Appendix C. Model of Accretion Disk Near Black Hole
Accretion Disk: Generally speaking, different types of accretion disks are discussed in the literature: thick disks (“doughnuts”), thin disks (Shakura-Sunyaev model), slim disks, and advection-dominated accretion flows. The standard thin accretion disk model is applied for an accretion rate that is substantially smaller than the Eddington limit. This model was one of the first studies of accretion disks in the vicinity of black holes, formulated in the early 1970s by Shakura and Sunyaev. The authors articulated the basic assumptions and equations to describe classic thin disks in the Newtonian approximation of gravity. Relativistic disks were first described by Bardeen, Press, and Teukolsky, and by Novikov and Thorne. Thorne introduced the “slim disk” model, whose equations include a number of terms neglected in the thin disk model equations. For more detail, see, for example, Refs. [
46,
47]; these reviews cover the main aspects of the black hole accretion disk theory.
In the model presented below, the following presumptions are made about the accretion disk: (a) It is thin, i.e., its characteristic vertical thickness is much smaller than its characteristic horizontal size; (b) It is located in the equatorial plane, implying that the averaged -component of 4-velocity becomes suppressed; (c) It rotates with the rotation axis perpendicular to the disk plane; (d) It is heated non-homogeneously, with higher temperatures at the outer layers (due to the processes of the tidal disruption of captured matter and strong dissipation in the shear flows).
Let us assume that, due to some mechanism, a locally “twisted” region of heated matter spontaneously formed in the rotating disk. Then, in the field of the centrifugal force, within the outer layers, “hot bubbles” form (i.e., localized plasma clusters whose temperature exceeds the “average” for the disk and whose density is lower), which move
towards the axis of the disk rotation. However, when these hot bubbles are also vorticial in nature (for example, due to the overall rotation of the disk), then each vortex bubble (via induction of the velocity field) “forces” other vortex bubbles to rotate around itself. Hence, the bubbles move “sideways”, rather than directly towards the rotation axis of the disk. As a result, since all vortices are affected by the cumulative velocity field, they gradually self-organize into zones of stability—a symmetric thermo-vorticial macro-structure which rotates as a whole around the mutual center of symmetry. The dynamics and longevity of this structure are linked to the thermal and vorticial properties of the system and its elements. Visually, the protruding components of this structure look like bright “hot petals”. For large-scale quasi-2D flows in “thin” fluid layers, with very high Reynolds numbers, the so-called “smoothing” effect is observed when the level of vorticity acquires an approximately constant value throughout the region of the vortex (see classical Ref. [
26], and subsequent developments in Refs. [
9,
10,
18,
27], and references therein).
The following assumptions are also made in the model: the mass of the accretion disk (
) is negligible compared to the black hole “mass”
, the radiative cooling does not strongly affect the dynamics of the fluid motion in the disk, the electrons and ions are very weakly coupled by Coulomb interaction, and therefore, ion and electron plasma components may have different temperatures, even with
much greater than
(see Ref. [
28]). Then, it is the electron plasma component which contributes the most to the equation of state of the accretion disk matter. Due to the large difference in the masses of electrons and protons, electrons are highly mobile and provide the quasi-neutrality of the plasma. Also, due to the high conductivity of the plasma, its own magnetic field can be considered “frozen-in”. A range of possibilities may exist. In hot flows like those around Sgr A* or M87, electron temperature is thought to be typically lower than the ion temperature due to radiative losses to synchrotron, inverse Compton, and bremsstrahlung processes.
Model Assumptions Regarding the Thickness Of the Disk: We consider the motion of the medium relatively far from the event horizon, i.e., for hydrodynamic structures, it may be assumed that , where is the Schwarzschild radius. In this paper, the approximation may be considered “far away”. When we are not dealing with phenomena occurring near the event horizon (which is ), then, for , we can restrict ourselves to the roughest approximation for the spacetime metric near a rotating (or not ) black hole.
It is assumed that the disk is geometrically thin——where h is the characteristic local thickness of the disk.
When (more specifically, ), then the vertical component of gravity acceleration can be estimated as . This is balanced by the pressure gradient. Here, G is the gravitational constant, c is the speed of light, , and is the radius dependent on the so-called Kepler parameter: the angular velocity of a test particle with a circular orbit at a distance r from in the Newtonian approximation of gravity.
