|
. This article has
been written using GNU TeXmacs [13].
Abstract
This document discusses the free boundary equilibrium fitting problem, fixed-boundary equilibrium problem, and magnetic coordinates.
Due to the divergence-free condition, i.e., , magnetic field can be expressed as the curl of a
vector field:
![]() |
(1.1) |
where is called the vector potential of
. (The usefullness of this
representation is that, once
is in this form, we
do not need to worry about the divergence-free constraint.) In
cylindrical coordinates
, the
above expression is written
![]() |
(1.2) |
We consider axisymmetric magnetic field, which means that, when
expressed in the cylindrical coordinate system , the components of
,
namely
,
, and
,
are all independent of
. For
this case, it can be proved that an axisymmetric vector potential
suffices for expressing the magnetic field, i.e., all
the components of the vector potential
can also
be taken independent of
.
Using this, Eq. (1.2) is written
![]() |
(1.3) |
In tokamak literature, direction is called the
toroidal direction, and
planes (i.e.,
planes) are called poloidal planes.
Equation (1.3) indicates that the two poloidal components
of , namely
and
, are determined by a
single component of
, namely
. This motivates us to define
a function
:
![]() |
(1.4) |
Then Eq. (1.3) implies the poloidal components, and
, can be
written as
![]() |
(1.5) |
![]() |
(1.6) |
(Note that it is the property of being axisymmetric and divergence-free
that enables us to express the two components of , namely
and
, in terms of a single function
.) Furthermore, it is ready to prove that
is constant along a magnetic field line, i.e.
. [Proof:
![]() |
![]() |
![]() |
|
![]() |
![]() |
||
![]() |
![]() |
]
(We note that is related to poloidal magnetic
flux, as we will discuss in Sec. 1.7.)
Using Eqs. (1.5) and (1.6), the poloidal
magnetic field is written as
Next, let's examine the toroidal component .
Equation (1.3) indicates that
involves both
and
.
This indicates that using the potential form does not enable useful
simplification for
. Therfore
we will directly use
. Define
(the reason that we define this quantity will
become clear when we discuss the forece balance), then the toroidal
magnetic field is written as
![]() |
(1.9) |
Combining Eqs. (1.8) and (1.9), we obtain
which is a general expression of axisymmetric magnetic field. Expression (1.10) is a famous formula in tokamak physics.
Let us discuss the gauge freedom of in the
axisymmetric case. Magnetic field remains the same under the following
gauge transformation of the vector potential:
![]() |
(1.11) |
where is an arbitrary scalar field. Here we
require that
be axisymmetric because, as
mentioned above, an axisymmetric vector potential suffices for
describing an axisymmetric magnetic field. In cylindrical coordinates,
is given by
![]() |
(1.12) |
Since is axisymmetric, it follows that all the
three components of
are independent of
, i.e.,
,
, and
, which implies that
is independent of
,
, and
, i.e.,
is actually a
spatial constant. Denote this constant by
.
Then
component of the gauge transformation (1.11) is written
(We see that the requirement of being axial symmetry greatly reduces
gauge freedom of .)
Multiplying Eq. (1.13) with
,
we obtain the corresponding gauge transformation for
,
![]() |
(1.14) |
which indicates is similar to that of
electrostatic potential, i.e., adding a constant to it does not make
difference. Note that the definition
does not
imply
because
can adopt
dependence under the gauge transformation (1.13). If
is finite at
, then
is zero there.
This is the case we encounter in the equilibrium reconstruction problem.
Because is constant along a magnetic field line
and
is independent of
, the projection of a magnetic field line onto
plane is a contour of
.
Inversely, a contour of
is also projections of a
magnetic field line onto the plane. [Proof. A contour of
on
plane satisfies
![]() |
(1.15) |
i.e.,
![]() |
(1.16) |
![]() |
(1.17) |
Using Eqs. (1.5) and (1.6), the above equation is written
![]() |
(1.18) |
i.e.,
![]() |
(1.19) |
which is the equation of the projection of a magnetic field line in plane. This proves that contours of
are projections of magnetic field lines in
plane.]
Figure 1.1 shows typical contours of
in a tokamak (CFETR tokamak as an example).
Points where the poloidal field is zero (i.e., =0) are called magnetic null points. There are two
types of null points : O-points and X-points, which can be visually
identified by viewing contours of
.
Mathematically, O-points and X-points are distinguished by the sign of
defined by
![]() |
(1.20) |
where corresponds to O-points and
corresponds to X-points.
Surfaces of revolution generated by rotating
contours around the axis of symmetry (
axis) are
called magnetic surfaces or flux surfaces. No field line intersects
these surfaces. We are only interested in flux surfaces within the
machine wall. Some
contours intersect the wall
before they can form closed curves. These flux surfaces are called
“open”. Otherwsie, they are called closed flux surfaces.
The value of is constant on a magnetic surface.
Meanwhile, the values of
on different magnetic
surfaces are usually different. These two properties enable
to be used as labels of magnetic surfaces. [In
cases that there are multiple magnetic surfaces of the same value of
, the ambiguity can be
resolved by specifying which region the flux surfaces lie in, e.g.,
within or outside the last-closed-flux surface region, near the
high-field side or low-field side, in the Scrape-Off Layer or the
private flux region (the area between the X-point and the material
divertor., i.e, the region between divertor legs that is unconnected to
the plasma).]
In most part of a tokamak plasma, contours of in
plane are closed before they touch the machine
wall. Figure 1.2 shows some examples of closed flux
surfaces.
![]() |
Figure 1.2. Closed magnetic
surfaces (blue) and various toroidal ribbons used to define the
poloidal magnetic flux discussed in Sec. 1.7. The
magnetic flux through the toroidal ribbons |
The innermost magnetic surface reduces to a curve, which is called
magnetic axis (in Fig. 1.2, labels
the magnetic axis).
is zero at the magnetic axis
since
reach maximum/minimum there. As a result,
the poloidal magnetic field is zero there (refer to Eq. (1.8)).
For closed flux surfaces, since
,
the condition
means
points in the anticlockwise direction (viewed along
direction), and
means
points in the clockwise direction.
Note that is defined by
, which is just a component of the vector potential
, thereby having no obvious
physical meaning. Next, we show that
has a
simple relation with the poloidal magnetic flux that can be be measured
in experiments.
![]() |
In Fig. 1.3, there are two magnetic surfaces labeled,
respectively, by and
. The poloidal magnetic flux through any toroidal
ribbons between the two magnetic surfaces is equal to each other (due to
the magnetic Gauss theorem). Denote this poloidal magnetic flux by
. Next, we calculate this flux. To
make the calculation easy, we select a plane perpendicular to the
axis, as is shown by the dash line in Fig. (1.3).
In this case, only
contribute to the poloidal
magnetic flux: (the positive direction of the plane is chosen to be
)
Equation (1.21) indicates that the difference of between two magnetic surfaces is equal to the poloidal
flux divided by
.
In experiments, we measure the poloidal magnetic flux through toroidal
loops around the central symmetric axis (discussed in Sec. 1.8).
Consider one of the loops that located at point , then, using (1.21), the flux through
the loop can be written as
![]() |
(1.22) |
The central symmetric axis is a field line. If
is finite at
,
then
is zero there. This is the case we
encounter in the equilibrium reconstruction problem. Then this flux is
written as
![]() |
(1.23) |
Due to this relation, is often called the
poloidal magnetic flux per radian (SI unit: web/rad), or simply
“the poloidal flux”. This relation allows the poloidal flux
measurements to be used to constrain the GS equation in the equilibrium
reconstrunction (discussed in Sec. 5). Note that the
positive normal direction of the surface (where the magnetic flux
is defined) is chosen in the
direction. This sign choice along with the
factor appearing in Eq. (1.23) are often confusing people
when they try to relate the experimentally measured flux with the
appearing in the GS equation. The
factor also appears when we relate the Green function of
to the mutual inductance, i.e.,
(discussed later).
A further compliction is that when one talks about “the poloidal
magnetic flux of a magnetic surface”, there are two possibilities
of the defintion. One of them is defined relative to the (as is discussed above), and another is relative to the
magnetic axis. The first definition is the flux through the central hole
of the magnetic surface, i.e., the poloidal flux through
in Fig. 1.2. In this case, as is discussed
above, the poloidal magnetic flux is related to
by
![]() |
(1.24) |
where the postive direction is .
The is the more often used defintion, and we will stick to this in the
equilibrium reconstruction problem.
The second definition is the flux enclosed by the closed flux surface,
i.e., the flux through the toroidal ribbon .
Denote this flux by
, then
![]() |
(1.25) |
where is the value of
at
the magnetic axis, the
orientation is in the
clockwise direction when an observer looks along the direction of
.
By measuring the voltage around a toroidal wire loop, we can obtain the time derivative of the poloidal flux and, after integrating over time, the flux itself.
Suppose that there is a toroidal wire loop located at
and denote the magnetic flux through the loop by
(only the poloidal magnetic field contribute to this flux, hence this
flux is called the poloidal flux). Then Faraday's law
![]() |
(1.26) |
where the direction of the loop integration and the direction of is related by the right-hand rule, is written as
![]() |
(1.27) |
where is often called electromotive force (emf).
If the loop is a coil with
turns, the induced
voltage
in the coil is
times the emf, i.e.,
. Then
Eq. (1.27) is written as
![]() |
(1.28) |
Integrating the above equation over time, we obtain
![]() |
(1.29) |
The starting time can be chosen as when
is easy to know (e.g., when there is no plasma).
Equation (1.29) tells us how to calcualte
from the measured loop voltage
.
Then
is obtained by
.
There are usually many flux loops (e.g. 35 on EAST[32]) at different locations in the poloidal plane (see Fig. 1.4). They are outside of the plasma region and thus are “external magnetic measurements”. The measured poloidal flux, along with the poloidal field measurement by magnetic probes, can be used as constraints in reconstructing the magnetic field within the plasma region. This is discussed in Sec. 5.
![]() |
A magnetic field line on a closed magnetic surface travel a closed curve
in the poloidal plane. For these field lines, we can define the safety
factor : the number of
toroidal loops a magnetic field line travels when it makes one poloidal
loop, i.e.
![]() |
(1.30) |
where as the change of the toroidal angle when a
magnetic field line travels a full poloidal loop.
For open field line region (where a field line touches the wall before its poloidal projection can close itself), the “connection length” is often used to characterize the magnetic field.
The equation of magnetic field lines is given by
![]() |
(1.31) |
where is the line element along the direction of
on the poloidal plane. Equation (1.31)
can be arranged in the form
![]() |
(1.32) |
which can be integrated over to give
![]() |
(1.33) |
where the line integration is along the poloidal magnetic field (the
contour of on the poloidal plane). Using this,
Eq. (1.30) is written
![]() |
(1.34) |
The safety factor given by Eq. (1.34) is expressed in terms
of the components of the magnetic field. The safety factor can also be
expressed in terms of the magnetic flux. Define
as the poloidal magnetic flux enclosed by two neighboring magnetic
surface, then
is given by
![]() |
(1.35) |
where is the length of a line segment in the
poloidal plane between the two magnetic surfaces, which is perpendicular
to the first magnetic surface (so perpendicular to the
). Note that
,
as well as
and
,
generally depends on the poloidal location whereas
is independent of the poloidal location.
Using Eq. (1.35), the poloidal magnetic field is written as
![]() |
(1.36) |
Substituting Eq. (1.36) into Eq. (1.34), we obtain
![]() |
(1.37) |
We know is a constant independent of the
poloidal location, so
can be taken outside the
integration to give
![]() |
(1.38) |
It is ready to realise that the integral appearing in Eq. (1.38)
is the toroidal magnetic flux enclosed by the two magnetic surfaces,
. Using this, Eq. (1.38)
is written as
![]() |
(1.39) |
Equation (1.39) indicates that the safety factor of a magnetic surface is equal to the differential of the toroidal magnetic flux with respect to the poloidal magnetic flux enclosed by the magnetic surface.
If the safety factor of a magnetic surface is a rational number, i.e.,
, where
and
are integers, then this magnetic surface is
called a rational surface, otherwise an irrational surface. A field line
on a rational surface with
closes itself after
it travels
poloidal loops. An example of a field
line on a rational surface is shown in Fig. 1.5.
In the above, the magnetic field is assumed to be axisymmetric. With
this assumption, the poloidal magnetic field (having two components) can
be expressed in terms of a single component of the vector potential
,
(specifically via
). This
kind of simplification is not achievable if the axisymmetricity
assumption is dropped, because other components of the vector potential
(namely
and
)
will appear in the expression of the poloidal magnetic field. Let us
re-examine Eq. (1.2) for a general magnetic perturbation:
When studying tearing modes and electromagnetic turbulence, most authors
narrow the possible perturbations by setting , i.e.,
![]() |
(2.2) |
![]() |
(2.3) |
![]() |
(2.4) |
where . Therefore this kind
of magnetic perturbation can still be written in the same form as the
equilibrium poloidal magnetic field:
![]() |
(2.5) |
The above approximation is widely used in practice, e.g., in turbulence
simulation, where is replaced by
. (Do we miss some magnetic perturbations that
is important for plasma transport when using the above specific form?)
The total magnetic field is then written as
![]() |
(2.6) |
**check**Can the projection of the total magnetic field line in the
poloidal plane can be traced by tracing the contour of ? No. The contours of
will not show island structures in the poloidal plane. To show the
expected island structures, we need to subtract non-reconnecting
poloidal magnetic field from the total poloidal field? **check** The
contours of the so-called helical flux will give the expected island
structures near the resonant surfaces?**check**
Next, we go back to discuss the 2D case (i.e., assuming axisymmetry).
When the displacement current term is neglectable (the case we consider here), the conductive current is just another representation of the magnetic field. Specifically, the current density is proportional to the curl of the magnetic field (Ampère's law):
where is vacuum magnetic permeability.
Use Eq. (3.1) and the definition , the poloidal components of the current density,
and
,
can be written as
![]() |
(3.2) |
and
![]() |
(3.3) |
respectively.
Ampere's law (3.1) indicates the toroidal current density
is given by
Define by
![]() |
(3.5) |
then Eq. (3.4) is written as
![]() |
(3.6) |
(Many authors incorrectly refer to Eq. (3.6) as the
Grad-Shafranov (GS) equation. Eq. (3.6) is just Ampere's
law, which has nothing to do with the force-balance. Only after we
express in terms of the plasma pressure, can Eq.
(3.6) be called the GS equation, as is discussed Sec. 4.3.)
The operator is different from the Laplacian
operator
. In terms of the
nabla operator
, the operator
is writen as
![]() |
(3.7) |
Let us consider what constraint the force balance imposes on the axisymmetric magnetic field discussed above. The MHD momentum equation is given by
![]() |
(4.1) |
where ,
,
,
,
, and
are mass density,
charge density, thermal pressure tensor, current density, electric
field, and magnetic field, respectively. The electric field force
is usually ignored due to either
or
. Further assume that
there is no plasma flow (
,
the flow effect is discussed in A.13) and the plasma
pressure is isotropic, then the steady state momentum equation (force
balance equation) is written as
![]() |
(4.2) |
where is the scalar plasma pressure.
Is the force balance (4.2) always satisfied in a real
toakamak discharge? To answer this question, we need to go back to the
original momentum equation (4.1). The imbalance between
and
will give rise to
the compressional Alfven waves, the time-scale of which,
, is much shorter than the time-scale
we are interested in. Therefore, on the time scale
(and for slow flow with
,
where
is the the sound speed), the leading order
of the momentum equation is the force balance (4.2).[26]. I.e. the inertial effect can be neglected (plasma mass is
approximately zero).
Consider the force balance in the direction of . Dotting the equilibrium equation (4.2)
by
, we obtain
![]() |
(4.3) |
which implies that is constant along a magnetic
field line. Since
is also constant along a
magnetic field line,
can be expressed in terms
of only
on a single magnetic line. Note that
this does not necessarily mean
is a
single-valued function of
,
(i.e.
). This is because
still has the freedom of taking different values on
different magnetic field lines with the same value of
while still satisfying
. This
situation can appear when there are saddle points (X points) in
contours (refer to Sec. A.22) and
takes different functions of
in islands of
sepearated by a X point. For
pressure within a single island of
,
can be safely assumed.
On the other hand, if , then
we obtain
i.e., Eq. (4.3) is satisfied, indicating
is a sufficient condition for the force balance in the parallel (to the
magnetic field) direction.
Consider the force balance in the toroidal direction. The component of Eq. (4.2) is written
![]() |
(4.4) |
Since , which implies
, equation (4.4)
reduces to
![]() |
(4.5) |
Using the expressions of the poloidal current density (3.2) and (3.3) in the force balance equation (4.5) yields
![]() |
(4.6) |
which can be further written
![]() |
(4.7) |
According to the same reasoning for the pressure, we conclude that is a sufficient condition for the toroidal force
balance. (The function
defined here is usually
called the “poloidal current function” in tokamak
literature. The reason for this name is discussed in Sec. A.11.)
Consider the force balance in direction. The
component of Eq. (4.2) is written
![]() |
(4.8) |
Using the expressions of the current density and magnetic field [Eqs. (1.6) and (3.3)], equation (4.8) is written
![]() |
(4.9) |
Assuming the sufficient condition discussed above, i.e., and
are a function of only
, i.e.,
and
, Eq. (4.9) is
written
![]() |
(4.10) |
which can be simplified to
![]() |
(4.11) |
which is the requirement of force-balance along the major radius. On the
other hand, we know can be expressed in
via Eq. (3.6). Combining this with Eq.
(4.11) yields
![]() |
(4.12) |
i.e.,
![]() |
(4.13) |
Equation (4.13) is known as Grad-Shafranov (GS) equation.
