0.028 (11) 9.0 Ð0.033 0.046 0.021 φ , ( )φ , = ∫ u*(r R) er g(r R)dr φ φ in the adiabatic wave functions g(r, R) and u(r, R) for tional levels vK, we should bear in mind that the main the electron in the axially symmetric two-proton field, contribution to the integral over internuclear distance (10) + + || H and H , where dug R (see [5] and [16], pp. 362Ð364) for a given transition frequency ω is made by a small and r is the electron radius vector measured from the + 2 + neighborhood of the point Rω where the H ( Σ ) + បω + 2 g center of mass of the H2 ion. + 2Σ+ and H2 ( ) potential-energy curves intersect; i.e., The electron-transition dipole matrix element (11) u បω is related to the corresponding oscillator strength Ug(Rω) + = Uu(Rω), f (Rω) of the transition at frequency ω = ∆U (Rω)/ប by (14) ug ug ∆ បω the expression Uug(Rω) = Uu(Rω) Ð Ug(Rω) = . ω 2me 2 Therefore, we can calculate the radial matrix element f (Rω) = ------d (Rω) . (12) ug 3 ប 2 ug of the dipole moment (10) using the quasi-classical e method of Landau ([16], pp. 399Ð401) or using an The nuclear wave functions in (10) are normalized as approach based on a quantum solution for the nonadia- follows: batic transitions in a model with linearly intersecting terms. This type of approach enables us to correctly ∞ 2 account for the contribution to the photodissociative χ(g) 2 ∫ vK(R) R dR = 1, ph.d ω absorption coefficient k + ( ), summed over all vK, of H 0 2 ∞ (13) all excited vibrationalÐrotational levels of the potential- energy curves (14) whose classical turning points R = a (χ(u) )∗χ(u) 2 δ( ) vK ∫ E'K(R) EK(R)R dR = E Ð E' . are both far from and near the intersection point Rω. 0 Exact solutions for the one-dimensional motion of When calculating the partial photodissociation nuclei in a uniform field (which, apart from normaliza- cross section (9) for the case of high vibrationalÐrota- tion coefficients, reduce to Airy functions î(x); see