Gas pressure along the direction perpendicular to the disk plane is determined by the hydrostatic equilibrium, . Generally speaking, electromagnetic radiation and the induced magnetic field may contribute to the pressure. In the simplest case, the dominant contributor is the gas thermal pressure; then the vertical temperature distribution is isothermal, which is an acceptable approximation when the disk is optically thick and externally heated. The equation of state of gas/plasma is then , where s is the “isothermal sound speed” (which is not a function of the transversal coordinate z).
If we further assume that , the equation of hydrostatic equilibrium becomes , the solution of which is . Here, the transversal space scale h is introduced: . Parameter represents the density at the disk mid-plane (at ), and parameter h is the characteristic local thickness of the accretion disk.
In order for the accretion disk to be considered thin, it is necessary that
. On the other hand, for great distances from the black hole (
), specific relativistic effects may be neglected or parametrized. Thus, we obtain the following natural bounds for the validity of our consideration:
. In other words, the presented model may be valid for the consideration of spots in accretion disks (resembling those in
Figure 1A, for example) if the spots are located within these limits.
Note that we work with the field, not with individual particles (their trajectories).
“Thin” Disk Temperature Distribution : The problem of how physical parameters, such as temperature, are distributed within a geometrically “thin” accretion disk may be examined, of course, within the framework of the hydrodynamical approximation, introducing, among others, the concepts of turbulent flow, Reynolds tensor, turbulent viscosity, and the closure of equations. Instead, since we are aiming to obtain only a model-specific estimate, we will first (1) neglect the dependence of physical parameters on the perpendicular to the disk-plane z-coordinate (i.e., integrate/average vertically) and (2) consider the disk without radial thermal advection (i.e., the transfer of heat with matter along the radius) and without the loss of mass from the lateral surfaces of the disk. In such a disk, the angular velocity of rotation of the disk matter at each location along the radius r is approximately equal to the angular velocity of rotation of a free test particle. In other words, . (Recall that key parameters that determine the structure of a geometrically thin disk are the mass of the gravitating center , the internal radius of the accretion disk , and the accretion rate .)
The next assumption is that the gas of the accretion disk spirals inward and this spiraling is very gradual, i.e., the orbits of the gas particles are almost circular. The orbital speed is Keplerian and is estimated as . Quantity is the so-called Keplerian parameter, which is radially dependent, determining the period () of revolution of the test particle located at distance r from the center of attraction .
The categorization of a (vertically isothermal) disk with characteristic thickness h as “geometrically thin” means that , where , is the isothermal sound speed for the gas, T is the absolute temperature, and the gas pressure is . Then, it follows from the static equilibrium for the motion in z-direction that . This yields , where is the density of the disk mid-plane.
The specific angular momentum of the test particle is . This means that, to flow inward (so that the radial velocity ), the gas must lose its angular momentum by redistributing the angular momentum within the disk. The “inner” gas transfers its angular momentum to the “outer” gas; the gas matter flows inward. The loss of the angular momentum by the entire system (via the outward wind, for example) results in the inward flow of the remaining disk matter (with increased viscosity, the process becomes diffusive).
For the unit mass located at the radial distance r, the acting potential produces the (inward-directed) force of attraction . When the mass is transported inward over the distance , its potential energy is changed by . The potential energy of the particle when moved from infinity to the location r decreases from zero to . One half of this energy is converted into kinetic energy; the other half is radiated away (this is the consequence of the so-called “virial theorem”). In fact, for any gravitationally bound system the time-averaged potential gravitational energy and the time-averaged kinetic energy of the motion of the particles of the system satisfy the following condition: . Thus, variations in the quantities are linked via the simple relationship: . Since the total energy of the disk is , then . In other words, any addition of energy to the disk reduces the kinetic energy of its particles (the disk components) and, conversely, energy radiation leads to an increase in , i.e., the temperature of the disk. Since one half of the variation in the gravitational energy goes to the kinetic energy of the gas, then the other half is radiated. The luminosity is thus . Divided by the radiating area, (the first factor 2 appears because the disk has two sides), this expression produces the luminosity per unit area.
In the approximation where the disk radiates as a black body, its radiation power may be characterized by the effective temperature T (from , where is the Stefan–Boltzmann constant). Equating the quantity L to the rate of energy loss via black-body radiation, we obtain . When is presumed to be independent of the coordinate r, this expression leads to the radial temperature distribution: .
This estimate was obtained for disks without radial advection (the transfer of heat with matter along the radius), without the loss of mass from the surfaces of the disk, without consideration of the boundary conditions at the inner edge of the disk, etc. For an extended source of mass (such that, within the disk, where , the rate and parameter )—i.e., when the inward mass transfer is balanced by the mass production at the disk periphery () and the mass is “devoured” by the black hole at the event-horizon ()—the radial temperature dependence is . This means that, depending on the value of the parameter s, the temperature curve may take various shapes. For , T is radius-independent (the disk is heated uniformly); no heat flow occurs along the r-direction. For , the disk is hotter at its periphery, where the tidal destruction of captured objects takes place.