[Note that the component of the force balance
equation is written
which turns out to be identical with the Grad-Shafranov equation. This
is not a coincidence. The reason is that the force balance equation has
been satisfied in three different directions (namely, ,
,
and
direction) and thus it must be satisfied in
all the directions.]
A general axisymmetric magnetic field (which does not necessarily satisfy the force balance), is given by Eq. (1.10), i.e.,
![]() |
(4.14) |
For the above axisymmetric magnetic field to be consistent with the
force balance equation (4.2), there are additional
requirements for and
. Specifically,
is
restricted by the GS equation and
should be a
function of only
. Therefore
an axisymmetric equilibrium magnetic field is fully determined by two
functions,
and
.
The
is determined by solving the GS equation
with specified RHS source terms and boundary conditions.
The RHS source terms in the GS equation (4.13) are and
, both
of which must be specified before the GS equation can be solved. For
most cases, the source terms are nonlinear about
and thus the GS equation is a two-dimensional (in
and
) nonlinear partial
differential equation for
.
For most choices of and
, the GS equation (4.13) has to be
solved numerically. For some particular choices of
and
profiles, analytical solutions are
available, one of which is the Solovév equilibrium and is
discussed in Appendix A.12.
Note that we solve the GS equation in order to obtain the poloidal
magnetic flux and thus the poloidal magnetic
field. The toroidal magnetic field must be specified in some way before
we can solve the GS equation. There are several ways of specifying the
toroidal magnetic field: (1) given
,
(2) given
, (3) given the
safety factor
. There are
simple relations between
,
, and
, which allows translation form one to another
(discussed later). In transport simulations,
is
obtained from current drive models and neoclassical bootstrap current
models. Note that the specification of the source terms (
,
,
, and
) usually involve the unknown
(via not only the explicit presence of
,
but also the flux-surface averaging which implicitly involves
). This indicates that iterations
are needed when numerically solving the GS equation.
A routine in tokamak operation is to reconstruct magnetic field under the constraints of MHD force balance and expermental measurements. This kind of task can be done by various codes, e.g., EFIT. I developed a similar code, heq (github.com/youjunhu/heq). I will discuss details on this code.
Ampere's law in Eq. (3.6) can be generalized to include discrete toroidal currents:
![]() |
where is the plasma toroidal current density,
is a toroidal curent filament located at
,
is
Dirac's delta function. The solution to the above equation is given by
![]() |
(5.2) |
where is the fundamental solution given by
which is often called Green's function in this context (which is obtained by using a formula similar to the Biot-Savart Law, see B.14).
If all the toroidal currents (plasma current + coil/vessel curents) are
known, then can be calculated by using Eq. (5.2). Unfortunately,
is usually
unknown. Meanwhile the force-balance indicates that
can be expressed in terms of
via Eq. (4.11),
i.e.,
![]() |
(5.4) |
Substituting this into Eq. (5.2) gives an implicit formula
for , which can be iterated
(with an initial guess of
,
and assuming the function forms,
and
, as well as the coil currents
, are known ). We can see
that this is a Picard iteration. Will the iteration converge? We do not
know for sure. Numerical experiments indicate it does in some cases. Let
us discuss some specific cases. We assume that the 1D functions,
and
, can
be modeled as (EFIT's model[17]):
![]() |
(5.5) |
![]() |
(5.6) |
where
![]() |
(5.7) |
and
are values of
at the magnetic axis and LCFS, the coefficients
and
are to be determined,
and
are integers chosen
by users. Expressions (5.5) and (5.6)
guarantee that
and
are
zero at and outside the LCFS, and thus no plasma current there. Note
that expressions (5.5) and (5.6) are nonlinear
functions of
even for
. This is because the unknowns,
and
, appear in the
denominator of expression (5.7). Another nonlinearity is
related to the fact that the unkonw
determines
where the LCFS is and thus determines the region where the current
desnsity is set to zero. This also gives rise to the name
“free-boundary” for this kind of problems.
Choosing an initial guess of on gridpoints, we
can get the values of
,
, and
by
solving a least square problem that minimises the difference between
quantities computed and the corresponding quantities measured in actual
experiments. (Details are given in Sec. 5.1.)
After the coefficients and
are obtained,
can be updated by using Eqs. (5.4)-(5.6). Then, we use the latest
and
in Eq. (5.2) to
re-compute the values of
on gridpoints by using
Eq. (5.2). The procedure is repeted until convergecne in
. This is called Picard
iteration. (Alternatively, the values of
on the
inner gridpoints can be updated by inverting the Laplace operator
, see Sec. 5.4. The
values of
on the bounary still have to be
obtained by using Eq. (5.2).)
I implemented the above method in a Python code (HEQ https://github.com/youjunhu/heq). The following subsections discuss details of the code: Sec. 5.1 discusses the least square problem. Sec. 5.3 discusses the vertical displacement instability stabilizer. Without the stabilizer, the Picard iteration diverges for many elongated configurations, due to the vertical displacement instability.
For , plugging expressions
(5.5) and (5.6) into (5.4), we
obtain
For notation ease, define the following basis functions:
![]() |
(5.9) |
and the corresponding expansion coefficients
![]() |
(5.10) |
then Eq. (5.8) is written as
![]() |
(5.11) |
where . Plugging expression
(5.11) into Eq. (5.2), we obtain
Collect all the free parameters as a column vector: . Denote the size of vector
by
, which is the number of
free parameters, i.e.,
, then
the rhs of Eq. (5.12) can be written as matrix-vector
product,
. Denote the total
number of measurements of the poloidal flux by
. Then matrix elements of
for
are given by:
![]() |
(5.13) |
for , and
![]() |
(5.14) |
for . Here
is the location where the
flux measurement is
made.
The matrix , which is of shape
, is called “response matrix”. (In
least square problems, this matrix is called “design matrix'.) The
value of
is equal to number of
measurements/constraints included (measurements/constraints are merged
to
by row stacking). We will include
measurements of the poloidal flux,
measurements of the poloidal magnetic field, 1 measurement of plasma
current, a constraint for the on-axis safety factor. Then
. (For EAST,
.)
Next, we consider the poloidal field measurement:
![]() |
(5.15) |
where is the location of the No. i probe,
is a unit vector denoting the positive normal
direction of the No.
magnetic probe,
,
.
can be inferred from
via
where
![]() |
(5.17) |
and and
is given by Eq.
(B.18) and (B.19). The rhs of Eq. (5.16)
indicates that the matrix elements of
are given
by:
![]() |
(5.18) |
for (), and
![]() |
(5.19) |
for (). I place the
magnetic field equations after the flux equations, i.,e the row number
is in the range
.
Next, we consider the constraint of plasma current. The plasma current
can be inferred from
via
![]() |
(5.20) |
The rhs of Eq. (5.20) indicates that the matrix elements of
are given by:
![]() |
(5.21) |
for (), and zero for (
). This equation is placed after
the equations of flux and magnetic field measurement, i.e., its row
number is
.
Let us consider the constraint of the on-axis safety factor . A formula for computing
is given by:
![]() |
(5.22) |
where
![]() |
(5.23) |
is the ellipticity (elongation) at the magnetic axis. Organizing Eq. (5.22) as
![]() |
(5.24) |
and using
![]() |
(5.25) |
equation (5.24) is written as
![]() |
(5.26) |
This provides one linear equation for ,
with the matrix elements given by
![]() |
(5.27) |
for (), and zero for (
).
In HEQ, the ellipticity is calculated
by computing the elongation of the innermost magnetic surface found by
the contour program.
Step 0. Initial guess of :
with
.
Step 1. Search for the magnetic axis and LCFS of
so that we get the values of
at those locations.
These values are needed in computing
.
The location of the LCFS are also needed because we need to set the
plasma current to zero outside the LCFS.
Step 2. Compute elements of response matrix
using
.
Step 3. Solve linear least-square problem to get .
Step 4. Update using the latest plasma and coil
currents:
or in its discrete form:
Step 5: and goto Step 1.
To stabilize the vertical displacement instability in the Picard
iteration, we add two virtual coils that carry oppositive currents
(denoted by and
),
and are up-down symmetric, located at
and
, respectively. And set the current
by
where is the location of plasma current center
defined by
is the
generated by all
the real PF coils,
is a positive constant chosen
by users. Note that
corresponds to that the
generated by the virtual coils exactly cancles the
generted by all the real coils. Usually we need
value larger than 2, i.e., over compensation. The virtual coils
locatons,
and
,
are usually chosen near the center of the top/bottom boundary of the
computational box.
If we want to invert the finite-difference version of the Laplace
operator to solve the GS equation for
, we encounter two issues: (1) the
RHS of the GS equation involves 2 unknown functions
and
, in addition to the main
unknown function
; (2) the
boundary condition needs to be given: i.e., the values of
on a rectangular computational boundry need to be
provided.
To address issue (1): function forms of and
are chosen by users, often as polymials in
with coefficients that are assumed known (which are
later obtained in the least-square method to make the resulting solution
approximately match some measurements, as is discussed in the previous
section).
To address issue (2): The value of on the
rectangular boundary is updated using the Green's function method
(assume that external coil currents are given), as is discussed in the
previous section.
Guess values of on both the inner gridpoints and
boundary points. After this, we can invert
to
update
within the computational box, and use the
Green's function method to update values of
on
the boundary. Then iterate to converge.
Two iterations: one for values on the inner
gridpoints, one over
values on the computational
boundary.
External PF coil currents enter the problem via the boundary condition:
specifically via its contribution to on the
boundary.
The above procedure is just a minor modification from the pure Green's function algorithm discussed in Sec. 5.2, i.e., replacing the Green's function method for the inner gridpoints with the finte-difference Laplace operator inverting.
For reference easy, the finite-difference schemes for the Laplace operator is listed below.
![]() |
(5.28) |
Method 1: discretize Eq. (5.29) as
![]() |
|||
![]() |
![]() |
which can be further discretized as
![]() |
|||
![]() |
![]() |
which can be organized as
where ,
,
,
. Method 2: discetize Eq. (5.30)
which can be organized as
where
,
,
My numerical experiments indicate this scheme is more stable than than the first one (for the same grid size, and without the VDE stabilizer, the first scheme blow up while the latter scheme works well).
Semi-free boundary equilibrium problems refer to the problems where the
LCFS shape and the value of on it are given, and
one is asked to solve the currents in the PF coils. This is similar to
the fitting problem discussed above: we consider the control points on
the LCFS as virtual flux loops that measure the poloidal flux. This mode
can be used to set the feedforward PF coil currents when a target plasma
state time evolution is given.
Using magnetic measurements as input, a realtime equilibrium
reconstruction code (e.g., rtefit) can infer the inner
structure of (observer). On the other hand, PF
coil currents can provide actions on
(actuactor). With the observer and actuator, we can feedback controll
the magnetic configuration (flux map).
For example, consider controlling the shape of LCFS (often called
iso-flux control). We choose a target surface. To make sure that the
target surface is a magnetic surface, the value of
on it should be made a spatial constant. Dentoe this constant by
. To make sure that
there is no closed flux surface outside the target surface, we select
the target surface as a surface that is tangential to the first wall
(for limiter configuration) or there is a X-point on it (for divertor
configuration). Then discretize the target surface by a series of
discrete coordinates
with
. The actions (coil current changes) are chosen to
minimise the following residual:
![]() |
(6.1) |
where
![]() |
(6.2) |
![]() |
(6.3) |
the superscritp (old) indicates values before the control actions, is obtained by a real-time equilibrium reconstruction
code. This assumes that the plasma current density remains the same when
the actions are imposed (i.e., no plasma response is included), so that
only the coil current changes contribute to the
change.
If we need to feedback control points to desired
positions, then the above sum can be extended to include the following:
![]() |
|||
![]() |
![]() |
where are desired poistions of X points.
Faraday's law
![]() |
(7.1) |
can be written in the integral form:
![]() |
(7.2) |
(The direction of the loop integration is related to the surface orientation by the right-hand rule.) I.e.,
![]() |
(7.3) |
where is the flux linked with the loop.
For the No. coil, the time derivative of the
flux through it can be expressed as
![]() |
(7.4) |
where is the mutual indctance between the No.
coil and No.
coil,
is the mutual inductance between the No.
coil and a plasma filament located at
, and
is the current
carried by the filament. The integration is over a fixed region (time
independent) that is large enough to include all the plasma and small
enough to not include any coil.
Define the effective mutual inductance between the plasma and the No.
coil by
![]() |
(7.5) |
where is the plasma current, then expression (7.4) is written as
![]() |
![]() |
![]() |
|
![]() |
![]() |
where the term proportional to the time derivative of
is ignored (is this a good approximation?). Including the above emf in
the circuit equation, we get
![]() |
(7.7) |
where is the resistence of the coil,
is the current in the present coil,
is the voltage imposed on the coil by the power supply. (Here the
positive direction of
and
must be chosen consistent with the loop integration discussed above
i.e., anti-clockwise in top view.)
The VDE controlling coils are made of two coils (one upper and one
lower, with updown symmetry), which are anti-conneted in series and
powered by one voltage source. For the upper
coil, the circuit equation is
![]() |
(7.8) |
Similarly, for the lower ()
coil, the circuit equation is
![]() |
(7.9) |
The anti-series connection implies that
![]() |
(7.10) |
and
![]() |
(7.11) |
where is the source voltage (assuming the
voltage source positive terminal is connected to the upper coil that
goes anti-clockwise in top view). Combining these equations, we obtain
![]() |
(7.12) |
This is used as an equation for the upper VDE coil current. Eq. (7.10) is used as an equation for the lower VDE coil current.
Because the mutual inductance between two objects is symmetric (i.e.,
), the mutual inductance
given by expression (7.5) can also be
used to calculate the effective emf induced in the plasma by the coils:
(Here can be understood as
weighted average of
. Why to
use
weighted, rather than just equal-weighted
spatial average? This is to ensure energy conservation: the power is
, so the current density
should serve as a weight.)
Similarly, the effective emf self-induced in the plasma is given by
![]() |
(7.13) |
where is the mutual inductance between
filament and
filament; both the 2D
integrations are over a fixed region that includes the plasma but does
not include any coil; one of the two 2D integrations can be regarded as
weighted average and the other is the
integration over the source. Expresion (7.13) motivates us
to define the effective self-inductance of plasma by
![]() |
(7.14) |
Then the effective emf in expression (7.13) is written as
where the term proportional to the time derivative of
is ignored. Then the circuit equation for the plasma current is written
as
![]() |
(7.15) |
where is the effective plasma resistivity.
Using the Euler implicit scheme, the coil current evolution equation
() is written as
Moving all the unknown (coil currents and plasma
current
) to the left-hand
side, we get
![]() |
![]() |
||
![]() |
![]() |
For each coil, we get a linear equation like the above. For coils, we get
equations. (Note
that
depends on
,
and thus is time dependent. If we use value of
at
in the above equation, then we need to
iterate because
is unkown. If we use the value
at
, then no iteration is
needed, whch is the case used in HEQ).
Similarly, the plasma current evolution equation (7.15) is discretized as
![]() |
(7.17) |
i.e.,
![]() |
(7.18) |
Define the plasma state by the position of the magnetic axis , total plasma current
, and the shape of the LCFS. Discretize the
LCFS in the cylindrical coordinates by
points.
Then the plasma state can be represented by
real
numbers:
,
,
,
and
for
.
Inputs to the policy network: the magnetic measurements at the present
time step (which serves as a proxy to the actual
plasma state), and target plasma states (provided by the designer) at
(where
may be a small
integer, say
).
Output of the policy network: voltage command to the coils imposed at
which can hopefully make the plasma follow the
target state in future time steps.
How to train the policy network on a plasma simulator?
In many studies of tokamak plasmas, one need construct a curvilinear
coordinate system based on a given magnetic cofiguration in order to
make the problem amenable to analytical methods or numerical methods.
Specifically, many theories and numerical codes use the curvilinear
coordinate systems that are constructed with one coordinate surface
coinciding with magnetic surfaces. In these coordinate systems, we need
to choose a poloidal coordinate and a toroidal
coordinate
. A particular
choice for
and
is one
that makes the magnetic field lines be straight lines in
plane. These kinds of coordinates are often called
magnetic coordinates. That is, “magnetic coordinates are defined
so they conform to the shape of the magnetic surfaces and trivialize the
equations for the field lines.”
A further tuned magnetic coordinate system is the so-called field aligned (or filed-line following) coordinate system, in which changing one of the three coordinates with the other two fixed would correspond to following a magnetic field line. The field aligned coordinates are discussed in Sec. 15.
Next, let us discuss some general properties about coordinates transformation[4].
In the Cartesian coordinates, a point is described by its coordinates
, which, in the vector form,
is written as
![]() |
(8.1) |
where is the location vector of the point;
,
,
and
are the basis vectors of the Cartesian
coordinates, which are constant, independent of spactial location. The
transformation between the Cartesian coordinates system,
, and a general coordinates system,
, can be expressed as
![]() |
(8.2) |
For example, cylindrical coordinates can be
considered as a general coordinate systems, which are defined by
.
The transformation function in Eq. (8.2) can be written as
![]() |
(8.3) |
A useful quality characterizing coordinate transformation is the Jacobian determinant (or simply called Jacobian), which, for the transformation in Eq. (8.3), is defined by
![]() |
(8.4) |
which can also be written as
![]() |
(8.5) |
It is easy to prove that the Jacobian in Eq. (8.5) can also be written (the derivation is given in my notes
on Jacobian)
![]() |
(8.6) |
Conventionally, the Jacobian of the transformation from the Cartesian
coordinates to a particular coordinate system is
called the Jacobian of
,
without explitly mentioning that this transformation is with respect to
the Cartesian coordinates.