ASTRONOMY REPORTS Vol. 44 No. 5 2000 342 LEBEDEV et al. [16], Para 24) should be taken for the nuclear wave func- In accordance with [16], we use the definition of the ( ) ( ) tions of the discrete χ g (R) and continuum χ u (R) . Airy function î(x) introduced by Fock (which differs vK EK from that presented in some handbooks by a factor of Thus, instead of the widely used quasi-classical for- π1/2). Note that, throughout the paper, we give only a mula (see [16], p. 401) for the transition probability at qualitative physical basis for the approximations used the point Rω, we use a more general method to calculate and the final formulas and will publish detailed calcu- the nonadiabatic transitions between two different elec- lations in a separate article. tron terms (this method is similar to that in Exercise 3 of Para 90 [16] in connection with collisions of the second Further, we use the well-known asymptotic of the kind). Applying this approach to radiative transitions Airy function î(x) in the classically permitted region + + ξ between the electron terms of the H2 ion leads to the fol- of motion for the H and H nuclei ( > 0), average the lowing expression for the cross section for photodisso- result (15) over the period of the rapidly oscillating part ciation from a given vibrationalÐrotational level vK: of the function î2(–ξ) in this region, and neglect its exponentially small decrease in the classically forbid- π2ω 2 ξ σph.d ω 8 dug(Rω) den region ( < 0). Then, we can see from (17) and (18) vK ( ) = ------∆------3cT vK Fug(Rω) that AvK 1 far from the inflection point (i.e., (15) for sufficiently large values of the energy difference Av × K 2 2 2 ------. Ev – U (Rω) – ប (K + 1/2) /2µRω ). In this case, a sim- ( ⁄ µ)[ ប2( )2 ⁄ µ 2 ] K g 2 EvK Ð Ug(Rω) Ð K + 1/2 2 Rω ple, quasi-classical expression for the cross section for πប | | + 2Σ+ Here, TvK = 2 / EvK – Ev ± 1, K is the period of the photodissociation of H2 ( g ) from a fixed vibrationalÐ vibrationalÐrotational motion of the nuclei in the state v ∆ rotational level K follows immediately from the quan- vK, and Fug(Rω) is the difference between the slopes tum formula (15). Note that this asymptotic of (15), of the potential-energy curves Ug(R) and Uu(R) at their which corresponds to AvK = 1, can also be derived point of intersection Rω: directly from the initial expression (9), if the results of Landau’s [16, pp. 399Ð401] quasi-classical theory are ∆ ∆ (K) dUu(R) dUg(R) Fug(Rω) = Fug (Rω) = ------Ð ------.(16) used to calculate the dipole matrix element for transi- dR dR R = Rω tions in the H–H+ system at the intersection point of the The quantity AvK in (15) can be expressed in terms of potential-energy curves. the square of an Airy function We can see from (15)Ð(18) that the quantum calcu- ξ 1/2 2 ξ AvK = 2 î (Ð ), lation of the partial photodissociation cross section differs ∞ from the quasi-classical calculation only at small values 3 (17) 2 2 2 1 u  for the difference [E – U (Rω) – ប (K + 1/2) /2µRω ]. In î(x) = ------∫ cos---- + ux dx, vK g π 3 particular, taking account of the factor A in (15) leads 0 vK to a finite value for the cross section at E – U (Rω) – where vK g 2 2 2 ប (K + 1/2) /2µRω = 0 and describes its rapid decrease 1/3 2µ  ប2(K + 1/2)2 in the classically forbidden region for nuclear motion. ξ ------ ω  =  2  EvK Ð Ug(R ) Ð ------2------ប  2µRω  (18) 2.3. Averaging over the Boltzmann Distribution ∆ 2/3 × Fug(Rω) , and the Photodissociative Absorption Coefficient ----(--K---)------(--K---)------Fu (Rω)Fg (Rω) A large number of vibrationalÐrotational levels vK are បΩ ӷ (K) (K) excited simultaneously at temperatures kT > e Be and Fg (Rω) and Fu (Rω) are the slopes of the even បΩ × –3 ( e = 0.288 eV, Be = 3.74 10 eV [13]), present in and odd terms taking into account the centrifugal the solar photosphere (kT ~ 0.5 eV). In this situation, if energy: we wish to find the photodissociative absorption coeffi- ph.d ω  2 2 cient k + ( ) (K) d ប (K + 1/2) H2 F (Rω) = Ð U + ------ , (19) g dR g µ 2  2 Rω ph.d ph.d R = Rω ω σ ω k + ( ) = T ( )N + , (21) H2 H2  2 2 (K) d ប (K + 1/2) it is of the most interest to calculate the cross section F (Rω) = Ð U + ------ . (20) u dR u µ 2  ph.d 2 Rω σ + (ω) averaged over v and K and normalized to the R = Rω H2

ASTRONOMY REPORTS Vol. 44 No. 5 2000 PHOTODISSOCIATIVE ABSORPTION 343 + γ total number density of H2 in the photosphere at tem- Here, (3/2, z) is an incomplete gamma function of perature T: order 3/2 (see, for example, [18]):