Model of Spacetime Metrics: Obviously, equations of fluid motion in the vicinity of a black hole must be written using the concept of relativistic dynamics. In our model (see details in Refs. [
29,
30]), the spacetime metric is fully characterized by the black hole’s mass parameter and “spin” (for an extensive discussion, see Refs. [
31,
32,
33,
34,
35,
36], and the bibliographies therein). Specifically, we use the Kerr metric—an exact, singular, stationary, and axially symmetric solution of the Hilbert–Einstein equations of the gravitational “field” in a vacuum.
The notations, here and presented below, are as follows: Latin indices and suffixes take the values ; parameter , c is always the speed of light; the Greek letters take the values and correspond to the spatial coordinates. The Galilean metric (special relativity) is characterized by a metric tensor . For the three-dimensional vector below, in Cartesian coordinates, there is no need to distinguish contra- and covariant components.
Using the Boyer–Lindquist four-coordinates
—and it is well-known that, besides the Boyer–Lindquist coordinate representation, other representations of spacetime locations exist—the square of the interval is written as
, where notations are standard and the components of
depend only on the dimensionless combination
and
. Here, as accepted,
is the Schwarzschild radius,
c is the speed of light,
G is the gravitational constant, and
is the “mass” of the black hole. The off-diagonal term
in the metric tensor is proportional to the rate of the black hole’s own rotation and to
. We use the metric signature
(see, for example, Ref. [
30], and references therein). To satisfy the principle of causality for moving material objects,
. The four timespace coordinates
provide the location of a world event from the viewpoint of a remote observer. The meaning of space coordinates
is clear once transitioned to the limit
. When the square of the interval becomes
, i.e., at infinity, parameters
may be interpreted as the standard spherical coordinates in flat spacetime. For parameter
r, strictly speaking, note that it is not the “distance”, in the usual sense, from the center of the black hole. This is because, for any material object, in the spacetime defined by equation
, no central point
exists in the sense of a world event on a valid world-line.
Relativistic Flows Of Perfect Fluids: We take into account the effects of special and general relativity to obtain the relativistic Euler equations. To avoid misunderstandings that may arise due to the underdefinition of some concepts (for example, the definition of the signature of the metric tensor), we include many details in this section, even though an advanced reader will certainly be aware of them.
As is known (see, for example, Ref. [
31]), the contravariant energy-momentum tensor of a perfect relativistic fluid is written as
. Here,
e is the internal energy of the fluid;
p is the pressure. The quantity
is the heat function (enthalpy);
is the contravariant metric tensor. The quantity
is the contravariant 4-velocity of the fluid flow;
is its covariant 4-velocity. The 4-velocity vectors of the flow are normalized by the condition
. The covariant metric tensor (included in the definition of the interval
via
) is the tensor
reciprocal to the tensor
; that is,
. The usual rule of summation is always used. The metric signature is chosen as
. The procedure of the raising and lowering of these indices follows the standard rule. Therefore, the 4-vector components’ links are as follows:
and
.
The relativistic internal energy e includes the rest energy of particles , where m is the rest mass of one particle and n is the proper number density of particles (i.e., is the volume per particle). The heat function w, normalized per , is written as . Here, captures the non-relativistic part of the heat function. Normalized pressure has the same dimension as normalized w: .
The explanations of why tensor
takes such a form can be found in Ref. [
12], §133. The mixed tensor
is thus
. A more complex model of the stress–energy tensor of a viscous relativistic fluid with an energy flux may be found in Ref. [
33], §22.3, or Ref. [
12], §136. A method for building analytical models of relativistic accretion disks may also be found in Ref. [
46], which also contains an extensive bibliography on the topic.