Using the defintion in Eq. (8.4), the Jacobian of the Cartesian coordinates can be calculated, yielding
. Likewise, the Jacobian of
the cylindrical coordinates
can be calculated as
follows:
If the Jacobian of a coordinate system is greater than zero, it is called a right-handed coordinate system. Otherwise it is called a left-handed system.
In a curvilinear coordinate system ,
there are two kinds of basis vectors:
and
, with
These two kinds of basis vectors satisfy the following orthogonality
relation:
![]() |
(8.7) |
where is the Kronical delta function. [Proof:
Working in a Cartesian coordinate system
with
the corresponding basis vectors denoted by
,
then the left-hand side of Eq. (8.7) can be written as
where the second equality is due to since
are constant vectors independent of spatial location;
the chain rule has been used in obtaining Eq. (8.8)]
[The cylindrical coordinate system is an example
of general coordinates. As an exercise, we can verify that the
cylindrical coordinates have the property given in Eq. (8.7).
In this case,
,
,
,
where
,
,
.]
It can be proved that is a contravariant vector
while
is a covariant vector (I do not prove this
and do not bother with the meaning of these names, just using this as a
naming scheme for easy reference).
The orthogonality relation in Eq. (8.7) is fundamental to
the theory of general coordinates. The orthogonality relation allows one
to write the covariant basis vectors in terms of contravariant basis
vectors and vice versa. For example, the orthogonality relation tells
that is orthogonal to
and
, thus,
can be written as
![]() |
(8.9) |
where is a unknown variable to be determined. To
determine
, dotting Eq. (8.9) by
, and using
the orthogonality relation again, we obtain
![]() |
(8.10) |
which gives
![]() |
![]() |
![]() |
|
![]() |
![]() |
Thus is written, in terms of
,
,
and
, as
![]() |
(8.12) |
Similarly, we obtain
![]() |
(8.13) |
and
![]() |
(8.14) |
Equations (8.12)-(8.14) can be generally written
![]() |
(8.15) |
where represents the cyclic order in the
variables
. Equation (8.15) expresses the covariant basis vectors in terms of the
contravariant basis vectors. On the other hand, from Eq. (8.12)-(8.14), we obtain
![]() |
(8.16) |
which expresses the contravariant basis vectors in terms of the covariant basis vectors.
Suppose is an arbitrary general coordinate
system. Following Einstein's notation, contravariant basis vectors are
denoted with upper indices as
![]() |
and the covariant basis vectors are denoted with low indices as
![]() |
Then the orthogonality relation, Eq. (8.7), is written as
![]() |
(8.19) |
In term of the contravairant basis vectors, is
written
![]() |
(8.20) |
where the components are easily obtained by taking scalar product with
, yielding
,
,
and
. Similarly, in term of
the covariant basis vectors,
is written
![]() |
(8.21) |
where ,
, and
.
Using the above notation, the relation in Eq. (8.15) is written as
![]() |
(8.22) |
![]() |
(8.23) |
![]() |
(8.24) |
where . Similarly, the
relation in Eq. (8.16) is written as
![]() |
(8.25) |
![]() |
(8.26) |
![]() |
(8.27) |
The gradient of a scalar function is readily
calculated from the chain rule,
![]() |
(8.28) |
Note that the gradient of a scalar function is in the covariant representation. The inverse form of this expression is obtained by dotting the above equation respectively by the three contravariant basis vectors, yielding
![]() |
(8.29) |
![]() |
(8.30) |
![]() |
(8.31) |
Using Eq. (8.28), the directional derivative in the
direction of is written as
![]() |
(8.32) |
To calculate the divergence of a vector, it is desired that the vector should be in the contravariant form because we can make use of the fact:
![]() |
(8.33) |
for any scalar quantities and
. Therefore we write vector
as
![]() |
(8.34) |
where ,
,
.
Then the divergence of
is readily calculated as
where the second equality is obtained by using Eqs. (8.29), (8.30), and (8.31).
The Laplacian operator is defined by .
Then
is written as (
is
an arbitrary function)
To proceed, we can use the divergence formula (8.36) to express the divergence in the above expression. However, the vector in the above (blue term) is not in the covariant form desired by the divergence formula (8.36). If we want to directly use the formula (8.36), we need to transform the vector (blue term in expression (8.37)) to the covariant form. This process seems to be a little complicated. Therefore, I choose not to use this method. Instead, I try to simplify expression (8.37) by using basic vector identities:
![]() |
![]() |
![]() |
|
![]() |
Using the gradient formula, the above expression is further written as
![]() |
![]() |
![]() |
|
![]() |
|||
![]() |
|||
![]() |
and can be simplified as
![]() |
![]() |
![]() |
|
![]() |
|||
![]() |
|||
![]() |
Assume are field-line following coordinates with
along the field line direction, then neglect all
the parallel derivatives, i.e., derivative over
, then the above expression is reduced to
![]() |
![]() |
![]() |
|
![]() |
![]() |
This approximation reduces the Laplacian operator from being
three-dimensional to being two-dimensional. This approximation is often
called the high- approximation, where
is the toroidal mode number (mode number along
direction).
To take the curl of a vector, it should be in the covariant
representation since we can make use of the fact that . Thus the curl of
is
written as
Note that taking the curl of a vector in the covariant form leaves the vector in the contravariant form.
Consider a general coordinate system .
I define the metric tensor as the transformation matrix between the
covariant basis vectors and the contravariant ones. Equations (8.15)
and (8.16) express the relation between the two sets of
basis vectors using cross product. Next, let us express the relation in
matrix from. To obtain the metric matrix, we write the contrariant basis
vectors in terms of the covariant ones, such as
![]() |
(8.43) |
Taking the scalar product respectively with ,
, and
, Eq. (8.43) is
written as
![]() |
(8.44) |
![]() |
(8.45) |
![]() |
(8.46) |
Similarly, we write
![]() |
(8.47) |
Taking the scalar product with ,
, and
, respectively, the above becomes
![]() |
(8.48) |
![]() |
(8.49) |
![]() |
(8.50) |
The same situation applies for the basis vector,
![]() |
(8.51) |
Taking the scalar product with ,
, and
, respectively, the above equation becomes
![]() |
(8.52) |
![]() |
(8.53) |
![]() |
(8.54) |
Summarizing the above results in matrix form, we obtain
![]() |
(8.55) |
Similarly, to convert contravariant basis vector to covariant one, we write
![]() |
(8.56) |
Taking the scalar product respectively with ,
, and
, the above equation becomes
![]() |
(8.57) |
![]() |
(8.58) |
![]() |
(8.59) |
For the second contravariant basis vector
![]() |
(8.60) |
![]() |
(8.61) |
![]() |
(8.62) |
![]() |
(8.63) |
For the third contravariant basis vector
![]() |
(8.64) |
![]() |
(8.65) |
![]() |
(8.66) |
![]() |
(8.67) |
Summarizing these results, we obtain
![]() |
(8.68) |
where
This matrix and the matrix in Eqs. (8.55) should be the inverse of each other. It is ready to prove this by directly calculating the product of the two matrix.
Suppose that are arbitrary general coordinates
except that
is the usual toroidal angle in
cylindrical coordinates. Then
is perpendicular
to both
and
.
Using this, Eq. (8.55) is simplified to
![]() |
(8.69) |
Similarly, Eq. (8.68) is simplified to
![]() |
(8.70) |
[Note that the matrix in Eqs. (8.69) and (8.70) should be the inverse of each other. The product of the two matrix,
![]() |
(8.71) |
can be calculated to give
where
By using the definition of the Jacobian in Eq. (8.6), it is
easy to verify that , i.e.,
![]() |
(8.72) |
]
The axisymmetric equilibrium magnetic field is given by Eq. (4.14), i.e.,
![]() |
(8.73) |
In a general coordinate system (not necessarily
flux coordinates), the above expression can be written as
![]() |
(8.74) |
where the subscripts denote the partial derivatives with the corresponding subscripts. Note that Eq. (8.74) is a mixed representation, which involves both covariant and contravariant basis vectors. Equation (8.74) can be converted to the contravariant form by using the metric tensor, giving
![]() |
(8.75) |
Similarly, Eq. (8.74) can also be transformed to the covariant form, giving
![]() |
(8.76) |
For the convenience of notation, define
![]() |
(8.77) |
then Eq. (8.76) is written as
![]() |
(8.78) |
A coordinate system , where
is the usual cylindrical toroidal angle, is
called a magnetic/flux coordinate system if
is a
function of only
, i.e.,
(we also have
since we are
considering axially symmetrical case). In terms of
coordinates, the contravariant form of the magnetic field, Eq. (8.75),
is written as
![]() |
(8.79) |
where . The covariant form of
the magnetic field, Eq. (8.76), is written as
![]() |
(8.80) |
The local safety factor is defined by
![]() |
(8.81) |
which characterizes the local pitch angle of a magnetic field line in
plane of a magnetic surface. Substituting the
contravariant representation of the magnetic field, Eq. (8.79),
into the above equation, the local safety factor is written
![]() |
(8.82) |
Note that the expression in Eq. (8.82)
depends on the Jacobian
.
This is because the definition of
depends on the
definition of
, which in turn
depends on the the Jacobian
.
In terms of , the
contravariant form of the magnetic field, Eq. (8.79), is
written
![]() |
(8.83) |
and the parallel differential operator is
written as
![]() |
(8.84) |
If happens to be independent of
(i.e., field lines are straight in
plane), then
the above operator becomes a constant coefficient differential oprator
(after divided by
). This
simplification is useful because different poloidal harmonics are
decoupled in this case. We will discuss this issue futher in Sec. 14.
The global safety factor defined in Eq. (1.34) is actually the poloidal average of the local safety factor, i.e.,
Note that and
defined
this way can be negative, which depends on the choice of the positive
direction of
and
coordinates (note that the safety factor given in G-eqdsk file is always
positive, i.e. it is the absolute value of the safety factor defined
here).
Next, let us transform the integration in
expression (8.86) to a curve integral in the poloidal
plane. Using the relation
and
[Eq. (8.94)], expression (8.86) is further
written
Expression (8.87) is used in the GTAW code to numerically
calculate the value of on magnetic surfaces (as
a benchmarking of the
profile specified in the
G-eqdsk file). Expression (8.87) can also be considered as
a relation between
and
. In the equilibrium problem where
is given (fixed-q equilibrium), we can use expression (8.87)
to obtain the corresponding
(which explicitly
appears in the GS equation):
![]() |
(8.88) |
We note that expression (8.88) involves magnetic surface
averaging, which is unknown before we know .
Therefore iteration is usually needed in solving the fixed-q equilibrium
(i.e., we guess the unknown
,
so that the magnetic surface averaging in expression (8.88)
can be performed, yielding the values of
.)
Using and
,
the absolute value of
in expression (8.87)
is written
![]() |
![]() |
![]() |
|
![]() |
![]() |
which is the familiar formula we see in textbooks.
Given the definition of a magnetic surface coordinate system , the Jacobian of this system is
fully determined. On the other hand, given the definition of
,
,
and the Jacobian, the definition of
is fully
determined (can have some trivial shifting freedoms). Next, let us
discuss how to calculate
in this case. In
coordinates, a line element is written
![]() |
(8.91) |
The line element that lies on a magnetic surface (i.e., ) and in a poloidal plane (i.e.,
) is then written
![]() |
![]() |
![]() |
|
![]() |
![]() |
We use the convention that and
take the same sign, i.e.,
![]() |
(8.93) |
Using the fact that and
are orthogonal and
, the
above equation is written as
![]() |
(8.94) |
Given , Eq. (8.94)
can be integrated to determine the
coordinate of
points on a magnetic surface.
–—
![]() |
(8.95) |
Once is known, the value of
of a point can be obtained by integrating expression (8.94),
i.e.,
![]() |
(8.96) |
where the curve integration is along the contour ,
is a reference point on
the contour, where value of the poloidal angle is chosen as
. The choice of the positive direction of
is up to users. Depending on the positive direction
chosen, the sign of the Jacobian of the constructed coordinates can have
a sign difference from the
appearing in Eq. (8.96). Denote the Jacobian of the constructed coordinates by
, then
![]() |
(8.97) |
This sign can be determined after the radial coordinate and the positive
direction of the poloidal angle are chosen. In GTAW
code, I
choose the positive direction of to be in
anticlockwise direction when observers look along the direction of
. To achieve this, the line
integration in Eq. (8.96) should be along the anticlockwise
direction. (I use the determination of the direction matrix, a well
known method in graphic theory, to determine the direction from a given
set of discrete points on a magnetic surface.)
The span of defined by Eq. (8.96)
is usually not
in one poloidal loop. This
poloidal angle can be scaled by
to define a new
poloidal coordinate
, whose
span is
in one poloidal loop, where
is a magnetic surface function given by
![]() |
(8.98) |
Then is written as
where . Sine we have modified
the definition of the poloidal angle, the Jacobian of the new
coordinates
is different from that of
. The Jacobian
of the new coordinates
is written as
The Jacobian can be set to various forms to
achieve various poloidal coordinates, which will be discussed in the
next section. After the Jacobian and the radial coordinate
is chosen, all the quantities on the right-hand side of
Eq. (8.99) are known and the integration can be performed
to obtain the value of
of each point on each
flux surface.
[The reference points and the values of poloidal
angle at these points can be chosen by users. One choice of the
reference points
are those points on the
horizontal ray in the midplane that starts from the magnetic axis and
points to the low filed side of the device and
at these points is chosen as zero (this is my choice in the
GTAW
code). In the TEK
code, the reference
points are chosen at the high-field side of the midplane and at the reference points.]
If the Jacobian is chosen to be of the following
form
![]() |
(8.101) |
Then in Eq. (8.99) is written
![]() |
(8.102) |
and the Jacobian of new coordinates ,
, which is given by Eq. (8.100), now takes the form
![]() |
(8.103) |
Equation (8.102) indicates a set of poloidal points with
equal arc intervals corresponds to a set of uniform
points. Therefore this choice of the Jacobian is called the
equal-arc-length Jacobian. Note that Eq. (8.102) does not
involve the radial coordinate
.
Therefore the values of
of points on any
magnetic surface can be determined before the radial coordinate is
chosen.
The volume element in coordinates is given by
. If we choose a Jacobian
that is independent of
, then
uniform
grids will correspond to grids with
uniform volume interval. In this case,
is
written as
![]() |
(8.104) |
where is a function independent of
. Then
in Eq. (8.99)
is written
![]() |
(8.105) |
and the Jacobian of the new coordinates ,
, is given by Eq. (8.100),
which now takes the following form:
![]() |
(8.106) |
Note that both and
are
independent of the function
introduced in Eq.
(8.104). (
is eliminated by the
normalization procedure specified in Sec. 8.15.1 due to the
fact that
is constant on a magnetic surface.)
The equal-volume poloidal angle is also called Hamada poloidal angle.
The equal-volume poloidal angle is useful in achieving loading balance
for parallel particle simulations. Assume that markers are loaded
uniform in space and the poloidal angle is domain decomposed and
assigned to different MPI process. Then the equal-volume poloidal angle
can make marker number in each MPI process be equal to each other and
thus work loading to each process be equal. (**check**If the domain
decomposition is also applied to the radial direction, to achieve
loading balance, then the radial coordinate
should be chosen in a way that makes
be
independent of
, so that
in Eq. (8.106) is constant in space.**)
If the Jacobian is chosen to be of the Boozer
form:
![]() |
(8.107) |
then the poloidal angle in Eq. (8.99) is written as
![]() |
(8.108) |
The final Jacobian is given by
![]() |
(8.109) |
The usefullness of Boozer poloidal angle will be further discussed in Sec. 14.8 after we introduce a gneralized toroidal angle.
If the Jacobian is chosen to be of the following
form
then Eq. (8.82) implies that the local safety factor, , is a magnetic surface function,
i.e., the magnetic field lines are straight in
plane. Then the poloidal angle in Eq. (8.99) is written
![]() |
(8.110) |
The Jacobian given by Eq. (8.100)
now takes the form
![]() |
(8.111) |
Let us denote an arbitrary poloidal angle by and
the above straight-field-line poloidal angle by
, then it is ready to find the following relation
between
and
:
![]() |
(8.112) |
where is the local safety factor corresponding
to the arbitrary poloidal angle
,
i.e.,
. [Proof: Using
, the poloidal angle
given in Eq. (8.110) is written as
![]() |
![]() |
![]() |
|
![]() |
![]() |
||
![]() |
![]() |
Using , where
is the Jacobian of
coordinates,
then the above
is reduced to expression (8.112).]
Note that Boozer poloidal angle is very close to the poloidal angel disccused here because the two Jacobians are very similar:
![]() |
(8.114) |
All Jacobians introduced above can be written in a general form:
![]() |
(8.115) |
The choice of gives the PEST coordinate,
give the Boozer coordinate,
gives the equal-arc coordinate,
gives the
Hammada coordinate.
Figure 8.1 compares the equal-arc-poloidal angle and the straight-line poloidal angle, which shows that the resolution of the straight-line poloidal angle is not good near the low-field-side midplane. Since ballooning modes take larger amplitude near the low-field-side midplane, better resolution is desired there. This is one reason that I often avoid using the straight-line poloidal angle in my numerical codes.
After the magnetic coordinates are constructed, we can evaluate the
Jacobian by using directly the definition of the
Jacobian, i.e.,
![]() |
(8.116) |
which can be further written as
![]() |
(8.117) |
where the partial differential can be evaluated by using numerical
differential schemes. The results obtained by this way should agree with
results obtained from the analytical form of the Jacobian. This
consistency check provide a verification for the correctness of the
theory derivation and numerical implementation. In evaluating the
Jacobian by using the analytical form, we may need to evaluate , which finally reduces to
evaluating
. The value of
is obtained numerically based on the numerical
data of
given in cylindrical coordinate grids.