(vK) z N + 1/2 Ðt σph.d ω σph.d ω H2 γ(3/2, z) = t e dt, Γ(3/2) = π ⁄ 2. (24) T ( ) = ∑ vK ( )------∫ N + H 0 vK 2 (22) ph.d ph.d (2K + 1) Ev + D σ ω σ ω  K 0 The average cross section T ( ) determined by = ∑ vK ( )------expÐ------ , Zv Ð r kT formula (23) is a function of the frequency ω and gas vK temperature T and enables us, using (21), to calculate the photodissociative absorption coefficient for a given where the statistical sum Zv – r is determined from for- mula (8). + value of the total number density of H2 in the discrete When calculating the photodissociation cross sec- spectrum. During specific calculations, it is convenient σph.d ω to express the total number density of H+ in the bound tion T ( ) averaged over a Boltzmann distribution, 2 the summation over the vibrationalÐrotational levels vK 2Σ+ state g in terms of the number densities of neutral in (22) can be replaced by integration over dv and dK, + and carried out using a method similar to that of [17] hydrogen atoms H(1s) and protons H in the continuum This corresponds to replacing the discrete spectrum of using relation (5). Then, with relation (12) for the elec- the vibrationalÐrotational energy levels with a quasi- tron-transition oscillator strengths, the corresponding ph.d –1 continuum. result for the absorption coefficient k + (ω) [cm ] (21) H2 Further, numerical computations demonstrate that, for photodissociation (3) takes the form at temperatures T = 4500Ð10000 K and for all transi- tion frequencies ω in the range 0.25 < បω < 5.7 eV 2 2 ph.d 3  ប  Rω f ug(Rω) k + (ω) = 4π α ------under consideration, the energy interval EvK where the H   ∆ 2 me Fug(Rω) classical inflection points R = avK are near the intersec- (25) γ( , ⁄ ) tion point Rω of the potential-energy curves (14) (i.e.,  Ug(Rω) 3/2 Ug(Rω) kT × exp Ð------N N +, ប2 2 µ 2   Γ H H where the factor [EvK – Ug(Rω) – (K + 1/2) /2 Rω ] in kT (3/2) (15) is close to zero) does not make an important con- tribution to the integral over dv and dK. This enables us where α = e2/បc = 1/137 is the fine-structure constant. to restrict our treatment to a quasi-classical approxima- tion when calculating the averaged cross section σph.d(ω) [cm2] and the corresponding absorption coef- 2.4. Contribution of the Continuous Spectrum T and the Resulting Photoabsorption Coefficient ph.d –1 ficient k + (ω) [cm ]; in other words, we can use the H2 Let us consider now the contribution of (4) to pho- value AvK = 1 in (15) and (17) in the classically permit- toabsorption resulting from freeÐfree transitions of the ted region and neglect the exponentially small decrease ç–H+ system. The initial quantum expression for the σph.d ω absorption coefficient for collisions of H(1s) and H+ of the partial cross section vK ( ) in the classically forbidden region. The integration over the rotational particles during transitions from a continuum state with energy E for the initial term U (R) to a continuum state quantum numbers in (22) is conducted over the range g បω with final repulsive term Uu(R) with energy E' = E + ≤ ≤ ប2 2 µ 2 takes the form [14, 15] 0 K Kmax (where Kmax /2 Rω = EvK – Ug(Rω)); the + ( , ) integration over the energy of the H2 ion in the discrete frÐfr u g Ð1 k + (ω) = 〈Vσ , 〉 [ cm ] , ≤ ≤ H ÐH E' E T spectrum is conducted over the range Emin E 0 2 2 2 2 (26) (where Emin = Ug(Rω) < 0). ប q ប (q') E = ------, E' = ------, The calculation result obtained using the above 2µ 2µ approximations can be written

 g +  4 2 2 (u, g) 1 H 8π បω π3 ω µ 3/2 σ ω  2  ph.d 16 Rω dug(Rω)  kT  E', E ( ) = ------2--- σ (ω) = ------2g g + T ∆  2  H H  3cq 3c Fug(Rω)Zv Ð r 2πប (23) (27) γ( , ⁄ ) (u, g) 2 (u, g) 2  D0 + Ug(Rω) 3/2 Ug(Rω) kT × [( ) ] × ∑ K + 1 dE', K + 1; E K + K dE', K Ð 1; E K . expÐ------ ------Γ------. kT (3/2) K