To better explain the idea of how to construct the equations of fluid motion, as the first step, we consider the flow in the flat spacetime. Then, if the Cartesian coordinates are used,
. The equations of fluid motion and the condition for the conservation of the proper number density of particles
n are contained in the following equations:
Reminder: this system becomes closed through the inclusion of the equation of state (for the ideal gas, Fermi gas, plasma, etc.). This is the most subtle part of the problem: one cannot just use some equation of state to “see what happens”, one must consider the real astrophysical situation in the proper spacetime for the problem. The magnetic field can also be included in the consideration. Then, tensor would contain an additional magnetic term. If the plasma is highly conductive, then the magnetic field may be considered “frozen-in” within the fluid. Then, the additional Lorentz force appears in the dynamic equations and an additional equation for the evolution of the “frozen-in” magnetic induction appears in the system of equations. Magneto-hydrodynamics is strictly applicable in configurations when the mean free path and mean free time of electrons and protons are much smaller than the characteristic scales and time intervals of the macroscopic motions in question. However, there exist situations when, even for systems with long free paths of current-carriers, the equations obtained from the kinetic theory are formally identical to the MHD equations. Such a situation is observed, for example, in a non-equilibrium plasma when the electron temperature considerably exceeds the ion temperature.
It is physically clear that when a conducting fluid moves in a magnetic field, electric fields are induced in it. Thus, electric currents begin to flow. The magnetic field exerts forces on these currents which, in principle, may considerably modify the flow characteristics. Conversely, the currents themself perturb and even strongly modify the magnetic field. Consequently, a complex interaction between the magnetic and fluid dynamics phenomena takes place. Thus, the fluid flow must be studied by combining the fluid dynamics equations with those of electromagnetic field equations—this is the MHD formulation. Such an approach covers a wide range of physical objects, from liquid metals in a magnetic field to cosmic plasmas. In most configurations, MHD processes are extremely complex.
The description radically simplifies if (a) all dissipative processes are neglected, i.e., no account is taken of thermal conduction and viscosity and the electrical conductivity is considered unbounded such that, in the case of a perfectly conducting fluid, the electrical field is completely screened and the magnetic field is “frozen-in” within the fluid; (b) the fluid is incompressible; (c) in the equations of conservation, the following terms are added: to the density of the hydrodynamical energy , and the Maxwell stress tensor to the hydrodynamical momentum flux density tensor . This leads to the appearance of the Lorentz force in the equation of fluid motion: .
The MHD formulation implies that the displacement current is neglected in the Maxwell equations, i.e., ; together with the “frozen-in” condition, this leads to the following condition: . Here, c is the light speed, l and are, respectively, the characteristic space scale of the considered flow structure and of the changes in its its evolution. From the Maxwell equation for the evolution of the magnetic field , which, in highly conductive plasma, becomes , it follows that . From these two inequalities, we find that , i.e., the flow must be non-relativistic. Because , i.e., , and taking into account , we find that it must be that . Thus, if the inequality (where is the speed of sound) is satisfied, the hot plasma may be considered non-relativistic and the effect of the magnetic field can be taken into account as a small perturbation and can be neglected in the leading approximation.
With respect to the modeling of black hole accretion disks in general, one caveat is critically important: the present understanding of the physical conditions within the disks (for example, the equation of state of the matter) is highly uncertain; there are no observational measurements that confirm any of the existing models of the equation of state. Unfortunately, as all studies note, the sensitivity of numerical models to the uncertainties in the parameters of the equation of state is also high. Hence, until some reliable observational data appear regarding the physical conditions in the vicinity of black holes, the overzealous obsession with numerical details remains meaningless and semi-qualitative estimates will suffice.
Taking the explicit expression for
and differentiating it (using notation
), we obtain the following from Equation (
A54):
We must remember that derivatives are to be regarded as the covariant components of the 4-vector-operator. In fact, the differential of the scalar f, namely , is also scalar. The scalar product of two 4-vectors, , is also scalar, making our assertion obvious. The operator of differentiation with respect to the coordinate, , determines the covariant components of the 4-vector operator . The index is raised according to the usual procedure: . In the Galilean metric, contravariant , where the enumerating index for space components, as mentioned earlier, is .
Next, project Equation (
A55) onto the direction of the 4-velocity, i.e., we multiply it by vector
. This product is obviously also zero:
In the first term, the coefficient is
. The second term is zero because
, i.e.,
, and, in the Galilean metric, in the absence of the spacetime curvature, all components of the Christoffel symbol are zero. Thus,
(Recall that
, where
is the Lorentz factor and
is the non-relativistic part of the internal energy.) When we take the equation of continuity, i.e., the second equation in the system Equation (
A54), into account, then Equation (
A57) is transformed into
In the system where the total number of particles is constant and there are no external heat sources, the fundamental thermodynamical identity is
. Here,
T is the absolute temperature,
s is the entropy per unit of proper volume,
is the volume per particle,
w is the enthalpy per unit of proper volume, and
is the entropy per one particle. Noting te equation of continuity, we obtain from Equation (
A57) the equation of conservation for the entropy flux:
This means there is no heat exchange between adjacent fluid particles (which are composed of enormous numbers of individual subatomic particles).