Then the cubic spline interpolating formula is used to obtain the value
of
at desired points. (
calculated by the second method (i.e. using analytic form) is used in
the GTAW code; the first methods are also implemented in the code for
the benchmark purpose.) In the following sections, for notation ease,
the Jacobiban of the constructed coordinate system will be denoted by
, rather than
.
The radial coordinate can be chosen to be
various surface function, e.g., volume, poloidal or toroidal magnetic
flux within a magnetic surface. The frequently used radial coordinates
include
, and
, where
is defined by
![]() |
(8.118) |
where and
are the values
of
at the magnetic axis and LCFS, respectively.
Other choices of the radial coordinates: the normalized toroidal
magnetic flux and its square root,
,
and
.
The volume between magnetic surfaces can also be used as a radial coordinate. The differential volume element is written as
![]() |
(8.119) |
Integrating over the toroidal angle, we obtain
![]() |
(8.120) |
Further integrating over the poloidal angle, we obtain
![]() |
(8.121) |
i.e.,
![]() |
(8.122) |
The cylindrical coordinates is a right-hand
system, with the positive direction of
pointing
vertically up. In GTAW code, the positive direction of
is chosen in the anticlockwise direction when observers look along the
direction of
. Then the
definition
indicates that (1)
is negative if
points from the magnetic axis to
LCFS; (2)
is positive if
points from the LCFS to the magnetic axis. This can be used to determine
the sign of Jacobian after using the analytical formula to obtain the
absolute value of Jacobian.
Given an axisymmetric tokamak equilibrium in
coordinates (e.g., 2D data
on a rectangular
grids
in G-file), we can construct a magnetic
surface coordinates
by the following two steps.
(1) Find out a series of magnetic surfaces on
plane and select radial coordinates for each magnetic surface (e.g. the
poloidal flux within each magnetic surface). (2) Specify the Jacobian or
some property that we want the poloidal angle to have. Then calculate
the poloidal angle of each point on each flux surface (on the
plane) by using Eq. (8.94) (if the Jacobian
is specified) or some method specified by us to achieve some property we
prefer for the poloidal angle (if a Jacobian is not directly specified).
Then we obtain the magnetic surface coordinates system
.
Two-dimensional data on a rectangular grids
is read from the G_EQDSK file (G-file) of EFIT code.
Based on the 2D array data, I use 2D cubic spline interpolation to
construct a interpolating function
.
To construct a magnetic surface coordinate system, I need to find the
contours of
, i.e., magnetic
surfaces. The values of
on the magnetic axis,
, and the value of
on the last closed flux surface (LCFS),
, are given in G-file. Using these two values,
I construct a 1D array “psival” with value of elements
changing uniform from
to
. Then I try to find the contours of
with contour level value ranging from
to
. This is done in the
following way: construct a series of straight line (in the poloidal
plane) that starts from the location of the magnetic axis and ends at
one of the points on the LCFS. Combine the straight line equation,
, with the interpolating function
, we obtain a one variable
function
. Then finding the
location where
is equal to a specified value
, is reduced to finding the
root of the equation
. Since
this is a one variable equation, the root can be easily found by using
simple root finding scheme, such as bisection method (bisection method
is used in GTAW code). After finding the roots for each value in the
array “psival” on each straight lines, the process of
finding the contours of
is finished. The
contours of
found this way are plotted in Fig.
9.1.
In the above, we mentioned that the point of magnetic axis and points on
the LCFS are needed to construct the straight lines. In G-file, points
on LCFS are given explicitly in an array. The location of magnetic axis
is also explicitly given in G-file. It is obvious that some of the
straight lines that pass through the location of
magnetic axis and points on the LCFS will have very large or even
infinite slope. On these lines, finding the accurate root of the
equation
is difficult or even impossible. The
way to avoid this situation is obvious: switch to use function
instead of
when the slope of
is large (the switch condition I used is
).
In constructing the flux surface coordinate with desired Jacobian, we
will need the absolute value of the gradient of ,
, on
some specified spatial points. To achieve this, we need to construct a
interpolating function for
.
The
can be written as
![]() |
(9.1) |
By using the center difference scheme to evaluate the partial
derivatives with respect to and
in the above equation (using one side difference scheme for the points
on the rectangular boundary), we can obtain an 2D array for the value of
on the rectangular
grids. Using this 2D array, we can construct an interpolating function
for
by using the cubic spline interpolation
scheme.
![]() ![]() |
![]() |
Metric elements of the coordinates, e.g.,
, are often needed in practical
calculations. Next, we express these metric elements in terms of the
cylindrical coordinates
and their partial
derivatives with respect to
and
. Note that, in this case, the coordinate
system is
while
and
are functions of
and
, i.e.,
![]() |
(9.2) |
![]() |
(9.3) |
Then and
are written as
![]() |
(9.4) |
![]() |
(9.5) |
wehre , etc. Equations (9.4) and (9.5) can be solved to give
![]() |
(9.6) |
![]() |
(9.7) |
Using the above expressions, the Jacobian of
coordinates,
, is written as
![]() |
![]() |
![]() |
|
![]() |
![]() |
||
![]() |
![]() |
i.e.,
![]() |
(9.9) |
Using this, Expressions (9.6) and (9.7) are written as
![]() |
(9.10) |
and
![]() |
(9.11) |
Then the elements of the metric matrix are written as
![]() |
(9.12) |
![]() |
(9.13) |
and
![]() |
(9.14) |
Equations (9.12), (9.13), and (9.14)
are the expressions of the metric elements in terms of ,
,
,
, and
.
[Combining the above results, we obtain
![]() |
(9.15) |
Equation (9.14) is used in GTAW code. Using the above
results, are written as
![]() |
(9.16) |
![]() |
(9.17) |
![]() |
(9.18) |
As a side product of the above results, we can calculate the arc length
in the poloidal plane along a constant surface,
, which is expressed as
Note that since we are considering the arc
length along a constant
surface in
plane. Then the above expression is reduced to
which agrees with Eq. (8.94).]
In some cases (e.g., turbulence simulation), model tokamak magnetic
field, which is not an exact solution to the GS equation, is often used.
In the model, the safety factor profile ,
toroidal field function
, and
the magnetic surface shape
are given. To use
this model in a simulation, we need to calculate its poloidal magnetic
field, which is determined by the poloidal magnetic flux function
. To determine
, we need to use Eq. (8.86), i.e.,
![]() |
(9.20) |
and re-organize the formula as
![]() |
(9.21) |
which can be integrated to obtain and thus the
poloidal magnetic field
,
where
. The toroidal magnetic
field can be obtained from
by
. In most papers,
is
chosen to be a constant.
A typical magnetic surface shape used in simulations is the Miller shape, which is given by
![]() |
(9.22) |
where is the radial coordinate,
,
,
and
are the Shafronov shift, triangularity, and
elongation profiles, which can be arbitrarily specified. For the special
case of
being a constant,
, and
,
this shape reduces to a concentric-circular magnetic field. An example
of calculating the poloidal field of the model field is given in Sec. 16. This kind of model field can be called theoretical
physicists' tokamak. Computational physicists often use this model for
code benchmarking purpose. The famous DIII-D cyclone base case is an
example, which was extensively used for benchmarking gyrokinetic
simulation of ion temperature driven turbulence.
Magnetic surface average of a physical quantity
is defined by
![]() |
(10.1) |
where the volume integration is over a small volume between two adjacent
flux surfaces with differing by
. [This formula (with finite
) is used in TEK code to calculate the
radial heat flux.]
The above definition can be written as
![]() |
(10.2) |
where the volume integration is over the volume enclosed by the magnetic
surface labeled by .
The above 3D volume integration can be written as a 2D surface
integration. The differential volume element is given by , where
is the Jacobian
of
coordinates. equation (10.1) is
written as
which is a 2D averaging over a magnetic surface. Noting that is independent of
,
expression (10.3) can be written as
![]() |
(10.4) |
where , which is the
harmonic of
,
where
is the toroidal mode number. Note that the
surface averaging of any
harmonic is zero. I.e.,
only the
harmonic can contribute to the magnetic
surface average.
In flux coordinates, Eq. (10.2) can be written as
where . (For the surface
average, in some parts of these notes, I assume that the Jacobian is
postive and thus the absolute value operator is dropped. In most part of
this document, axisymmetry is assumed for
,
i.e.,
.)
The volume within a magnetic surface is written
![]() |
(10.6) |
Then the differential of with respect to
is written as
![]() |
(10.7) |
Sometimes, we do not want the Jacobian to explicitly appear in the formula. This can be achieved by using Eq. (8.94), i.e.,
![]() |
(10.8) |
Then expression (10.4) is written as
(Equation (10.9) is used in GTAW code[14].)
The parallel current density appears in the flux diffusion equation. Let
us derive a formula for it in the flux coordinates. Using , along with magnetic field expression (8.80) and the curl formula (8.42), we obtain
where ,
. Expression (10.10) is the
contravariant form of
.
Taking dot product between Eqs. (8.80) and (10.10),
we obtain
![]() |
![]() |
![]() |
|
![]() |
|||
![]() |
![]() |
||
![]() |
|||
![]() |
![]() |
||
![]() |
![]() |
Using the defintion of the magnetic surface average, Eq. (10.5), i.e.,
we obtain
which agrees with Eq. (5) in Ref. [23].
This is an interesting way of defining flux within a magnetic surface using volume integration, instend of surface integration. Examine the meaning of the following volume integral:
![]() |
(11.1) |
where the volume is the volume within the
magnetic surface labeled by
.
Using
, the quantity
can be further written as
![]() |
(11.2) |
Note that is not a single-value function of the
spacial points. In order to evaluate the integration in Eq. (11.2),
we need to select one branch of
,
which can be chosen to be
.
Note that function
is not continuous at the
branch cut
. Next, we want to
use the divergence theorem to convert the above volume integration to
surface integration. Noting the discontinuity of the integrand
at
, the volume
should be cut there, thus, generating two surfaces. Denote these two
surfaces by
and
,
then equation (11.2) is written as
where the orientation of surface is in the
negative direction of
, the
orientation of
is in the positive direction of
, and the surface
is the toroidal magnetic surface
. The surface integration through
is obviously zero since
lies in this surface.
Therefore, we have
Eq. (11.3) indicates that is the
magnetic flux through the
surface. This is the
flux
defined by Eq. (1.25), i.e.,
![]() |
(11.4) |
[We can also directly verify defined by Eq. (11.4) is the poloidal flux, without using the divergence
theorem. Using the expression of the volume element
,
Note that the sign of the Jacobian appears in Eq. (11.5),
which is due to the positive direction of surface
is determined by the positive direction of
,
which in turn is determined by the sign of the Jacobian (In my code,
however, the positive direction of
is chosen by
me and the sign of the Jacobian is determined by the positive direction
of
). We can verify the sign
of Eq. (11.5) is consistent with that in Eq. (1.25).]
Similarly, the toroidal flux within a flux surface is written as
![]() |
![]() |
![]() |
|
![]() |
![]() |
In flux coordinates, is written as
from which, we can also obtain
![]() |
(11.9) |
and
![]() |
(11.10) |
Using , we obtain
![]() |
(11.11) |
Similarly, the toroidal current within a flux surface is written as
![]() |
(11.12) |
Let us express in flux coordinates:
Use Eq. (3.6), i.e.,
![]() |
(11.15) |
and Eq. (13.4), i.e.,
![]() |
(11.16) |
then is written as
Then the surface average is written as
![]() |
![]() |
![]() |
|
![]() |
![]() |
||
![]() |
![]() |
||
![]() |
![]() |
||
![]() |
![]() |
Using this in Eq. (11.14), we obtain
Faraday's law
![]() |
(12.1) |
can be written in the integral form:
![]() |
(12.2) |
where is the flux linked with the loop. Choosing
an axisymmetric loop along
,
and assuming
is axisymmetric, Eq. (12.2)
is written as
i.e.,
Using , the above equation is
written as
![]() |
(12.3) |
Use Ohm's law
(where stands for all the terms on the rhs of
Ohm's law, typically
, where
is the total current density and
is the non-inductive part), to eliminate
, then Eq. (12.3) is written as
![]() |
(12.4) |
Using to eliminate
on
the rhs, we obtain
![]() |
(12.5) |
Noting that , the above
equation is simplified to
![]() |
(12.6) |
This is an evolution equation for the poloidal flux function in the lab frame.
Next, let us derive an evolution equation for the toroidal field
function . The
projection of Faraday's law (12.1) is written
as
![]() |
(12.7) |
i.e.,
![]() |
(12.8) |
Use Ohm's law to eliminate ,
then Eq. (12.8) is written as
![]() |
(12.9) |
i.e.,
![]() |
(12.10) |
Next, let us write Eqs. (12.6) and (12.10) in
a general magnetic coordinate system that is
time-dependent (relative to the laboratory static coordinate system).
Why do we need a magnetic coordinate system? Because it makes it easy to
perform the magnetic surface average when we derive 1D transport
equations. Then why do we require the magnetic coordinates to be
time-dependent? Because the magnetic surface shapes are evolving, and
thus a static transform can not remain conform to the evolving shapes.
For a time-dependent coordinate transform:
where stands for a laboratory static Cartesian
coordinate system
, it
follows that
![]() |
![]() |
![]() |
|
![]() |
![]() |
where
![]() |
(12.12) |
is the coordinate velocity. Further define
![]() |
(12.13) |
which is the fluid velocity observed in
coordinate system.
Note that the flux evolution equation (12.6) is written in a laboratory static coordinate system, i.e.,
![]() |
(12.14) |
Then in time-dependent coordinate system, the
above equation is written as
![]() |
(12.15) |
Noting , the above equation
is simplified to
![]() |
(12.16) |
Now we restrict to be flux coordinates (i.e.,
contours of
are contours of
) and remain flux coordinates when both
and
evolve. Then the rhs of
Eq. (12.16) must be chosen to be independent of
, i.e.,
![]() |
(12.17) |
where is a function to be determined. For later
use, multiplying the above equation by
:
![]() |
(12.18) |
and performing the magnetic surface average on it, we obtain
![]() |
(12.19) |
Next, consider the evolution equation for ,
Eq. (12.10), which can be written as
![]() |
(12.20) |
We note that . Then the above
equation is written as
![]() |
(12.21) |
To facilitate later use, we multiply the above equation by the Jacobian
J of coordinates:
i.e.,
Tansforming to time-dependent coordinate system,
the above equation is written as
For notation ease, the subscripts ()
of the time derivative is dropped in the following, with the
understanding that the time derivatives are taken with (
) held fixed. Equation (12.22)
simplifies to
Noting that , the above
equation is written as
which simplifies to
Use the well known relation (to be proved):
![]() |
(12.24) |
then Eq. (12.23) simplifies to
![]() |
(12.25) |
Integrating Eq. (12.25) over :
we obtain
![]() |
(12.26) |
Using the defintion of the surface average, Eq. (10.5), i.e.,
(where ,
is the volume enclosed by surface
),
Eq. (12.26) is written as
![]() |
(12.27) |
Noting the relation (proof is given in Sec. A.1):
![]() |
(12.28) |
Eq. (12.27) is written as
Noting that , the above
equation is simplified to
![]() |
(12.29) |
The toroidal magnetic flux is given by Eq. (11.8), i.e.,
Then the time partial derivative in coordinates
is written as
Use Eq. (12.29), then Eq. (12.30) is written as
where use has been made of the fact that at
.
If we choose as the toroidal magnetic flux,
i.e.,
(or more general
, where
is an arbitrary
time-independent monotonical function), then, by defintion
![]() |
(12.32) |
Combining this with Eqs. (12.31), we get
![]() |
(12.33) |
This is a condition that the coordinate velocity
(and thus
) satisfies when
you choose
as
.
Meanwhile, as is discussed above, the requirement of
being flux coordinates imposes a constraint given by Eq. (12.17).
Next, we combine this constraint with Eq. (12.33) to
uniquely determine
appearing in Eq. (12.17).
Mutiplying Eq. (12.33) by
and then
adding it to Eq. (12.19), we obtain
which simplifies to ( terms cancel out)
i.e.,
![]() |
(12.34) |
Then the poloidal magnetic flux diffusion equation is written as
![]() |
(12.35) |
Note that it is in the moving flux coordinates
that the poloidal magnetic flux diffusion can be written in the form
given by Eq. (12.35). Before I did the above derivation, I
considered Eq. (12.35) (which appears in many papers) as an
approximate equation in static lab frame (approximating the toroidal
component by the parallel component). It turns out that it is an exact
equation in the moving coordinate system.
We can switch the role of and
and do similar derivation as the above. I.e., we choose
and derive an evolution equation for the toroidal flux
. (It would be interesting to do this
derivation. to be done.) But most researchers have chosen
. Some reasons behind this choice
are as follows. Like the poloidal flux, the toroidal flux is contributed
by two parts: coils and plasma. Different from the poloidal flux, the
toroidal flux is mainly contributed by the coils. This is because the
plasma poloidal current (which contributes to the toroidal flux) is much
smaller than the TF coil currents. Because TF coil currents usually do
not change during one discharge, the toroidal flux
remains almost constant in a discharge (small variation is contributed
by the varying plasma poloidal current, which is much smaller than the
TF coil current). We usually prefer to our coordinate not moving
significantly with respect to the lab frame. So we prefer
over
.
A confusion I once had: if you choose ,
then
can also be considered as a function of
because
is a function of
. Then why does it matter
which one is chosen in the above derivation? The reason why we can not
easily switch between
and
is that thier relation is time-dependent, i.e.,
. This time-dependence will show up when we want to
switch from one to another.