ASTRONOMY REPORTS Vol. 44 No. 5 2000 344 LEBEDEV et al. ( , ) σ u g 4 Here, E', E [cm s] is the effective cross section for the coefficient for processes (3) and (4). In a quasi-contin- photoabsorption process (4), V = (2E/µ)1/2 is the rela- uum approximation, the general formula takes the form tive velocity of the colliding H(1s) and H+ particles at ω ph.d ω frÐfr ω kT( ) = k + ( ) + k + ( ) R ∞, and K is the orbital quantum number of these H2 H ÐH particles (in the quasi-classical limit, K qρ, where ρ = ប2 2 (31) is the impact parameter). The averaging in (27) is car- 3   Rω f ug(Rω)  Ug(Rω) = 4π α ------exp Ð------N N +.   ∆   H H ried out over a Maxwell distribution for the velocities V. me Fug(Rω) kT The matrix element of the dipole moment of the freeÐ free transition differs from (10) only in the exchange of This result has a clear physical interpretation: The the initial nuclear wave function of the discrete spec- absorption coefficient is proportional to the oscillator 2 + 2 + ( ) Σ Σ g strength fug(Rω) of the g u electron transition at trum χv (R) with the corresponding wave function of K the intersection of the terms (14). It is also proportional to χ(g) δ the continuum EK (R), normalized to an energy func- ∝ π 2 the probability WT (Rω) 4 Rω exp[–Ug(Rω)/kT] that the tion. + H(1s) and H particles are separated by a distance Rω in The factor of one-half in the braces in (27) is due to the initial bound state Ug(R). The Boltzmann factor the symmetry of process (4) with respect to permuta- exp(–Ug/kT) in (31) specifies the temperature depen- tion of the H and H+ nuclei, similar to the factor of one- dence of the total absorption coefficient. half in the statistical sum (8) for the discrete spectrum. + Formula (31) represents the total contribution of the Collisions of the H and H can occur with equal proba- direct processes (3) and (4) to the ç–H+ system photo- bilities via the two channels absorption coefficient. Taking into account stimulated ω emission, the resulting absorption coefficient KT( ) + (H + H )2Σ+ under conditions of thermodynamic equilibrium is H + H+ g (28) given by the well-known relation (see [14]) + (H + H )2Σ+. u  បω K (ω) = k (ω) 1 Ð exp Ð------. (32) T T  kT  We can calculate the absorption coefficient (26) for freeÐfree transitions (4) using the same approximation We are primarily interested in a comparative analysis of as for (23) and (25). The result is the roles of the photoabsorption processes (3), (4) and (1), (2) as functions of the frequency ω. The factor in 2 2 square brackets in (32) affects the formulas for the frÐfr 3  ប  Rω f ug(Rω) k + (ω) = 4π α ------absorption coefficients for the direct and inverse pro- H ÐH   ∆ me Fug(Rω) cesses in the same way. Therefore, we present our (29) Γ , ⁄ results for the absorption coefficients below, without  Ug(Rω) (3/2 Ug(Rω) kT) × exp Ð------N N +, taking into account stimulated emission.   Γ H H kT (3/2) Combining (31) and (32) leads to the well-known result of Bates [7] for photoabsorption by the ç–H+ ∞ system, derived using a simple model with fixed Cou- Γ(3/2, z) = ∫t1/2eÐtdt = Γ(3/2) Ð γ(3/2, z). (30) lomb centers and a classical Boltzmann distribution function for their coordinates. Our treatment was based z on the initial quantum formulas for the cross sections for photodissociation and freeÐfree transitions. Our As in the case of photodissociation, the total contri- results show that, in a quasi-classical limit for the nuclear bution of the continuum to the absorption coefficient is wave functions, the total contribution of these processes described by a simple formula in the quasi-classical can be described by the simple analytic expression (31), approximation. Comparison of expressions (25) and if, in addition, we use a quasi-continuum approximation (29) for the absorption coefficients for boundÐfree (3) for the vibrationalÐrotational levels. and freeÐfree (4) transitions shows that, in the quasi- classical approximation, they differ only in the factors Our numerical computation was based on the more of γ(3/2, z) and Γ(3/2, z) corresponding to the contribu- exact formula (15) for the partial photodissociation + cross section and an analogous formula for the contri- tions of the discrete and continuous spectra of the H2 bution of the continuum. The computation shows that | | ion (where z = Ug(Rω) /kT). In particular, both contribu- the correction factor (17), allowing for the quantum character of the nuclear motion in the classically for- tions are proportional to the product N N + of the H H bidden region near the turning points, affects the total number densities of hydrogen atoms H(1s) and protons absorption coefficients in the visual, near-infrared, and + H in the continuum. This enables us to derive a more near-ultraviolet only slightly for the gas temperatures compact analytical expression for the total absorption T = 4500Ð10000 K under consideration. This immedi-

ASTRONOMY REPORTS Vol. 44 No. 5 2000 PHOTODISSOCIATIVE ABSORPTION 345

ω –1 ω λ η Table 3. Dependence of kT ( ), cm on (or wavelength ) for temperatures in the range T = 4500Ð10000 K. Values of T Ð39 5 (ω) = k (ω)/N H + calculated using (31) are in units of 10 cm T H H