The next step is to project the first equation in Equation (
A54) onto the direction perpendicular to the 4-velocity. For this, we use the projection operator, i.e., we build
, which is also zero. Thus,
Using the definition of
and simplifying some terms, we find four equations:
Here, the three space components of these four equations are the relativistic generalization of Euler equations; the time component for
is the consequence of the other three. Equation (
A61) was obtained from Equation (
A60) and the definition for
, as follows:
Here, the second term cancels the fifth term because
, and the fourth term is zero because
, for the same reason; from this, we can obtain Equation (
A61).
The equations of relativistic fluid dynamics in the general theory of relativity are obtained from Equations (
A59) and (
A61) by simply replacing the ordinary derivatives with the covariant ones (see Ref. [
12], §134) and keeping in mind that the 4-velocity expression is modified as well:
Recall that the covariant derivatives for vectors and are respectively determined as and . The quantities of are the Christoffel symbols. Also, we recall that these quantities are expressed via the metric tensor as In Galilean coordinates, Christoffel symbols ; therefore, the covariant differentiation reduces to the ordinary differentiation.
Because the covariant derivative of a scalar function produces the same result as an ordinary derivative, Equations (
A63) are rewritten as
or, for a contravariant component of 4-velocity of flow
, the first equation becomes
When the coordinate grid and spacetime metric are specified, then the group of the last terms in the parenthesis in Equation (
A65) manifests itself as a specific “force” acting on a fluid relativistic particle in the “Euler description.” If, in analogy with the dynamics of relativistic particles with
, we wrote, in the left part of Equation (
A65), the quantity
, this would have been incorrect. Properly, the left part should look as written in Equation (
A65). The covariant quantity
is not the specific covariant force acting on the fluid particle; this force is the quantity
.
It is helpful to see what will happen in the case of low velocities and in the absence of fields, i.e., when , . For the -spatial contravariant component of 4-velocity, we obtain (with the coefficient ) the “material” derivative: . To find the covariant component of the “material” acceleration , we use the standard rule for the lowering of indices.
Thus, we become capable of calculating the spatial “force” components, using
, based on the following equation:
which has a transparent physical meaning: the left side of the equation determines the kinematic characteristics of the process—the evolution in time and space of the 4-velocity—which characterizes the state of the fluid particle; the right side of the equation specifies the causes of this change, namely, the pressure gradient and the presence of an external force field.
The subsequent calculations are conceptually rather transparent: for a given metric tensor
found from interval
, defined via
, we calculate the Christoffel symbols
; everything is substituted into Equations (
A66), which are solved for the selected model of the equation of state of (disk) matter for the quantities
. Equations (
A66), which can be applied in many physically interesting situations, are quite complicated. The equations of fluid dynamics in the first approximation beyond the Newtonian (i.e., in the post-Newtonian, via expansion with respect to parameter
) were obtained in Ref. [
37], and are discussed in Ref. [
33]. Additional information can be also found in Refs. [
38,
39,
40,
41,
44].
Explicit Form of Metric Coefficients for Kerr Geometry with Rotation: For the Kerr metric (see Ref. [
30] and references therein), the components of the metric tensor may be found from the following expression:
If we need to transition to an uniformly rotating frame of reference, then we make the following transformation
, where parameter
is the angular velocity of the rotation co-linear with the
z-axis. The same notations remain:
. Then,
This dimensionless expression is expressed in units of length
and time
. Therefore,
,
,
. Thus, Equation (
A68) can be written as
using the following:
As is known, for the Kerr model, the parameter is always
. From Equation (
A69), it follows that dimensionless spacial parameter
x is bounded as follows:
if
. Obviously, setting
and
in Equation (
A69) provides the Schwarzschild metric. Also, the limit value
corresponds to the external event horizon.
Finally, we provide the expressions for the components of metric tensor
in the leading approximation for
and
:
From these expressions, it follows that if the fluid flow is away from the event horizon (i.e., in dimensionless units) and parameter satisfies conditions and (approximately) , then the principal contributions to the components of the metric tensor come from the terms which depend on ; the stand-alone gravitational part in the metric is not the focal one.