Assume , then Eq. (12.35)
is written as
![]() |
(12.36) |
Using Eq. (10.12), the above equation is written as
![]() |
(12.37) |
Define by
![]() |
(12.38) |
where is typical value of the toroidal field (a
space-time constant). Since
is a
time-independent function of
,
so
can be used as the radial coordinate, i.e.,
. Then Eq. (12.37)
is written as
![]() |
(12.39) |
which agrees with Eq. (8) in Ref. [23]. The radial derivative of volume is written as
![]() |
(12.40) |
where use has been made of Eq. (11.10), i.e., .
![]() |
(12.41) |
We know that and
at the
magnetic axis. We also know that the safety factor
is nonzero at the magnetic axis. Then Eq. (12.41) implies
![]() |
(12.42) |
which serves as an inner boundary for the diffusion equation (12.39).
Using Eq. (11.18), the outer boundary condition can be written as
![]() |
(12.43) |
where
The fixed boundary equilibrium problem (also called the “inverse equilibrium problem” by some authors) refers to the case where the shape of a boundary magnetic surface is given and one is asked to solve the equilibrium within this magnetic surface. To make it convenient to deal with the shape of the boundary, one usually uses a general coordinates system which has one coordinate surface coinciding with the given magnetic surface. This makes it trivial to deal with the irregular boundary. To obtain the equilibrium, one needs to solve the GS equation in the general coordinate system.
Next we derive the form of the toroidal elliptic operator in a general curvilinear coordinate system. Specifically,
we consider a curvilinear coordinate system
, which is an arbitrary coordinate system except
that
is perpendicular to both
and
.
The toroidal elliptic operator takes the form given by Eq. (3.7), i.e.,
![]() |
(13.1) |
The gradient of is written as (note that
is independent of
)
![]() |
(13.2) |
Using this expression and the divergence formula (8.36), the elliptic operator in Eq. (13.1) is written
where the subscripts denotes partial derivatives,
is the Jacobian of the coordinate system
.
In a flux coordinate system, is independent of
, i.e.,
. Then the toroidal elliptic operator is
written as
![]() |
(13.4) |
Using Eq. (13.3), the GS equation (4.13) is written
which is the form of the GS equation in
coordinate system.
The toroidal elliptic operator in Eq. (13.3) can be written
![]() |
(13.6) |
where is defined by Eq. (8.77),
i.e.,
![]() |
(13.7) |
Next, we derive the finite difference form of the toroidal elliptic
operator. The finite difference form of the term
is written
![]() |
![]() |
![]() |
|
![]() |
![]() |
where
![]() |
(13.9) |
The finite difference form of is written
![]() |
![]() |
![]() |
|
![]() |
![]() |
where
![]() |
(13.11) |
The finite difference form of is written as
![]() |
(13.12) |
Approximating the value of at the grid centers
by the average of the value of
at the neighbor
grid points, Eq. (13.12) is written as
![]() |
(13.13) |
where
![]() |
(13.14) |
Similarly, the finite difference form of is
written as
![]() |
![]() |
![]() |
|
![]() |
![]() |
Using the above results, the finite difference form of the operator is written as
The coefficients are given by
![]() |
(13.16) |
![]() |
(13.17) |
and
![]() |
(13.18) |
where the Jacobian
![]() |
(13.19) |
The partial derivatives, ,
,
, and
,
appearing in Eqs. (13.16)-(13.19) are
calculated by using the central difference scheme. The values of
,
,
and
at the middle points
are approximated by the linear average of their values at the neighbor
grid points.
The function and
in the
GS equation are free functions which must be specified by users before
solving the GS equation. Next, we discuss one way to specify the free
functions. Following Ref. [15], we take
and
to be of the forms
![]() |
(13.20) |
![]() |
(13.21) |
with and
chosen to be of
polynomial form:
![]() |
(13.22) |
![]() |
(13.23) |
where
![]() |
(13.24) |
with the value of
on the
boundary,
the value of
on the magnetic axis,
,
,
,
,
, and
are free parameters.
Using the profiles of
and
given by Eqs. (13.20) and (13.21), we obtain
![]() |
(13.25) |
where , and
![]() |
(13.26) |
Then the term on the r.h.s (nonlinear source term) of the GS equation is written
![]() |
(13.27) |
The value of parameters ,
, and
in Eqs. (13.20) and (13.21), and the value of
and
in Eqs. (13.22)
and (13.23) can be chosen arbitrarily. The parameter
is used to set the value of the total toroidal
current. The toroidal current density is given by Eq. (A.79),
i.e.,
![]() |
(13.28) |
which can be integrated over the poloidal cross section within the boundary magnetic surface to give the total toroidal current,
Using
![]() |
(13.30) |
Eq. (13.29) is written as
![]() |
(13.31) |
from which we solve for ,
giving
![]() |
(13.32) |
If the total toroidal current is given, Eq. (13.32) can be used to determine the value of
.
In the fixed boundary equilibrium problem, the shape of the boundary magnetic surface (it is also the boundary of the computational region) is given while the shape of the inner flux surface is to be solved. A simple analytical expression for a D-shaped magnetic surface takes the form
![]() |
(13.33) |
![]() |
(13.34) |
with changing from
to
. According to the definition
in Eqs. (A.21), (A.22), and (A.24)
we can readily verify that the parameters
,
,
appearing in Eqs. (13.33) and (13.34) are
indeed the minor radius, major radius, and ellipticity, respectively.
According to the definition of triangularity Eq. (A.23),
the triangularity
for the magnetic surface
defined by Eqs. (13.33) and (13.34) is written
as
![]() |
(13.35) |
Another common expression for the shape of a magnetic surface was given by Miller[7, 21], which is written as
![]() |
(13.36) |
![]() |
(13.37) |
Note that Miller's formula is only slightly different from the formula
(13.33). For Miller's formula, it is easy to prove that the
triangularity is equal to
(instead of
as given in Eq. (13.35)).
In the iterative metric method[10] for solving the fixed boundary equilibrium problem, we need to provide an initial guess of the shape of the inner flux surface (this initial guess is used to construct a initial generalized coordinates system). A common guess of the inner flux surfaces is given by
![]() |
(13.38) |
![]() |
(13.39) |
where is a parameter,
is
a label parameter of flux surface. If the shape of the LCFS is given by
Eqs. (13.33) and (13.34), then Eqs. (13.38)
and (13.39) are written as
![]() |
(13.40) |
![]() |
(13.41) |
Fig. 13.1 plots the shape given by Eqs. (13.40)
and (13.41) for 0.4,
,
,
, and
with
varying from zero
to one.
![]() ![]() |
Figure 13.1. Shape of flux surface
given by Eqs. (13.40) and (13.41).
Left figure: points with the same value of |
The tokamak equilibrium problem where the shape of the LCFS is given is called fixed boundary equilibrium problem. I wrote a numerical code that uses the iterative metric method[10] to solve this kind of equilibrium problem. Figure 13.2 describes the steps involved in the iterative metric method.
To benchmark the numerical code, I set the profile of
and
according to Eqs. (A.37) and
(A.38) with the parameters
,
,
, and
.
The comparison of the analytic and numerical results are shown in Fig.
13.3.
![]() |
Figure 13.3. Agreement between the
magnetic surfaces (contours of |
Note that the parameter in the Solovev
equilibrium seems to be not needed in the numerical calculation. In fact
this impression is wrong: the
parameter is
actually needed in determining the boundary magnetic surface of the
numerical equilibrium (in the case considered here
is chosen as
).
With the pressure increasing, the magnetic axis usually shifts to the low-field-side of the device, as is shown in Fig. 13.4.
![]() ![]() |
In the GS equation, is one of the two free
functions which can be prescribed by users. In some cases, we want to
specify the safety factor profile
,
instead of
, in solving the
GS equation. Next, we derive the form of the GS equation that contains
, rather than
, as a free function. The safety factor defined
in Eq. (8.86) can be written
Equation (13.44) gives the relation between the safety
factor and the toroidal field function
. This relation can be used in the
GS equation to eliminate
in favor of
, which gives
![]() |
(13.45) |
![]() |
(13.46) |
Note that this expression involves the flux surface average, which
depends on the flux surface shape and the shape is unknown before is determined.
Multiplying Eq. (A.102) by gives
![]() |
(13.47) |
Surface-averaging the above equation, we obtain
![]() |
(13.48) |
![]() |
(13.49) |
![]() |
(13.50) |
![]() |
(13.51) |
![]() |
(13.52) |
Substitute Eq. (13.46) into the above equation to eliminate
, we obtain
![]() |
(13.53) |
Eq. (13.53) agrees with Eq. (5.55) in Ref. [15].
![]() |
(13.54) |
![]() |
(13.55) |
where
![]() |
(13.56) |
The GS equation is
![]() |
(13.57) |
![]() |
(13.58) |
![]() |
(13.59) |
Using Eq. (13.55) to eliminate in
the above equation, the coefficients before (
) is written as
![]() |
![]() |
![]() |
|
![]() |
![]() |
Substituting the expression of into the above
equation, we obtain
which is equal to the expression (5.58) in Ref. [15]. The
coefficient before is written as
Define
![]() |
(13.62) |
![]() |
(13.63) |
Using
![]() |
(13.64) |
Eq. () is written as
![]() |
(13.65) |
![]() |
(13.66) |
![]() |
(13.67) |
![]() |
(13.68) |
![]() |
(13.69) |
![]() |
(13.70) |
![]() |
(13.71) |
![]() |
(13.72) |
![]() |
(13.73) |
![]() |
(13.74) |
But the expression of is slightly different from
that given in Ref. [15] [Eq. (5.57)]. Using the above
coefficients, the GS equation with the
-profile
held fixed is written as
![]() |
(13.75) |
In Sec. 8.12, we introduced the local safety factor . Equation (8.82)
indicates that if the Jacobian is chosen to be of the particular form
, then the local safety
factor is independent of
,
i.e., magnetic line is straight in
plane. On the
other hand, if we want to make field line straight in
plane, the Jacobian must be chosen to be of the specific form
. We note that, as mentioned in
Sec. 8.14, the poloidal angle is fully determined by the
choice of the Jacobian. The specific choice of
is usually too restrictive for achieve a desired poloidal resolution
(for example, the equal-arc poloidal angle can not be achieved by this
choice of Jacobian). Is there any way that we can make the field line
straight in a coordinate system at the same time ensure that the
Jacobian can be freely adjusted to obtain desired poloidal angle? The
answer is obvious: Yes, we can achieve this by defining a new toroidal
angle
that generalizes the usual (cylindrical)
toroidal angle
. Define a new
toroidal angle
by[9]
![]() |
(14.1) |
where is a unknown function to be determined by
the constraint of field line being straight in
plane. Using Eq. (8.83), the new local safety factor in
coordinates is written as
To make the new local safety factor be independent of , the right-hand side of Eq. (14.2)
should be independent of
,
i.e.,
![]() |
(14.3) |
where can be an arbitrary function of
. A convenient choice for
is
, i.e.,
making the new local safety factor be equal to the original global
safety factor, i.e.,
. In
this case, equation (14.3) is written as
![]() |
(14.4) |
which, on a magnetic surface labed by ,
can be integrated over
to give
![]() |
(14.5) |
where is an starting poloidal angle arbitrarily
chosen for the integration, and
is the constant
of integration. In the following, both
and
will be chosen to be zero. Then the above equations
is written
![]() |
(14.6) |
Substituting the above expression into the definition of (Eq. 14.1), we obtain
![]() |
(14.7) |
which is the formula for calculating the general toroidal angle. If is a straight-field line poloidal angle, then
in Eq. (14.7) reduces to the usual
toroidal angle
.
In summary, magnetic field line is straight in
plane with slope being
if
is defined by Eq. (14.7). In this method, we make the field
line straight by defining a new toroidal angle, instead of requiring the
Jacobian to take particular forms. Thus, the freedom of choosing the
form of the Jacobian is still available to be used later to choose a
good poloidal angle coordinate. Note that the Jacobian of the new
coordinates
is equal to that of
. [Proof:
] Also note that defined by Eq. (14.6)
is a periodic function of
.
[Proof: Equation (14.6) implies that
![]() |
![]() |
![]() |
|
![]() |
![]() |
||
![]() |
![]() |
||
![]() |
![]() |
]
[In numerical implementation, the term appearing
in
is computed by using
![]() |
![]() |
![]() |
|
![]() |
![]() |
||
![]() |
![]() |
For later use, from Eq. (14.6), we obtain
This formula is used in GTAW code, where the derivative
is calculated numerically by using the central difference scheme.]
Recall that the contravariant form of the magnetic field in coordinates is given by Eq. (8.83), i.e.,
![]() |
(14.11) |
Next, let us derive the corresponding form in
coordinates. Using the definition of
,
equation (14.11) is written as
![]() |
![]() |
![]() |
|
![]() |
![]() |
||
![]() |
![]() |
Using Eq. (14.4), the above equation is simplified as
Equation (14.13) is the contravariant form of the magnetic
field in coordinates.
The expression of the magnetic field in Eq. (14.13) can be
rewritten in terms of the flux function and
discussed in Sec. (). Equation (14.13)
is
![]() |
(14.14) |
which, by using Eq. (), i.e., ,
is rewritten as
![]() |
(14.15) |
which, by using Eq. (), i.e., ,
is further written as
![]() |
(14.16) |
Noting the simple fact that
![]() |
(14.17) |
where is a constant, we conclude that
![]() |
(14.18) |
(since , where the part
acts as a constant when we hold
and
constant), i.e., the symmetry property with
respect to the new toroidal angle
is identical
with the one with respect to the old toroidal angle
. On the other hand, generally,
![]() |
(14.19) |
and
![]() |
(14.20) |
In the special case that is axisymmetric (i.e.,
is independent of
in
coordinates), then two sides of Eqs. (14.19)
and (14.20) are equal to each other. Note that the partial
derivatives
and
in Sec.
14.1 and 14.2 are taken in
coordinates. Because the quantities involved in Sec. 14.1
and 14.2 are axisymmetric, these partial derivatives are
equal to their counterparts in
coordinates.
In Sec. 9, we have provided the steps to construct the
magnetic surface coordinate system .
Only one additional step is needed to construct the straight-line flux
coordinate system
. The
additional step is to calculate the generalized toroidal angle
according to Eq. (14.1), where
is obtained from Eq. (14.6). Also note that
the Jacobian of
coordinates happens to be equal
to that of
coordinates.
The usefulness of the contravariant form [Eq. (14.13] of
the magnetic field lies in that it allows a simple form of operator in a coordinate system. (The operator
is usually called magnetic differential operator.) In
coordinate system, by using the contravariant form
Eq. (14.13), the operator is written as
Next, consider the solution of the following magnetic differential equation:
![]() |
(14.22) |
where is some known function. Using Eq. (14.21), the magnetic differential equation is written as
![]() |
(14.23) |
Note that the coefficients before the two partial derivatives of the
above equation are all independent of and
. This indicates that different
Fourier harmonics in
and
are decoupled. As a result of this fact, if
is
Fourier expanded as
![]() |
(14.24) |
(note that, following the convention adopted in tokamak literature[9], the Fourier harmonics are chosen to be , instead of
),
and the right-hand side is expanded as
![]() |
(14.25) |
then Eq. (14.23) can be readily solved to give
![]() |
(14.26) |
The usefulness of the straight line magnetic coordinates lies in that, as mentioned previously, it makes the
coefficients before the two partial derivatives both independent of
and
,
thus, allowing a simple solution to the magnetic differential equation.
Equation (14.26) indicates that, for the differential
equation (14.22), there is a resonant response to a
perturbation on a magnetic surface with
. Therefore, the magnetic surface
with
is called the “resonant
surface” for the perturbation
.
The phase change of the perturbation along a
magnetic field is given by
,
which can be written as
.
Since
on a resonant surface, this indicates that
there is no phase change along a magnetic field line on a resonant
surface, i.e., the parallel wavenumber
is zero
on a resonant surface.
Next, we discuss a special poloidal angle, which is useful in
describling a perturnbation of single harmonic . This poloidal angle is defined by
![]() |
(14.27) |
where are the mode numbers of the perturbation.
The poloidal angle
is often called helical angle
and is special in that its definition is associated with a perturbation
(the mode numbers of the perturbation appear in the definition) while
the definition of the poloidal angles discussed previously only involve
the equilibrium quantities.
The poloidal angle is designed to make 3D
perturbations of the form
reduce to 2D
perturbations, i.e.,
![]() |
(14.28) |
It is ready to verify that the Jacobian of coordinates
is equal to that of coordinates
[proof:
].
The component of along
direction (i.e., the covariant component) is written
At the resonant surface ,
equation (14.29) implies
.
The direction
defines the reconnecting component
of the magnetic field?
On the other hand, the component of along
direction is written
Using (14.30) and (14.29), the relation
between and
is written
as
![]() |
(14.31) |
In the above, we have obtained the covariant form of the magnetic field
in coordinates (i.e., Eq. (8.80)).
Next, we derive the corresponding form in
coordinate. In order to do this, we need to express the
basis vector in terms of
,
, and
basis vectors. Using the definition of the generalized toroidal angle,
we obtain
Using Eq. (14.32), the covariant form of the magnetic field, Eq. (8.80), is written as
![]() |
(14.33) |
This expression can be further simplified by using equation (14.4)
to eliminate , which gives
![]() |
![]() |
![]() |
|
![]() |
![]() |
Using , the above equation is
written as
Equation (14.35) is the covariant form of the magnetic
field in coordinate system. For the particular
choice of the radial coordinate
and the Jacobian
(i.e., Boozer's Jacobian, discussed in Sec. 14.9), equation (14.35) reduces to
![]() |
(14.36) |
with . The magnetic field
expression in Eq. (14.36) frequently appears in tokamak
literature[30]. In this form, the coefficients before both
and
depends on only the
radial coordinate. In terms of
,
the Jabobian can also be written as
![]() |
(14.37) |
In solving the MHD eigenmode equations in toroidal geometries, besides
the operator, we will also encounter another
surface operator
. Next, we
derive the form of the this operator in
coordinate system. Using the covariant form of the equilibrium magnetic
field [Eq. (14.35)], we obtain
![]() |
(14.38) |
Using this, the operator is written as
which is the form of the operator in coordinate
system.