T, K បω λ µ , eV , m Rω [a0] 4500 5000 5500 5800 6000 6500 8000 10000 η ω Ð39 5 T( ), 10 cm 5.701 0.217 3.0 16.09 9.33 5.98 4.76 4.13 3.02 1.49 0.807 4.925 0.252 3.2 12.40 7.55 5.03 4.08 3.59 2.69 1.41 0.809 4.256 0.291 3.4 9.54 6.08 4.21 3.48 3.10 2.39 1.33 0.805 3.676 0.337 3.6 7.44 4.96 3.56 3.01 2.71 2.14 1.27 0.803 3.173 0.391 3.8 5.93 4.12 3.06 2.63 2.39 1.94 1.21 0.806 2.736 0.453 4.0 4.80 3.47 2.67 2.33 2.14 1.78 1.17 0.811 2.356 0.526 4.2 3.97 2.98 2.36 2.09 1.94 1.64 1.13 0.820 2.027 0.612 4.4 3.39 2.63 2.14 1.92 1.80 1.55 1.12 0.840 1.744 0.711 4.6 2.95 2.36 1.96 1.79 1.69 1.48 1.11 0.863 1.497 0.828 4.8 2.64 2.17 1.85 1.70 1.62 1.45 1.12 0.898 1.282 0.967 5.0 2.41 2.03 1.78 1.64 1.57 1.42 1.14 0.940 0.580 2.137 6.0 2.04 1.88 1.75 1.69 1.66 1.58 1.41 1.29 0.253 4.900 7.0 2.38 2.29 2.22 2.18 2.16 2.11 2.01 1.92

∆ ately suggests that the quasi-classical approximation Uug(R0), the classically permitted region of internu- can be applied to calculations of absorption coefficients clear distances does not affect the photodissociation in the photospheres of the Sun and hotter stars and con- cross section, and only the contribution from the con- firms the validity of the semiclassical approach used by tinuum is present. We do not consider this case of large Bates. transition frequencies here. Further development of the theory of absorption by a medium in equilibrium containing atomic and molec- 3. RESULTS AND DISCUSSION ular components must go beyond the BornÐOppenhe- imer approximation, which assumes that the electron We calculated the total molecular (quasi-molecular) and nuclear motions are completely separate. This ion H+ absorption coefficient for several values of ω approximation is most justified for low vibrationalÐ 2 v and T using formula (31) and the data from Table 2. The rotational states. As the quantum numbers and K ω increase, the interdependence between the electron and results for kT( ) are presented in Table 3. nuclear motions ceases to be negligible and can lead to The results show that, at the temperatures T = 4500Ð considerable additional effects. These effects have been 6500 K typical for the solar photosphere, after a very studied in the physics of atomic collisions, where meth- ≤ បω + ods for taking into account electronÐnuclear correla- small decrease at 0.25 < 1 eV, the H2 photoab- tions were developed. Applying these methods to pho- sorption coefficient (31) increases with the frequency toabsorption by diatomic systems could increase the ω. This behavior is primarily determined by the Boltz- | | ω absorption coefficient at long wavelengths. Another mann factor exp[ Ug(Rω) /kT]. The increase in is fol- improvement of the theory is to use methods making it lowed by a decrease in the internuclear distance Rω cor- possible to go beyond the limitations of models with responding to the phototransition, and, accordingly, by linearly intersecting terms (see, for example, [19]). an increase in the absolute value of the potential energy |U (Rω)|. (Recall that U (Rω) < 0 in the frequency range Note also that formulas (25), (29), and (31) refer to g g under consideration; see Table 2.) Absorption will occur the frequency range of interest to us, where the quan- v tum energy បω is considerably less than the splitting from the lower levels K, whose populations increase as ∆ the vibrational quantum number v decreases. The absorp- term Uug at the point R0 where Ug(R0) = 0 (R0 = 1.12 a0 ω tion coefficient kT( ) monotonically decreases as ∆ + បω | | and Uug(R0) = 22.2 eV for the H2 ion). When > exp[ Ug(Rω) /kT] as the temperature rises.