The expressions in Equation (
A70) make it possible to determine the size of the strip of a thin accretion disk, within which the approximation of the classical hydrodynamics can be determined, and where the main contribution to the flow dynamics is made not by the fine structure of the metric in the vicinity of a rotating black hole, but by the effect of the rotation of the accretion disk, characterized by parameter
. Parameter
appears in these expressions because the accretion disk mainly forms through the captured matter, which has a non-zero angular momentum; therefore, the (averaged) angular velocity of the disk—which is a material formation of finite mass and limited dimensions—is non-zero. Hence, it is logical to transition to a coordinate system rotating with such angular velocity. However, the very concept of a rotating coordinate system contains two implied caveats (see Ref. [
31], § 84): (1) some material bodies must exist (or the coordinate system has no “anchors”) and (2) they exist within the spacetime domain, which is necessarily bound (or the material bodies outside
would rotate with velocities exceeding the speed of light, which is impossible; the limit
is where component
of the metric tensor turns negative). This means that a rotating coordinate system cannot extend to infinity. In other words, it is important to remember that all considerations are always made for a finite spacetime domain surrounding the axis of rotation—beyond this domain lies the “forbidden” zone. (Near the axis of rotation is the other “forbidden” zone—the ergosphere.)
As the rotation speeds up (
), the “forbidden” zones (where
) transform from two detached zones (
Figure A1A) into one merged zone (
Figure A1B).
Figure A1.
Domains of spacial coordinates where the -component of the metric tensor becomes negative (darkened zones). Calculations are made for the conditional values of parameters: left panel (A) ; right panel (B) .
Figure A1.
Domains of spacial coordinates where the -component of the metric tensor becomes negative (darkened zones). Calculations are made for the conditional values of parameters: left panel (A) ; right panel (B) .
With respect to the relativistic consideration in general, several critically important caveats exist: (1) A “rigid coordinate grid” (such as the Boyer–Lindquist parametrization) cannot span everywhere, from infinity and into the ergosphere. The meanings of the parameters defining the world-point for a body do change when transitioning from one domain of spacetime into the other. This means that a “singularity” may appear. (2) The singularities of both the event horizon and the static limit in the Kerr metric are illusory. They become singularities only within the “unfortunate” choice of coordinates. Indeed, just like in the case of being far away from the black hole (at infinity) where the Boyer–Lindquist coordinates conveniently turn into normal spherical coordinates, in the case of being near the black hole, the spacetime may be smoothly described (without singularities) by differently chosen coordinates, such as Kruskal–Szekeres, Lemaitre, or Eddington–Finkelstein coordinates (see, for example, Ref. [
36]). Indeed, besides the Boyer–Lindquist coordinates, a number of representations of spacetime locations exist (see, for example, Refs. [
42,
43]). To clarify this idea: consider the spherical coordinate system. When angular parameter
or
(think of the North/South Poles on the globe), the contribution to the interval
(via
) from the term
always equals zero, i.e., any value of
describes the same world-point (the Pole). Therefore, if working with spherical coordinates, at the Pole, we can see an illusory singularity (the
-coordinate appears to “fold”). However, this is clearly a product of the “unfortunate” choice of coordinates; the “singularity” disappears in another coordinate system—in the Cartesian, for example.
Linkage Between 4-velocity and 3-velocity: To express the 4-velocities of flow
(in the curved spacetime with rotation) via 3-velocities, we separate the coordinates of time and space (denoted by the Greek letters) in the general form of the square of interval
:
For the special case of the Kerr geometry with rotation, quantities
are easily expressed via
. Next, we rewrite general Equation (
A71) as follows:
where the distance
between two close points is found from
. Here, tensor
(which should not be confused with the Lorentz factor
) includes the 3-vector
whose covariant components are defined as
. Then, Equation (
A72) can be rewritten as follows:
The expression for
is apparent from Equations (
A72) and (
A73) and reveals that, in order to measure the magnitude of flow velocity
v, one must simultaneously measure the distance between neighboring points
(in Equation (
A72)) and the interval of the corresponding time. This is why the expression
is introduced.
From Equation (
A73), it follows that
The link between the components of the general tensor
and the components of the specific case of the Kerr metric with rotation
is given by the following:
Tensor
, introduced in Equation (
A73), becomes, in this cas,
. It should be noted that tensor
determines the metric in 3D space: the distance between two close points can be found from
. The contravariant 3D metric tensor is
(see Ref. [
31], §84). Also, when
, the contravariant component of the same vector
is determined as follows:
. We can also obtain that
.