Examining Eq. (14.40), we find that the coefficients before
the two partial derivatives will be independent of
and
if the Jacobian
is
chosen to be of the form
,
where
is some magnetic surface function. It is
obvious that the independence of the coefficients on
and
will be advantageous to some applications.
The coordinate system
with the particular choice
of
is called the Boozer coordinates. The
usefulness of the new toroidal angle
is
highlighted in Boozer's choice of the Jacobian, which makes both
and
be a constant-coefficient
differential operator. For other choices of the Jacobian, only the
operator is a constant-coefficient differential
operator.
In solving the MHD eigenmode equations in toroidal geometry, we also
need the radial differential operator .
Next, we derive the form of the operator in
coordinates. Using
the radial differential operator is written as
where and
are given
respectively by Eqs. (14.10) and (14.4). Using
the above formula,
is written as
![]() |
(14.42) |
This formula is used in GTAW code.
In coordinates, a magnetic field line is
straight in
plane with the slope being
. So the equation for magnetic
field lines is given by
,
where
is a constant in
plane. Note that
can be used to label different
magnetic field lines on a magnetic surface. This motivates us to use
, i.e.,
![]() |
(15.1) |
to as a new coordinate to replace .
Then we obtain a new set of coordinate:
,
which are called field-line-following coordinates. Using Eqs. (15.1)
and (14.7),
can be written as
where is the local safety factor. (If we choose
the straight-field-line
,
then
is written as
.)
Define
, which is called
tor_shift in TEK code, then
.
In TEK, I choose with
corresponding to the high-field-side midplane, and
is increasing along the counter-clockwise direction wehn viewed along
. There are two reasons why
is chosen this way: 1. The toroidal shift
remain small near the low-field-side
. This may be good for spatial resolution in
this important region where ballooning modes often have larger
amplitude. 2. The
cut, i.e.,
, is far away from the important
low-field-side. The gyrokinetic field equations are not solved at
. Perturbations there need to
obtained by using interpolations at the
cut,
i.e., infer values of perturbations on
gridpoints at
from the corresponding values at
. Numerical errors more
likely appear there. So we prefer that the
cut
is located in less important area (area where mode amplitude is small).
In coordinates, the magnetic field in Eq. (14.13) is written as
![]() |
(15.3) |
which is called the Clebsch form. The direction
![]() |
(15.4) |
is parallel (or anti-parallel) to the magnetic field direction. Due to
this fact, coordinates are usually called
“field-line-following coordinates” or “field-aligned
coordinates” [2, 5].
Equation (15.3) implies that
![]() |
(15.5) |
and
![]() |
(15.6) |
i.e., both and
are
constant along a magnetic field line. Taking scalar product of Eq. (15.3) with
, we
obtain
![]() |
(15.7) |
which is nonzero, i.e., only among
is changing along a magnetic field line. (Here
is the Jacobian of the coordinate system
, which happens to be equal to the Jacobian of
coordinates.)
Using Eq. (15.3), the magnetic differential operator in the new coordinate system
is written
![]() |
(15.8) |
which is just a partial derivative over ,
as is expected, since only
is changing along a
magnetic field line.
By the way, note that , i.e.,
the magnetic field lines are straight with zero slope on
plane.
The fact implies that
is
constant along a magnetic field line. At first glance, a magnetic line
on an irrational surface seems to sample all the points on the surface.
This seems to indicate that
is a constant on
irrational surfaces. However,
must be varying so
that it can provide a suitable toroidal coordinate. I had once been
confused by this conflict for a long time. The key point to resolve this
confusion is to realize that it is wrong to say there is only one
magnetic line on an irrational surface, i.e. it is wrong to say a
magnetic line on an irrational surface samples all the points on the
surface. There are still infinite number of magnetic field lines that
can not be connected with each other on an irrational surface. Then the
fact
does not imply that
must be the same on these different magnetic field lines. In fact,
although
, the gradient of
on a flux-surface along the perpendicular
direction is nonzero, i.e.,
.
[Proof:
which is obviously nonzero.] This indicates that
is not constant on a flux-surface. Equation (15.10) also
indicates that the perpendicular gradient becomes infinity at points
where
(i.e., X-points).
In coordinates,
is
perpendicular to
. However,
in field-line-following coordinates
,
is not perpendicular to
. Therefore
is not along the
binormal direction
.
It is widely believed that turbulence responsible for energy transport
in tokamak plasmas usually has ,
where
and
are the
parallel and perpendicular wavenumbers, respectively. Due to this
elongated structure along the parallel direction, less grids can be used
in the parallel direction than that in the perpendicular direction in
turbulence simulation. In this case, the field-aligned coordinates
provide suitable coordinates to be used, where less
gridpoints can be used for
coordinate in
simulations and even some
derivatives can be
neglected (high-n approximation), which simplifies the equations that
need to be solved. There is another reason why almost all gyrokinetic
codes use field-aligned coordinates: the stability of numerical
algorthims is improved when we use coarse grids in the parallel
direction because the parallel Courant condition (for explicit schemes)
can be more easily satisfied (especially for the
cases with kinetic electrons), where
is the
parallel grid spacing, which is larger when coarse grids are used in the
parallel direction. This is also mentioned in Ref. [22] and
it seems right from my experiences of testing several algorithms. This
can also be understood in the following way: the coarse parallel grid
automatically filters out physically irrelevant but numerically
problematic high-
modes, permitting much longer
time steps for explicit time stepping, in both particle and fluid
codes[11].
The generalized toroidal angle is defined by Eq.
(15.2), i.e.,
,
where
. The gradient of
is then written as
Then
![]() |
(15.12) |
![]() |
(15.13) |
Another method of computing the above quantities: Using Eqs. (9.10) and (9.11), expression (15.11) is written as
![]() |
![]() |
![]() |
|
![]() |
![]() |
(Note that is discontinuous across the
cut.) Then
![]() |
![]() |
![]() |
|
![]() |
![]() |
![]() |
![]() |
![]() |
|
![]() |
![]() |
The figures below plot the spatial variation of some quantities.
|
|
In GEM[6] and GENE[12] codes, the
field-aligned coordinates are defined by
![]() |
(15.17) |
![]() |
(15.18) |
![]() |
(15.19) |
where is an arbitrary flux surface label with
length dimension, which is often chosen in
GEM
to be the
minor radius of a magnetic surface in the midplane. Here and
are constant quantities of
length dimension,
is the minor radius of a
reference magnetic surface (usually corresponding to the center of the
radial simulation box),
is the major radius of
the magnetic axis,
is the safety factor value on
the
surface. The constant length
introduced in the definition of
is
to make
approximately correspond to the length
along the field line in the large-aspect ratio limit. The constant
length
introduced in the definition of
is to make
corresponds to the
arc-length in the poloidal plane traced by a field line when its usual
toroidal angle increment
is
. This explanation makes
look like a poloidal coordinate whereas
is
actually a toroidal coordinate.
Next, let us calculate the wave number along the
direction,
, for a mode with
toroidal wave number
. The
wavelength along the
direction,
, is given by
![]() |
(15.20) |
Then the wavenumber is written as
![]() |
(15.21) |
On the other hand, the poloidal wavenumber is
given by
![]() |
(15.22) |
where is the poloidal mode number of the mode in
coordinates. If the mode has the property
, i.e.,
, then
is equal to the
. The motivation of introducing the
constant length
in the definition of
is to make
for a mode with
.
Some authors call or
by
the name “binormal wavenumber”, which is not an appropriate
name in my opinion. Some authors call
the
binormal direction, which is also an inappropriate name since neither
nor
is along the
binormal direction
.
In this section, I try to visualize gridpoints in the field aligned
coordinates. The directions of the covariant basis vectors of coordinates are as follows:
![]() |
(15.23) |
![]() |
(15.24) |
![]() |
(15.25) |
Here is a combination of the usual radial and
toroidal direction, which needs some clarification. Note that,
is related to
by Eq. (15.2),
i.e.,
![]() |
(15.26) |
(where the second equality becomes exact if is
the straight-field-line poloidal angle defined in Sec. 8.15.5.),
which indicates that, for
and
, the usual toroidal angle
is changing when changing
and holding
and
fixed. Figure 15.4b
shows how the usual toroidal angle
changes when
we change
and hold
and
fixed.
The relation given by Eq. (15.26)
indicates that the toroidal shift along
for a
radial change form
to
is
given by
, which is larger on
isosurface with larger value of
. An example for this is shown in Fig. 15.5
for
, where
has larger toroidal shift than that in Fig. 15.4 for
.
|
The curves can be understood from another
perspective. Examine a family of magnetic field lines that start from
and
but different radial
coordinates. These starting points all have the same value of
, which is equal to
. When following these field lines to another
isosurfce of
, the
intersecting points of these field lines with the
isosurface will trace out a
line with
. Examine another family of
magnetic field lines similar to the above but with the starting toroidal
angle
. They will trace out
another
line (with
)
on the
isosurface. Continue the process, we
finally get those curves in Fig. 15.4b and Fig. 15.5b.
For a harmonic in coordinates given by
, the radial wave number is
. Let us calculate the
corresponding radial wavenumber
in the new
coordinates
, which is
defined by
![]() |
(15.27) |
where the phase is given by .
Then the above expression is written as
![]() |
![]() |
![]() |
|
![]() |
![]() |
Using Using , the above
expression is written as
![]() |
where . This result indicates
that, compared with the radial wavenumber in coordinate system
, the radial wavenumber in the new
coordinate system
has an increment
. For nonzero magnetic shear (
) and nonzero poloidal location (
), the increment
can be
large for modes with toroidal mode number
.
Then we need to use more radial grid number (compared what is needed in
coordinates) to resolve the radial variation.
This is one of disadvantage of using field-aligned coordinates. If the
saving associated with using less parallel grid number out-weights the
cost associated with using more radial grid number, we obtain a net
saving in using the field-aligned coordinates.
Let us examine how many grid points are needed
to resolve the
dependence in
coordinates on the high-field side (
).
Assume
, then
at
is given by
. The corresponding wave-length is given by
. The grid spacing
should be less than half of this wave-length (sampling
theorem). Then the grid number should satisfy that
![]() |
(15.30) |
where is the radial width of the computational
domain.
The number of Fourier harmonics that need to be included is given by
![]() |
(15.31) |
For DIII-D cyclone base case, choose the radial coordinate as
. At the
radial location
,
,
,
then
. Then
for the radial width
.
Figure 15.6 plots lines on the
isosurfaces, which are chosen to be on the
low-field-side midplane. On
surface,
lines are identical to
lines. On
surface, each
line has
large
shift. In old version of my code,
surfaces are chosen as the
cuts (in the new version
). A
connection condition for the perturbations is needed between these two
surfaces. This connection condition is discussed in Sec. 15.6.2.
Since and
correspond to
the same spatial point, a real space continuous quantity
expressed in terms of coordinates
, i.e.,
,
must satisfy the following periodic conditions along
:
![]() |
(15.32) |
Since and
correspond to
the same spatial point,
must satisfy the
following periodic conditions along
:
![]() |
(15.33) |
Since and
correspond to
the same spatial point, a real space continuous quantity
expressed in terms of field-line-following coordinates
, i.e.,
, must satisfy the following periodic condition
along
:
![]() |
(15.34) |
However, generally there is no periodic condition along ,
![]() |
(15.35) |
because and
are
generally not the same spatial point. In fact, equation (15.2)
implies, for point
, its
toroidal angle
is given by
![]() |
(15.36) |
while for point , its
toroidal angle
is given by
![]() |
(15.37) |
i.e., and
are different
by
. From this, we know that
and
correspond to the
same spatial point. Therefore we have the following periodic condition:
![]() |
(15.38) |
or equivalently
![]() |
(15.39) |
For the fully kinetic ion module of GEM code that I am developing, is chosen in the range
.
The condition (15.39) imposes the following boundary
condition:
![]() |
(15.40) |
If is on a grid,
is
usually not on a grid. Therefore, to get the value of
, an interpolation of the discrete date over
the generalized toroidal angle
(or equivalently
) is needed, as is shown in
Fig. 15.7.
Figure 15.8 compares a small number of
contours and
contours on a magnetic surface.
As is shown in the left panel of Fig. 15.8, with fixed, an
curve reaches its
starting point when
changes from zero to
. However, as shown in the right
panel of Fig. 15.8, with
fixed, an
curve (i.e. a magnetic field line) does not
necessarily reach its starting point when
changes from zero to
. There
is a toroidal shift,
,
between the starting point and ending point. Therefore there is
generally no periodic condition along
since
is not always an integer. A mixed periodic condition
involves both
and
is
given in (15.38).
In field-line-following coordinates ,
a toroidal harmonic of a physical perturbation can be written as
![]() |
(15.41) |
where is the toroidal mode number,
, which is not necessary an integer, is
introduced to describe the variation along a field line. The periodic
condition given by Eq. (15.38) requires that
![]() |
(15.42) |
To satisfy the above condition, we can choose
![]() |
(15.43) |
where is an arbitrary integer, i.e.,
![]() |
(15.44) |
We are interested in perturbation with a slow variation along the field
line direction (i.e., along )
and thus we want the value of
to be small. One
of the possible small values given by expression (15.44) is
to choose
, so that
is given by
![]() |
(15.45) |
where N is a function that return the nearest
integer of its argument,
and
is the maximal and minimal value of the safety factor in the radial
region in which we are interested. [In the past, I choose
. However,
in this case
is not a continuous function of
and thus is not
physical.] This form is used to set the initial density perturbation in
the fully kinetic code I am developing. Note that
in Eq. (15.45) depends on the radial coordinate
through
. Also
note that
here is different from the poloidal
mode number
in
coordinate system. It is ready to show that the perturbation given by
Eq. (15.41) with
and
has large poloidal mode number
when expressed in
coordinates. [Proof:
Expression (15.41) can be written as
![]() |
(15.46) |
Assume that is the straight-field-line poloidal
angle in
coordinate system, then
and the above equation is written as
![]() |
(15.47) |
which indicates the poloidal mode number in
coordinates is given by
.
For the case with
and
,
is much larger than one.]
Since contours on a magnetic surface are
magnetic field lines, they span out the 3D shape of magnetic surface
when there are many
contours on a magnetic
surface, as is shown by the right-panel of Fig. 15.9.
Figure 15.10 compares the
coordinate surface of
coordinates with the
coordinate surface of
coordinates.
![]() ![]() |
Figure 15.11. The same plot as in Fig. 15.10 but with a larger radial range. |
|
|
The above grid is often called “flux-tube”, since it is created by following field lines, and it looks like a tube if the field lines are very near to each other (e.g., when we choose a very narrow radial and toroidal region to start from). Since no magnetic field line passes through the sides of the tube, the flux through any cross section of the tube is equal. The term “flux tube” is often used in astrophysics.
Most gyrokinetic simulation codes use field-line-following coordinates in constructing spatial mesh. The mesh is conceptually generated by the following three steps: (1) selecting some initial points; (2) tracing out magnetic-field-lines passing through these points; (3) choose the intersecting points of these field-lines with a series of chosen surfaces as the final grid-points. The initial points and the chosen surfaces differ among various codes and thus the resulting grid differs. Next, let us discuss some examples.
Given the definition of coordinates, first
choose
points at
with
. (The
plane is chosen to be near the low-field-side midplane.) Then trace out
the magnetic-field-lines passing through these points for one poloidal
circuit (poloidal angle range:
)
and record the intersection points of these field-lines with various
planes. This gives a 2D grid [
surface for a fixed value of
].
After this, we rotate the 2D grid toroidally to get the 3D grid.
Ususally we consider a toroidal wedge rather than a full torus (to
reduce computational cost). So the 2D grid is rotated toroidally by
, where
.
Note that the gridpoints on surface usually do
not coincide those on
surface due to the
toroidal shift along the field line arising when the safety factor is
irrational. Interpolation can be used to map physical variables defined
on gridpoints in
plane to those in
plane.
Given the definition of coordinates, 2D
grid-points can be chosen on
plane based on
coordinates. Then trace out the magnetic-field-lines
starting from these points for one toroidal circuit and record the
intersection points of these field-lines with
planes, where
,
. It is obvious that the resulting mesh are not
toroidally symmetrical. And also the grids on
plane differ from those on
plane. Interpolation
can be used to mapping physical variables defined on grid-points of
to those of
plane. In this
case, the number of “toroidal grid-points”
(i.e., the number of poloidal planes) is actually the number of
grid-points in the parallel direction within one toroidal circuit.
XGC1
can handle the region outside of the LCFS. Here we
only discuss the region inside the LCFS. At each radial gridpoint on
plane, follow the magnetic field line starting
form this point for one poloidal loop and record the intersection points
of this field-line with
planes, where
,
.
For the case
, one
magnetic-field-line will have more than one intersection points on some
poloidal planes. Repeat tracing the field line for each radial location.
Then project (toroidally) all the intersection points on different
poloidal planes to a single poloidal plane and rotate (toroidally) these
2D grids to define a 3D grid that are toroidally symmetrical.
The generalized toroidal angle is numerically
calculated in my code. To verify
along a
magnetic field-line, figure 15.14 plots the values of
along a magnetic field line, which indicates that
is constant. This indicates the numerical
implementation of the field-aligned coordinates is correct.