ASTRONOMY REPORTS Vol. 44 No. 5 2000 346 LEBEDEV et al. β ω Table 4. T( ) calculated using (31), (33), and (34) for T = determined by (31), (33), and (34). It follows from

4500, 5800, 8000, and 10000 K, assuming the number den- these formulas that, when N = N + , β (ω) is a dimen- e H T sities of N and N + are equal e H sionless parameter that depends only on the frequency ω β ω and the temperature T. Values for T( ) are presented T, K in Table 4. We used the photoejection cross sections

បω, eV λ, µm 4500 5800 8000 10000 σ Ð (ω) from [2] and the results of [3] for 〈v σ 〉 for H e e–H β T, % the absorption coefficient (34) for freeÐfree transitions (2). 5.70 0.22 63.8 41.1 30.3 27.5 β ω Thus, the values of T( ) in Table 4 characterize the 4.93 0.25 40.3 28.9 23.5 22.5 additional photoabsorption by molecular and quasi- 4.26 0.29 25.7 20.4 18.4 18.5 + molecular H ions over the total photoabsorption by 3.68 0.34 16.7 14.7 14.5 15.2 2 the negative ions H– and collisions of electrons with 3.17 0.39 11.1 10.7 11.5 12.6 + 2.74 0.45 7.7 8.0 9.4 10.6 neutral hydrogen atoms H. Note that the H2 absorption spectrum extends to frequencies considerably greater 2.36 0.53 5.4 6.2 7.7 9.1 than those in Tables 2Ð4. However, a comparison of the 2.03 0.61 4.1 5.0 6.9 8.0 H+ and H– absorption coefficients is meaningless at ener- 1.74 0.71 3.2 4.2 5.4 7.2 2 gies បω > 4.93 eV, since absorption from the triplet level 1.50 0.83 2.7 3.7 5.5 6.8 3s3p 3P of the Mg atom becomes dominant (see [1]). 1.28 0.97 2.6 3.6 5.4 6.7 The frequency range បω ~ 0.5–3 eV, which approx- 0.86 1.43 5.6 7.2 8.3 9.2 imately corresponds to the half-width of the Planck 0.58 2.14 10.0 8.0 6.6 6.0 distribution for black-body emission at temperature 0.54 2.28 9.0 7.2 6.0 5.5 kT ~ 0.5 eV, is of particular interest for studies of solar β ω 0.38 3.22 4.7 4.0 3.5 3.3 emission. In this frequency range, T( ), which speci- fies the relative contributions of photoabsorption by the 0.25 4.90 2.2 2.1 1.7 1.6 H–H+ system and photoabsorption via (1) and (2), is, on average, equal to ~6−7% and is only weakly depen- dent on temperature. There is a considerable increase in Let us compare the resulting values for the H+ β (ω) (by ~30Ð40%) at បω ~ 3–5 eV due to both the 2 T ω absorption coefficient k (ω) and the total coefficient for increase in kT( ) [see (31) and Table 3] and the T decrease in the absorption coefficients for H– ions and absorption by the negative ion H– via the process (1) free–free electron transitions in the field of the hydro- bÐfr gen atom, which are determined by (33) and (34). How- k Ð (ω) = σ Ð(ω)N Ð = σ Ð(ω) H H H H ever, the range បω ~ 3−5 eV is of little interest for cal- culations of the radiation integrated over the spectrum, 2 3/2 (33)  πប E Ð  since its contribution is appreciably suppressed by the × 1 2   H  [ Ð1] -------- exp------ NH Ne cm exponential dependence of the Planck distribution. 4 mekT kT  For stars hotter than the Sun with temperatures T = បω and via freeÐfree transitions (2) 8000Ð10000 K, the range of photon energies 3 < < 5 eV either partially (at T = 8000 K) or completely (T = 10000 K) overlaps the half-width of the Planck distri- frÐfr ω 〈 σ 〉 [ Ð1] keÐH( ) = v e eÐH T NH Ne cm . (34) bution. The results presented abo v e (T able 4) sho w that + 2 photoabsorption by H2 ions substantially increases the Here, σ Ð (ω) [cm ] is the cross section for the photo- H total absorption coefficient. In this case, the character- – β ω ejection (1) of an electron from an H ion and istic values of T( ) are approximately 7Ð12% at visual 〈 σ 〉 4 ve e−H [cm s] is the effective cross section for pho- wavelengths and can be as large as 20Ð30% at short toabsorption (2), normalized to a unit flux of electrons wavelengths. with velocity ve colliding with the H(1s) atom (for At temperatures of about 3000 K, the contribution details, see [14]). + of photodissociation of H2 ions and freeÐfree pho- Let us consider the ratio totransitions in the H–H+ system integrated over the spectrum is approximately 3% of the total contribution ph.d ω frÐfr ω k + ( ) + k + ( ) k (ω) of photoabsorption by H– and freeÐfree electron transi- β ω H2 H ÐH ------T------T( ) = -----bÐfr------frÐfr------= bÐfr frÐfr , (35) tions. However, the relative contributions of photoab- k _ (ω) + k (ω) k _ (ω) + k (ω) H eÐH H eÐH sorption by the H–H+ system and of photoabsorption