Synchronized Time: The introduction of parameter
defined by
—the so-called
synchronized time—is not some kind of mathematical manipulation. It reflects the fundamental need within the framework of relativism to simultaneously measure positions at two separate points and, therefore, to measure the particle velocity. This means that the measured time interval is
, divided by
c; the concept of the velocity of fluid particle movement is formulated using this very time interval. Equation (
A71) allows us to conclude that 3D contravariant and covariant velocities take the following form:
Furthermore, from Equations (
A74) and (
A76), the contravariant components of 4-velocity are not introduced arbitrarily, but, following the procedure above, are expressed via the 3-velocity vector
as follows:
Equation (
A77) provides the link between the 4-velocities
and 3-vectors of flow velocities
. Note that, in this case, the operator
can be written as
, where
and
. In fact,
Thus, the operator D is nothing more than a “material” derivative with respect to time, in which, instead of differentiation with respect to the time t of a remote observer, differentiation with respect to synchronized time is implied.
“Specific Force” and Coriolis Effect: Let us return to Equation (
A65), which we write using the following form:
where, after some simple manipulations,
becomes a remarkably symmetric expression
The term
in Equation (
A78) is non-zero in a curved spacetime and manifests itself as a certain specific force acting on a moving fluid particle.
The key question that immediately arises is whether Equation (
A78) provides similar expressions to the centrifugal and Coriolis forces which appear in the classical hydrodynamics.
We expand the expression for
, explicitly selecting terms with
:
The first term in Equation (
A80) can be rewritten as follows:
Indeed, this expression turns into the simple model of Newtonian gravity in the limit case of the weak field, non-rotation of the black hole, and when the very last term is zero. This last term reflects the combined effect of the rotation of “everything” (via ), movement (via ), and gravity (via ).
We rewrite the second term in Equation (
A80) as follows:
Here, in a system with the right orientation of basis vectors, the Levi–Civita symbol
(recall that this is not a tensor) has the components
, and its sign changes when index transposition is not even. In curvilinear 3D coordinates, the unit antisymmetric tensor is defined as
, where
d is the determinant of the matrix of the metric tensor
. Respectively, the tensor
. The expression in parenthesises in Equation (
A82), i.e.,
, is
-component of the 3D contravariant vector
(see Ref. [
48]; or Ref. [
31], footnote on p. 252). In combination with other multipliers, Equation (
A82) produces the cross-product of vectors
and
with coefficient
. Indeed, if
, then
. Thus, the second term in Equation (
A80) can be written as
, in accordance with the definition of the cross-product in 3D curvilinear coordinates. The second term in Equation (
A80) has an analogous form (absent external fields) to the Coriolis specific force that would appear in a frame rotating with angular velocity
.
In the limit case of low velocities and in the absence of fields (i.e., when
), the non-zero component of
, i.e.,
, becomes
, which is the angular velocity of the rotation of the coordinate system. In this case, Equation (
A82) becomes
, i.e., we can obtain the expression for the Coriolis force in spherical coordinates, with the presence of coefficients 2 and
. Recall that, in the full expression for the radial component of the cumulative “force,” the following terms are summarized: the radial component of the Coriolis force (
), the radial component of the centrifugal force (
), and the term (
) from the material derivative for the flow velocity. By summing up, we can obtain the radial component
.
The third term—which provides the relativistic correction to the leading terms in Equation (
A80) —appears due to both the large flow velocities and the curvature of spacetime. This term is present when the derivatives with respect to the 3-coordinates of the spatial part of the metric tensor,
, are non-zero. After the substitution
, some of the terms are canceled, and the remaining ones contribute to the Christoffel symbol over a spatial basis. In Cartesian coordinates, this term disappears.
Each term in Equation (
A80) is drastically simplified in the case of low (
) fluid flow velocities, i.e., when
and when
: the first term in
is then transformed into a specific potential “force” (acting on a fluid particle of flow) which is a strictly radial one in the case of the Schwartzschild metric. In the case of the Kerr metric, this term includes both the “gravitational” action and the “centrifugal” effect dependent on
and
(see
in Equation (
A69)). The second term, as noted above, is analogous to the Coriolis force in the classic hydrodynamics. The third term is negligible for low flow velocities.
The subsequent simplification of Equation (
A80) involves the expansion of all dynamic quantities into series with respect to small parameters
and
, which produces rather cumbersome expressions. The contribution to the equation of fluid motion from the terms containing the higher orders of flow velocity
will be negligible. The key difference compared to the classic hydrodynamic treatment is the replacement of the classic gravity acceleration
with
and the replacement of the classic angular velocity
with parameter
, as written above.