![]() ![]() |
Let us introduce the binormal wavenumber, which is frequently used in
presenting turbulence simulation results. Define the binormal direction
by
which is a unit vector lying on a magnetic surface and perpendicular to
. The binormal wavenumber of
a mode is defined by
![]() |
(15.48) |
where is the phase of the mode. Consider a mode
given by
, then the phase
. Then
is written as
![]() |
![]() |
![]() |
|
![]() |
![]() |
where the radial phase does not appear since
. The above expression can be
further written as
Equation (15.50) is the general expression of the binormal
wavenumber. On the resonant surface of the mode, i.e., , then the above expression is written as
![]() |
(15.51) |
Using Eq. (14.12), i.e., ,
the above expression is written as
Using , the above equation is
written
![]() |
(15.52) |
which indicates the binormal wavenumber generally depends on the
poloidal angle. For large aspect-ratio tokamak, we have ,
.
Then Eq. (15.52) is written
![]() |
(15.53) |
which indicates the binormal wavenumber are approximately independent of
the poloidal angle. Since on a resonant surface,
the above equation is written
,
which is the usual poloidal wave number. Due to this relation, the
binormal wavenumber
is often called the poloidal
wavenumber and denoted by
in papers on tokamak
turbulence. In the GENE code,
coordinate is
defined by
. Then the
of a mode of toroidal mode number
is given by
where
and
. Then
is written as
, which is
similar to the binormal defined above. For this reason,
of GENE code is also called binormal wave-vector, which is in fact not
reasonable because neither
or
is along the binormal direction.
Assume magnetic surfaces of a magnetic configuration are known and given by
![]() |
(16.1) |
![]() |
(16.2) |
where are two parameters and
is magnetic surface label (i.e.,
).
The above parametric equations specify a series of concentric-circular
magnetic surfaces.
Assume the toroidal field function is given.
Then the toroidal magnetic field is determined by
. Further assume the safety factor profile
is given, then the magnetic field is fully determined.
Next, let us derive the explicit form of the poloidal magnetic field
, which is given by
![]() |
(16.3) |
which involves the poloidal magnetic flux .
Therefore our task is to express
in terms of
and
.
Using
, we obtain
Integrate the above equation over ,
we obtain
![]() |
(16.4) |
which an be written as
![]() |
(16.5) |
where use has been made of .
Using
and
,
the above equation is written
![]() |
(16.6) |
Using maxima
(an open-source computer algebra system), the
above integration over can be performed
analytically, giving
![]() |
(16.7) |
Using this, equation (16.6) is written as
![]() |
(16.8) |
which can be simplified as
![]() |
(16.9) |
This is what we want—the expression of the poloidal magnetic flux
in terms of and
.
[Another way of obtaining Eq. (16.9) is to use Eq. (8.86), i.e.,
![]() |
(16.10) |
where is the Jacobian of the
coordinates and is given by
.
Then Eq. (16.10) is simplified as
![]() |
(16.11) |
which, after being integrated over ,
gives Eq. (16.9).]
Using Eq. (16.9), the poloidal magnetic field in Eq. (16.3) is written as
[Using the formulas and
, where
is the Jacobian of
the
coordinates, we obtain
and
,
. Then Eq. (16.12) is written as
![]() |
(16.13) |
This is the explicit form of the poloidal magnetic field in terms of
and
.
The magnitude of
is written as
![]() |
(16.14) |
Note that both and
depend on the poloidal angle
.]
I use Eq. (16.9) to compute the 2D data of
(
) on the poloidal plane when
creating a numerical G-eqdsk file for the above magnetic configuration
(Fortran code is at
/home/yj/project_new/circular_configuration_with_q_given
).
Assume that the poloidal plasma current is zero, then
is a constant independent of
.
This is always assumed by the authors who use concentric-circular
configuration but is seldom explicitly mentioned.
In analytical work, dependence on
is often approximated as
where is the local inverse aspect ratio.
Since the above field is derived from the general form given by Eq. (1.10), it is guaranteed that the field is divergence-free. In
case of any doubt, let us directly verify this. Write
as
![]() |
(16.15) |
where is the Jacobian of
coordinates;
,
, and
are given by
![]() |
(16.16) |
![]() |
(16.17) |
![]() |
(16.18) |
Use given by (16.12), then
,
,
and
are written as
![]() |
(16.19) |
![]() |
(16.20) |
and
![]() |
(16.21) |
respectively. Then, by using the divergence formula in
coordinates,
is written as
![]() |
![]() |
![]() |
|
![]() |
![]() |
||
![]() |
![]() |
||
![]() |
![]() |
i.e., in this case is indeed divergence-free.
The answer is no. It is ready to verify that the poloidal magnetic flux
function given by Eq. (16.9) is not
a solution to the GS equation (A.119) even if the plasma
pressure in the GS equation is set to zero. The above expression is a
solution to the GS equation in the limit of infinite aspect ratio and
zero plasma pressure. The finite aspect ratio and plasma pressure
requires the magnetic surfaces to have the Shafranov shift in order to
satisfy the GS equation (see Sec. A.27).
For the above magnetic field, the toroidal shift involved in the definition of the generalized toroidal angle can be expressed in simple analytical form. The toroidal shift is given by
![]() |
(16.23) |
where the local safety factor can be written as
![]() |
(16.24) |
Using and
![]() |
(16.25) |
The local safety factor in Eq. (16.24)
is written as
![]() |
(16.26) |
Using this, expression (16.23) is written
![]() |
(16.27) |
Assume , then the integration
can be analytically performed (using maxima),
yielding
![]() |
(16.28) |
Then expression (16.27) is written
![]() |
(16.29) |
where use has been made of .
Using this, the generalized toroidal angle can be written as
The results given by the formula (16.29) are compared with the results from my code that assumes a general numerical configuration. The results from the two methods agree with each other, as is shown in Fig. 16.1, which provides confidence in both the analytical formula and the numerical code.
![]() |
Figure 16.1. The results of |
In passing, we note that the straight-field-line poloidal angle can also be considered to be defined by
![]() |
(16.31) |
i.,e,
![]() |
(16.32) |
Then using Eq. (16.30), is written
as
![]() |
(16.33) |
which agrees with Eq. (A2) in Gorler's paper[12].
Let and define
,
, etc. Explicit expressions
for these elements can be written as
![]() |
![]() |
![]() |
|
![]() |
![]() |
||
![]() |
![]() |
||
![]() |
![]() |
||
![]() |
![]() |
![]() |
(16.35) |
Using expression (16.29), can be
evaluated analytically, yielding
where use has been made of
(I did not remember this formula and I use SymPy to obtain this.) These expressions are used to benchmark the numerical code that assume general flux surface shapes. The results show that the code gives correct result when concentric circular flux surfaces are used.
Taking the derivative of
, equation (16.29) is written as (using
Sympy)
![]() |
(16.36) |
where
![]() |
(16.37) |
Equation (16.26) should be equal to
given by Eq. (16.26). This was verified numerically.
Taking the derivative of Eq. (16.9),
we obtain
![]() |
(16.38) |
i.e.,
![]() |
(16.39) |
![]() |
(16.40) |
|
The magnetic shear for a concentric-circular configuration is defined by
![]() |
(16.41) |
where is the minor radius of a magnetic surface.
The above expression can be re-arranged as
![]() |
(16.42) |
Integrating the above equation over and assuming
is a constant, we obtain
![]() |
(16.43) |
Performing the integration, the above equation is written as
![]() |
(16.44) |
where . Equation (16.44)
can be finally written as
![]() |
(16.45) |
This is a profile with a constant magnetic shear . In Ben's toroidal ITG simulation, the following
profile is used:
![]() |
(16.46) |
with . This is a linear
profile over
, with the
values of
and the shear at
being
and
,
respectively.
Prove that
![]() |
(A.1) |
where is an arbitary axisymmetric vector.
Proof: Using the divergence formula in
coordinates, i.e.,
where ,
,
.
Then
is written as
where use has been made of that is a periodic
function of
with period
. Then
is further written as
![]() |
![]() |
![]() |
|
![]() |
![]() |
||
![]() |
![]() |
||
![]() |
![]() |
which proves relation (A.1).
The GS equation is given by Eq. (4.13), i.e.,
![]() |
(A.3) |
If a solution to the GS equation is obtained, the solution can be scaled
to obtain a family of solutions. Given an equilibrium with ,
,
, then it is ready to prove
that
,
, and
is also a solution to
the GS equation, where
is a constant. In this
case, both the poloidal and toroidal magnetic fields are increased by a
factor of
, and thus the
safety factor remains unchanged. Also note that the pressure is
increased by
factor and thus the value of
(the ratio of the therm pressure to magnetic
pressure) remains unchanged. Note that
,
which indicates that the direction of the toroidal magnetic field can be
reversed without breaking the force balance. Also note that
and
can be negative, which
indicates that the direction of the toroidal current can also be
reversed without breaking the force balance.
The second kind of scaling is to set ,
, and
. It is ready to prove that the scaled expression is
still a solution to the GS equation because
. This scaling keep the pressure and the poloidal
field unchanged and thus the poloidal beta
remains unchanged. This scaling scales the toroidal field and thus can
be used to generate a family of equilibria with different profiles of
safety factor.
Another scaling, which is trivial, is to set ,
, and
. This scaling can be used to
test the effects of the pressure (not the pressure gradient) on various
physical processes.
When a numerical equilibrium is obtained, one can use these scaling methods together to generate new equilibria that satisfy particular global conditions. Note that the shape of magnetic surfaces of the scaled equilibrium remains the same as the original one.
The above scaling is made under the constraint that the the GS equation
(A.3), is satisfied. In practice, we may scale by a factor while fixing
and
. This does not satisfy the GS
equation, but allows more flexibility in changing the the safety factor
profile. Scaling
by a factor corresponds to
scaling the plasma current and the poloidal magnetic field.
For a large aspect-ratio, circular cross section tokamak, the on a magnetic surface is nearly constant,
. The poloidal angle dependence of the magnetic
field can be neglected, i.e.,
,
and
, where
is the minor radius of the relavent magnetic surface. Using these, the
safety factor in Eq. (1.34) is approximated to
![]() |
(A.4) |
EAST has 14 superconducting poloidal field (PF) coils and 2 in-vessel copper coils. The layout is shown in Fig. A.1.
![]() |
Figure A.1. Coils system of EAST tokamak. Coils PF1 to PF6 have 140 turns/coil. Both PF11 and PF12 have 64 turns, Both PF13 and PF14 have 32 turns. PF7 and PF9 are adjacent to each other and are connected in series, and share one power supply. Therefore they are considered as one independent coil, with total turns being 248. The same situation applies for PF8 and PF10. Locations of PF coils are from Refs. [26, 8]. Maximum current per turn in PF coils is 14.5kA. The inner green double D-shaped structures correspond to the vaccum vessel wall, and the outer red dashed D-shape corresponds to the TF-coils. Note that PF1-14 coils are all located outside of the TF-coils. Red coils are RMP copper coils, which generate 3D magnetic perturbations. |
All the PF coils can be considered as shaping coils since they all have effects in shaping the plasmas. In practice, they are further classified according to their main roles. PF1 to PF6 form a solenoid in the center of the torus and thus called Central Solenoide (CS) coils. Their main role is to induce electric field to drive current in the plasma and heat the plasma. As a result, they are often called “Ohm heating coils”.
PF13 and PF14 are mainly used to control (slow) vertical plasma displacement and thus are often called “position control coils”. PF11 and PF12 are used to triangularize the plasma and thus is called “shaping coils”. In order to squeeze the plasma to form desied triangularity, their currents need to be in the opposite direction of plamsa current (since two opposite currents repel each other). PF7+PF9 and PF8+PF10 are often called (by EAST operators) as “big coils” or “divertor coils” since they have the largest number of turns and current and are used to elongate the plasma to diverter configurations.
Besides the 14 superconducting PF coils outside the vaccum vessel, EAST has two small copper coils (2 turn/coil) within the vessel, which share a power supply and are connected in anti-series, and thus have opposite currents. They are closer to the plasma (than other PF coils) and are used to control fast plasma displacements, specifically VDEs (vertical displacement events). They are often called “fast control coils”.
Using Ampere's circuital law
![]() |
(A.5) |
along the toroidal direction and assuming perfect toroidal symmetry, we obtain
![]() |
(A.6) |
i.e.,
![]() |
(A.7) |
Neglecting the poloidal current contributed by the plasma, the poloidal
current is determined solely by the current in the TF coils. The EAST
tokamak has 16 groups of TF coils with 132turns/coil (I got to know the
number of turns from ZhaoLiang Wang: ).
Denote the current in a single turn by
,
then Eq. (A.7) is written
![]() |
(A.8) |
Using this formula, the strength of the toroidal magnetic field at for
is calculated to be
. This was one of the two scenarios
often used in EAST experiments (another scenario is
). (The limit of the current in a single turn
of the TF coils is
(from B. J. Xiao's paper [31]).
Note that the exact equilibrium toroidal magnetic field
is given by
. Compare this
with Eq. (A.7), we know that the approximation made to
obtain Eq. (A.8) is equivalent to
, i.e. assuming
is a
constant. The poloidal plasma current density
is
related to
by
.
The constant
corresponds to zero plasma poloidal
current, which is consistent to the assumption used to obtain Eq. (A.8).
Let us estimate the safety factor value near the plasma edge using the
total plasma current and the current in a single turn of TF coils . For divertor magnetic
configuration, the plasma edge is at the saperatrix, where
. To get a characteristic safety factor value
that is finite, one often chooses the edge to be the magnetic surface
that encloses
of the poloidal magnetic flux.
Denote this surface by
and the value of
at this surface by
,
which is given by
![]() |
(A.9) |
where is the minor radius of the surface
, and
the
major radius of the magnetic axis,
is the the
magnitude of toroidal magnetic field at the magnetic axis, and
is the average poloidal magnetic field on the surface,
. Using Eq. (A.8),
Eq. (A.9) is written as
![]() |
(A.10) |
For EAST, tipically and
. Using this, we obtain
![]() |
(A.11) |
The so-called resonant magnetic perturbation (RMP) coils are 3D coils that are used to suppress or mitigate edge localized modes. The shape and location of RMP coils of EAST tokamak are plotted in Fig. A.2.
The size of EAST is similar to that of DIII-D tokamak. The main parameters are summarized in Table. A.1. A significant difference between EAST and DIII-D is that DIII-D has a larger minor radius, which makes DIII-D able to operate with a larger toroidal current than that EAST can do for the same current density. Another significant difference between EAST and DIII-D is that the coils of EAST are supper-conducting while the coils of DIII-D are not. The supper-conducting coils enable EAST to operate at longer pulse.
|
|||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||
Table A.1. Comparison of main
parameters of EAST, DIII-D, JET, and ITER tokamaks. The |
|
||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||
Table A.2. Continued from Table A.1. |
|
||||||||||||||||||||||||||||||||||||||||||||||||||||||||||||
Table A.3. Continued from Table A.2. |
DIII-D has 24 groups of TF coils with 6turns/coil, i.e., total turns are
, with a maximum current of
in a single turn[20]. Using formula
(A.7), the toroidal filed at
can be
calculated, giving
.
DIII-D is special in that its poloidal field (PF) coils are located inside of the TF-coils, which makes the PF-coils more close to the plasma and thus more efficient in shaping the plasma. However, this nested structure is difficult to assemble. In superconducting tokamaks (e.g., EAST, KSTAR, ITER), PF coils are all placed outside of the TF-coils.
I noticed that HL-2M also has the PF coils located within the TF-coils, similar to DIII-D. This remind me that this layout may apply to all non-superconducting tokamaks (to be confirmed, No, non-superconducting tokamak ASDEX-U has PF coils outside of TF coils).
KSTAR has 16 TF coils and 14 PF coils. Both of the TF and PF coil system
use internally cooled superconductors. The nominal current in TF coils
is with
and all coils
connected in series[25]. Using these information and
formula (A.7), the toroidal filed at
can be calculated, giving
.
The PF coil system consists of 8 Central Solenoide (CS) coils and 6
outer PF coils and can provide 17 V-sec.
ITER has 18 TF coils with number of turns in one coil being 134 and
current per turn [4].
Using these information and formula (A.7), the toroidal
filed at
can be calculated, giving
According to Refs. [7, 21], Miller's formula for a series of shaped flux surfaces is given by
![]() |
(A.12) |
![]() |
(A.13) |
where and
are elongation
and triangularity profile,
is the Shafranov
shift profile, which is given by
![]() |
(A.14) |
where is a constant,
is
the major radius of the center of the boundary flux surface. The
triangularity profile is
![]() |
(A.15) |
and the elongation profile is
![]() |
(A.16) |
The nominal ITER parameters are ,
and
.
I wrote a code to plot the shapes of the flux surface
(/home/yj/project/miller_flux_surface). An example of the results is
given in Fig. A.3.
<image|/home/yj/project/miller_flux_surface/plt.eps||||> |
An analytic expression for the pressure profile of double (inner and external) transport barriers is given by
![]() |
(A.17) |
where is the normalized poloidal flux,
and
are the width of the inner and
external barriers,
and
are the locations of the barriers,
and
is the height of the barriers,
is
a constant to ensure
at
.
![]() |
Figure A.4. Pressure profile of double (inner
and external) transport barriers given by Eq. (A.17)
with |
![]() |
To construct a periodic function about ,
we introduces a function
which is defined over
and vanishes sufficiently fast as
so that the following infinite summation converge:
![]() |
(A.18) |
If we use the above sum to define a function
![]() |
(A.19) |
then it is obvious that
![]() |
(A.20) |
i.e., is a periodic function about
with period of
.