ASTRONOMY REPORTS Vol. 44 No. 5 2000 PHOTODISSOCIATIVE ABSORPTION 347 via processes (1) and (2) appreciably increases at short 7. D. R. Bates, Mon. Not. R. Astron. Soc. 112, 40 (1952). β ω wavelengths, and the corresponding values of T( ) 8. L. D. Landau and E. M. Lifshitz, Statistical Physics reach ~50–150% at energies 4.5 < បω < 5.5 eV. Note (Nauka, Moscow, 1976; Pergamon, Oxford, 1980). also that the freeÐfree electron phototransitions in the 9. R. A. Buckingham, S. Reid, and R. Spence, Mon. Not. R. fields of neutral hydrogen molecules H2, whose number Astron. Soc. 112, 382 (1952). density sharply increases with decreasing temperature, 10. R. L. Kurucz, Atlas: A Computer Program for Calculat- must be also taken into account when considering ing Model Stellar Atmospheres, Spec. Report No. 309 of absorption in stellar atmospheres with temperature Smith. Astrophys. Obs. T ~ 3000 K. 11. Yu. D. Oksyuk, Opt. Spektrosk. 23, 214 (1967). 12. G. H. Dunn, Phys. Rev. A: Gen. Phys. A 172, 172 (1968). 13. K. P. Huber and G. Herzberg, Molecular Spectra and ACKNOWLEDGMENTS Molecular Structure: IV. Constants of Diatomic Mole- We are grateful to A.A. Boyarchuk for stimulating dis- cules (Van Nostrand Reinhold, New York, 1979; Mir, cussions. This work was supported by the Russian Foun- Moscow, 1984), Chap. 1. dation for Basic Research (project no. 99-02-16602) and 14. I. I. Sobel’man, Introduction to the Theory of Atomic the “Integratsiya” Federal Program (grant 2.1-35). Spectra (Nauka, Moscow, 1977). 15. V. S. Lebedev and V. S. Marchenko, Tr. Fiz. Inst. im. P. N. Lebedeva 145, 79 (1984). REFERENCES 16. L. D. Landau and E. M. Lifshitz, Quantum Mechanics: 1. A. Unsöld and B. Baschek, The New Cosmos (Springer, Nonrelativistic Theory (Nauka, Moscow, 1974; Perga- Berlin, 1991). mon, Oxford, 1977). 2. M. Venuti and P. Decleva, J. Phys. D 30, 4839 (1997). 17. V. S. Lebedev, J. Phys. B: At. Mol. Opt. Phys. B 24, 1993 3. K. L. Bell and K. A. Berrington, J. Phys. B 20, 801 (1987). (1991). 4. R. Wildt, Observatory 64, 195 (1942); R. Wildt, Astron. 18. Handbook of Special Functions, Ed. by M. Abramovits J. 54, 139 (1949). and I. Stigan (Nauka, Moscow, 1979). 5. D. R. Bates, J. Chem. Phys. 19, 1122 (1951); D. R. Bates, 19. E. E. Nikitin and S. Ya. Umanskiœ, Nonadiabatic Transi- K. Ledsham, and A. L. Stewart, Philos. Trans. R. Soc. tions in the Presence of Slow Atomic Collisions (Atom- A 246, 215 (1953). izdat, Moscow, 1979). 6. D. M. Bishop and L. M. Cheung, J. Phys. B 11, 3133 (1978). Translated by Yu. Dumin

ASTRONOMY REPORTS Vol. 44 No. 5 2000