If conditions are satisfied, then, leaving the leading terms in the expansion , we obtain that . Here, terms of a higher order of smallness are omitted. The first term describes the centrifugal effect, the second describes the contribution of the force of attraction in the Newtonian approximation of gravity, the third term is derived from the Schwarzschild model of non-rotating black holes, the last term takes into account the rotation of both the black hole and the disk. For , but , the principal contribution provides the first term , i.e., the centrifugal effect, independently of the exact structure of the spacetime near the rotating, or not rotating, black hole.
Derivation of Equations for Relativistic Flow: In the subsequent treatment, the following expressions are useful. For a 3-vector with covariant components , the contravariant components of vector are . Also, . When the fluid can be considered “incompressible”, the current potential may be introduced: .
By rearranging the terms, we rewrite Equation (
A78) in the following form:
Remember here that (for the spacial components) and .
Presuming that the metric tensor components
and
are time-independent, when low-velocity flows (
, i.e.,
and
) are considered at some distance from the event horizon (i.e.,
, where
is a characteristic magnitude of the Christoffel symbols), then the right side in Equation (
A83) may be dropped. On the other hand, the thermodynamical quantity
gives
. For low velocities,
, and the last term on the left side of Equation (
A83) becomes
. The equation for the evolution of flow velocities then becomes:
The simplest formulation is the one made in the Cartesian coordinates. Then, we may not distinguish between the contra- and covariant components of a vector and simply write the following: and . So, when (for example, when there is only one component of the vector , which is not dependent on the polar coordinate ). Here, is the Laplace operator.
2D Flows: Consider the flow in a thin layer of an ideal fluid in the plane . In this case, the flow velocity has only two non-zero components in the -plane. Then, (in terms of components: ) has only one non-zero component , i.e., is perpendicular to the -plane.
We assume that a state of dynamical equilibrium exists:
We will now write
. Combining Equation (
A85) with Equation (
A84), we find that a perturbed state is described by the following:
In this equation, the small term
is omitted for similar reasons to those articulated in Ref. [
12], §13, or [
14]: We suppose that the space-scale of structures which are of interest to us is small in comparison with the distances over which the force-field from
causes a noticeable change in density, and we can regard the fluid itself as incompressible. This means that, during this process, we can neglect the change in the density caused by the pressure change. The change in density caused by thermal expansion,
, cannot be neglected, because this is the one which causes the phenomenon.
All the necessary attributes in the evolution equation Equation (
A86) are presented: on the left is the kinematic quantity, which is dependent on the flow velocity
, which describes the change in the state of the fluid particle in the Euler description, and on the right are the terms characterizing the causes which drive this change—the gradient pressure and the force characteristic determined by the metric coefficient
.
Next, the following well known expressions are used:
, where
and
which provide
Then, we can move forward. Next, we have
. Obviously,
. In this case, for 2D flows in the
plane,
. Full vorticity can be determined as
. The evolution equation, Equation (
1), for function
(in
) is found after some manipulations, excluding
using the operation
.
The condition
does not mean that the density of the medium is constant. This condition simply means that the density evolution takes place according to the following equation:
. Using
, after some simple manipulation, the evolution equation for temperature perturbations takes the form of the third equation in Equation (
1). For the equation of state we shall assume, in accordance with the assumption of
, that the density essentially depends only on the temperature and not on the pressure. Thus, we set
, where subscript 0 denotes the reference values. The coefficient of thermal expansion
is presumed constant. Since the quantity
is generally small, one may neglect the density variations, and hence replace
with the constant value
in all terms, except in the “buoyancy” term.
Condition
is not merely wishful. This condition is applicable in situations where, on the one hand, the magnitude of the instantaneous current velocity
v may be considered small compared to the speed of sound
s, i.e.,
, and, on the other hand, the time
during which the flow configuration changes meaningfully is large compared to the time which is necessary for the sound signal to travel distances of the order of the size
l of the vorticial structure in the flow, i.e.,
(see Ref. [
12]). Then, we can assume that information about the disturbance of the medium is transmitted by an acoustic signal, as if instantly, i.e., the medium can be considered incompressible. Obviously, the requirement must also be met that the speed of sound
s, although large compared to the value of the local current velocity in the medium, must be less than the speed of light
c.
Condition means that the flow velocity is uniquely determined by specifying the vorticity . Indeed, from the well known identity for any vector quantity , it follows that when , then , and, consequently, . Here, is the Green function of the problem with boundary conditions. The final remark is that since the velocity is determined through the integral of the product of the vorticity q and the Green’s function G, even if the model vorticity distribution function is chosen to be “bumpy”, the integration process—smoothing—produces a resulting function that is “smooth”, i.e., the function is “well-defined” and correctly (at least qualitatively) describes the examined process.