If we use the right-hand-side of Eq. (A.19) to represent
, then we do not need to
worry about the periodic property of
(the
periodic property is guaranteed by the representation)
Let us introduce parameters characterizing the shape of a magnetic
surface in the poloidal plane. The “midplane” is defined as
the plane that passes through the magnetic axis and is perpendicular to
the symmetric axis ( axis). For a up-down
symmetric (about the midplane) magnetic surface, its shape can be
roughly characterized by four parameters, namely, the
coordinate of the innermost and outermost points in the midplane,
and
; the
coordinators of the highest point of the
magnetic surface,
. These
four parameters are indicated in Fig. A.6.
![]() |
In terms of these four parameters, we can define the major radius of a magnetic surface
![]() |
(A.21) |
(which is the coordinate of the geometric center
of the magnetic surface), the minor radius of a magnetic surface
![]() |
(A.22) |
the triangularity of a magnetic surface
![]() |
(A.23) |
and, the ellipticity (elongation) of a magnetic surface
![]() |
(A.24) |
Usually, we specify the value of ,
,
, and
,
instead of
, to characterize
the shape of a magnetic surface. The value of the triangularity
is usually positive in traditional tokamak
operations, but negative triangularity is achievable and potentially
useful, which is under active investigation.
Besides, using and
,
we can define another useful parameter
,
which is called the inverse aspect ratio. Note that the major radius
of the LCFS is usually different from
(the
coordinate of the magnetic
axis). Usually we have
due to the so-called
Shafranov shift.
We note that is the covariant toroidal component
of
in cylindrical coordinates
. The proof is as follows. Note that the
covariant form of
should be expressed in terms
of the contravariant basis vector (
,
, and
), i.e.,
![]() |
(A.25) |
where is the covariant toroidal component of
. To obtain
, we take scalar product of Eq. (A.25)
with
and use the orthogonality relation (8.7), which gives
![]() |
(A.26) |
In cylindrical coordinates ,
the position vector is written as
![]() |
(A.27) |
where ,
, and
are unit vectors
along
,
, and
,
respectively, i.e.
![]() |
(A.28) |
Using this, we obtain
![]() |
(A.29) |
Use Eq. (A.29) in Eq. (A.26) giving
![]() |
(A.30) |
with defined by
.
Equation (A.30) indicates that
is
the covariant toroidal component of the vector potential.
As is discussed in Sec. (), to satisfy the force balance, must be a magnetic surface function, i.e.,
. Using this, expression (3.2) and
(3.3) are written
![]() |
(A.31) |
and
![]() |
(A.32) |
respectively. The above two equations imply that
![]() |
(A.33) |
which implies that the projections of lines and
lines in the poloidal plane are identical to
each other. This indicates that the
surfaces
coincide with the magnetic surfaces.
The poloidal plasma current density is usually small (compared with the
toroidal plasma current density) but can be important for some cases of
interest and thus could not be safely neglected. Many model equilibria
(e.g., Solovev equilibria, DIII-D cyclone base cases) frequently used in
simulations assume that is a spatial constant,
i.e., neglecting the poloidal plasmas current. The conclusions drawn
from these simulations could be misleading.
Using this and , and
following the same steps in Sec. 1.7, we obtain
![]() |
(A.34) |
where is the poloidal current enclosed by the
two magnetic surfaces, the positive direction of
is chosen to be in the clockwise direction when observers look along
. Equation (A.34)
indicates that the difference of
between two
magnetic surface is proportional to the poloidal current. For this
reason,
is usually call the “poloidal
current function”.
In the above, we see that the relation of with
the poloidal electric current is similar to that of
with the poloidal magnetic flux. This similarity is due to that
is isomorphic to
The poloidal plasma current density can be
further written as
Using , Eq. (A.35)
can also be written as
![]() |
![]() |
![]() |
|
![]() |
![]() |
For most choices of and
, the GS equation (4.13) has to be
solved numerically. For the specific choice of
and
profiles given by
![]() |
(A.37) |
![]() |
(A.38) |
where ,
, and
are constants,
there is an analytical solution, which is given by[15]
![]() |
(A.39) |
where is an constant. [Proof: By direct
substitution, we can verify
of this form is
indeed a solution to the GS equation (4.13).] A useful
choice for tokamak application is to set
,
, and
. Then Eq. (A.39) is written
![]() |
(A.40) |
For this case, the toroidal field function is a
constant. (For the Solovev equilibrium (A.40), I found
numerically that the value of the safety factor at the magnetic axis
(
) is equal to
. This result should be able to be proved
analytically. I will do this later. In calculating the safety factor, we
also need the expression of
,
which is given analytically by
![]() |
![]() |
![]() |
|
![]() |
![]() |
)
Define , and
, then Eq. (A.40) is written as
where ,
. From Eq. (A.42), we obtain
![]() |
(A.43) |
Given the value of ,
, for each value of
, we can plot a magnetic surface on
plane. An example of the nested magnetic surfaces is shown
in Fig. A.7.
![]() |
Figure A.7. Flux surfaces of
Solovév equilibrium for |
Equation (A.40) can be solved to give explicit form of the
contour of in
plane:
![]() |
(A.44) |
The minor radius of a magnetic surface of the Solovev equilibrium can be calculated by using Eq. (A.44), which gives
![]() |
(A.45) |
![]() |
(A.46) |
and thus
![]() |
(A.47) |
where . In my code of
constructing Solovev magnetic surface, the value of
is specified by users, and then Eq. (A.47) is solved
numerically to obtain the value of
of the flux
surface. Note that the case
corresponds to
, i.e., the magnetic axis, while
the case
corresponds to
. Therefore, the reasonable value of
of a magnetic surface should be in the range
. This range is used as the interval bracketing
a root in the bisection root finder.
Using Eq. (A.47), the inverse aspect ratio of a magnetic
surface labeled by can be approximated as[15]
![]() |
(A.48) |
Therefore, the value of of a magnetic surface
with the inverse aspect ratio
is approximately
given by
![]() |
(A.49) |
In deriving the Grad-Shafranov equation, we have assumed that there is no plasma flow. Next, let us examine whether this assumption is justified for plasmas in EAST tokamak.
The complete momentum equation is given by
![]() |
(A.50) |
where term can be usually neglected due to
either
or
,
is a pressure tensor, which is different from
the scalar pressure considered in this note. The equilibrium with
pressure tensor can be important for neutral beam heating plasma, where
pressure contributed by NBI fast ions can be a tensor. With plasma flow
and scalar pressure and neglecting electric force term, the steady state
momentum equation is written
![]() |
(A.51) |
where the term on the left-hand side is the contribution of plasma flow to the force balance. Let us estimate the magnitude of this term. Macroscopic flows in tokamak are usually along the toroidal direction (the poloidal flow is usually heavily damped). The toroidal flow usually has the same toroidal angular frequency on a magnetic surface, i.e., the flow can be written as
![]() |
(A.52) |
where is the toroidal angular frequency of the
flow, which can have radial variation. Using this expression, the
left-hand side of Eq. (A.51) is written as
![]() |
![]() |
![]() |
|
![]() |
![]() |
||
![]() |
![]() |
which is just the centripetal force. For typical EAST plasma, the
rotation frequency is less than
. For
,
and
[Mass density of
EAST#38300@3.9s,
], the above
fore is about
newton. On the other hand, the
pressure gradient force in Eq. (A.51) is about
newton, which is one order larger than the force
contributed by the rotation. [Pressure gradient force is estimated by
using
, where
is the thermal pressure at the magnetic axis and
is the minor radius of plasma. For typical EAST plasmas
(EAST#38300@3.9s),
,
. Then
.] This indicates the rotation has little influence
on the force balance. In other words, the EAST tokamak equilibrium can
be well described by the static equilibrium without rotation.
Plasma is is the ratio of thermal pressure to
magnetic pressure, i.e.,
![]() |
(A.54) |
Since pressure in tokamak plasmas is not uniform, volume averaged
pressure is used to define the beta. The toroidal beta
and the poloidal beta
are defined, respectively,
by
![]() |
(A.55) |
![]() |
(A.56) |
where is the volume average,
is the vacuum toroidal magnetic field at the magnetic axis (or
geometrical center of the plasma),
is the flux
surface averaged poloidal magnetic fied,
is the
length of the poloidal perimeter of the boundary flux surface. Then
can be written as
![]() |
(A.57) |
An alternative defintion of use the LCFS
cross-sectional average of
rather than the
volume average
(Wesson tokamaks). This
definition is adopted in many codes, including HEQ code I developed
Beta can be considered as a quantity characterizing the efficiency of the magnetic field of tokamaks in confining plasmas.
In tokamaks, the toroidal magnetic field is dominant and thus the the
toroidal beta (not
)
is the usual way to characterize the the efficiency of the magnetic
field in confining plasmas. (Why do we need
? The short answer is that
is proportional to an important current, the bootstrap current, which is
important for tokamak steady state operation.)
Tokamak experiments have found that it is easier to achieve higher in low
plasmas than in higher
plasmas, which indicates that the efficiency of
the magnetic field in confining plasma is a decreasing function of the
magnitude of the magnetic field.
Beta limit means there is a limit for the value of beta beyond which the
plasma will encounter a serious disruption. Early calculation of the
beta limit on JET shows that the maximal
obtained is proportional to
,
where all quantities are in SI units. This scaling relation
is often called Troyon scaling, where the coefficient
was determined numerically by Troyon to 0.028. Often
is expressed in percent, in which case
. This motivates us to define
![]() |
(A.58) |
which is called “normalized beta”. The normalized beta is an operational parameter indicating how close the
plasma is to reach destabilizing major MHD activities. Its typical value
is of order unit. As mentioned above,
calculated
by Troyon is
. Empirical
evaluation from the data of different tokamaks raises this value
slightly to 3.5, although significantly higher values, e.g.,
, have been achieved in the low
aspect ratio tokamak NSTX[24].
The value of indicates how close one is to the
onset of deleterious instability . The ability to increase the value of
can be considered as the ability of controlling
the major MHD instabilities, and thus can be used to characterize how
well a tokamak device is operated. One goal of EAST tokamak during
2015-2016 is to sustain a plasma with
for at
least 10 seconds.
(check** The tearing mode, specifically the neoclassical tearing mode (NTM) is expected to set the beta limit in a reactor.)
(**check: Tokamak experiments have found that it is easier to achieve
high in large
plasmas
than in small
plasmas. However, experiments
found it is easier to achieve high
in small
plasmas than in large
plasmas. Examining the expression of
and
given by Eqs. (A.57) and Eq. (),
respectively, we recognize that pressure limit should have a scaling of
with
.
)
is the ratio of the pressure to the squre of
plasma current, and thus characterizes the efficiency of the plasma
current in confining the plasma. Is there a limit for
?
As is mentioned in Sec. A.14, the beta limit study on JET
tokamak shows that the maximal obtained is
proportional to
. This means
the maximal plasmas pressure obtained is proportional to
, i.e.,
![]() |
(A.59) |
The total plasma energy is given by ,
where
is the major radius of the device. Using
Eq. (A.59), we obtain
![]() |
(A.60) |
Since fusion power is proportional to the plasma energy, the above relation indicates, to obtain larger fusion power, we need bigger tokamaks with larger plasma current.
The reason why larger plasma current is desired can also be appreciated
by examining an empirical scaling of the the energy confinement time
given by
![]() |
(A.61) |
which is proportional to .
Another fundamental reason for building larger tokamaks is that the
energy confinement time increases with the
machine size, where
is the heat diffusitivity.
In addition, the heat diffusitivity
decreases
with increasing machine size if the diffusitivity satisfies the
gyro-Bohm scaling, which is given by
![]() |
(A.62) |
which is inverse proportional to the machine size . However, if the diffusitivity satisfies the Bohm
scaling, which is given by
![]() |
(A.63) |
then, the diffusitivity is independent of the machine size. The Bohm
diffusitivity is
times
larger than the gyro-Bohm diffusitivity
,
where
is the gyro-radius. Heat diffusitivity
scaling in the low confinement operation (
mode)
in present-day tokamaks is observed to be Bohm or worse than Bohm.
The maximum density that can be obtained in stable plasma operations (without disruption) is empirically given by
![]() |
(A.64) |
where is the Greenwalt density, which is given
by
![]() |
(A.65) |
where is the plasma current,
is the minor radius, all physical quantities are in SI units. The
gives the density limit that can be achieved for a
tokamak operation scenario with plasma current
and plasma minor radius
. The
Greenwalt density limit is an empirical one, which, like other empirical
limits, can be exceeded in practice. Equation (A.65)
indicates that the Greenwalt density is proportional to the current
density. Therefore the ability to operate in large plasma current
density means the ability to operate with high plasma density.
Note that neither the pressure limit nor the density limit is determined by the force-balance constraints. They are determined by the stability of the equilibrium. On the other hand, since the stability of the equilibrium is determined by the equilibrium itself, the pressure and density limit is determined by the equilibrium.
The self-inductance of a current loop is defined as the ratio of the
magnetic flux traversing the loop and its
current
:
![]() |
(A.66) |
where
![]() |
(A.67) |
and is generated by the current
. It can be proved that
is independent of the current
in the loop, i.e.,
is fully determined by the shape of the loop.
On the other hand, the energy contained in the magnetic field produced by the loop current is given by
![]() |
(A.68) |
where the volume includes all space where is not
negligible. It can be proved that
,
, and
are related to each other by:
![]() |
(A.69) |
i.e.,
![]() |
(A.70) |
which can be considered an equivalent definition of the self-inductance.
The internal inductance of tokamak plasma is
defined in such a way that
only includes the
poloidal magnetic energy within the plasma. Specifically,
is defined by
![]() |
(A.71) |
where only energy of the poloidal magnetic field in included, and the
integration over the plasma volume .
Its complement is the external inductance (
).
The normalized internal inductance is defined as
![]() |
(A.72) |
where is the surface averaged major radius.
Using Eq. (A.71), expression (A.72) is written
as
![]() |
(A.73) |
which is the definition of used in the ITER
design. Using the defintion of
,
i.e.,
![]() |
(A.74) |
where is the cross section of boundary flux
surface, expression (A.73) is written as
![]() |
(A.75) |
Another definition of is
![]() |
(A.76) |
where is the surface average over the plasma
boundary. Using Ampere's law, we obtain
![]() |
(A.77) |
Using this Eq. (A.76) is written as
![]() |
(A.78) |
For circular cross section, .
Then the above expression is written as
which agrees with Eq. (A.75).
The normalized internal inductance reflects the peakness of the current
density profile (I have not verified this): smaller value of corresponds to broader current profile.
Since , the current density
can be inferred from a given magnetic field. The
components of
(expressed in terms of
and
) are given
by Eqs. (3.2), (3.3) and (3.6).
These expressions can be further simplified by using the equilibrium
constraints, such as
and
. The simplified
expressions
are given in A.20.
On the other hand, in the kinetic equilibrium reconstruction[19], it is the current that is first (partially) specified (e.g., by summing sources of current drive and bootstrap current) and then the current is used as constraints for the GS equation, i.e, constraints for the magnetic field.
Due to the force balance condition, the plasma current is related to the plasma pressure gradient:
The parallel (to the magnetic field) current density is written as
![]() |
![]() |
![]() |
|
![]() |
![]() |
||
![]() |
![]() |
||
![]() |
![]() |
||
![]() |
![]() |
||
![]() |
![]() |
For later use, define
Equation (A.81) is used in GTAW code to calculate (actually calculated is
)[14]. Note that the expression for
in
Eq. (A.81) is not a magnetic surface function. Define
as
![]() |
![]() |
![]() |
|
![]() |
![]() |
||
![]() |
![]() |
where is called Pfirsch-Schluter (PS) current.
In cylindrical geometry, due to the poloidal symmetry, the
Pfiersch-Schluter current is zero. In toroidal geometry, due to the
poloidal asymmetry, the PS current is generally nonzero. Thus, this
quantity characterizes a toroidal effect.
Another useful quantity is ,
which is written as
![]() |
where is flux surface averaging operator.
In the vacuum region that is between the plasma and the first wall, there is no current, i.e.,
![]() |
(A.85) |
Next, let us examine what constraint this condition imposes on the magnetic field. Using expression (1.10), the above equation is written as
![]() |
(A.86) |
It is not obvious how to draw useful information from the above equation.
In the above, we use the vector potential approach. Next, let us try the scalar potential approach:
![]() |
(A.87) |
then the constraint is written as
![]() |
(A.88) |
[check***As discussed in Sec. (), the force balance equation of
axisymmetric plasma requires that .
From this and the fact
, we
conclude that
is a function of
, i.e.,
.
However, this reasoning is not rigorous. Note the concept of a function
requires that a function can not be a one-to-more map. This means that
indicates that the values of
must be equal on two different magnetic field lines that have the same
value of
. However, the two
equations
and
do not
require this constraint. To examine whether this constraint removes some
equilibria from all the possible ones, we consider a system with an
point. Inside one of the magnetic islands, we use
![]() |
(A.89) |
and inside the another, we use
![]() |
(A.90) |
Then solve the two GS equations respectively within the boundary of the
two islands. It is easy to obtain two magnetic surfaces that have the
same value of respectively inside the two
islands. Equations (A.89) and (A.90) indicate
that the values of
on the two magnetic surfaces
are different from each other. It is obvious the resulting equilibrium
that contain the two islands can not be recovered by directly solving a
single GS equation with a given function
.**check]
On both an irrational surface and a rational surface, there are infinite number of magnetic field lines that are not connected with each other. It is wrong to say there is only one magnetic field line on a irrational surface.
Consider a perturbation that satisfies
, i.e., the value
is a constant along any one of the magnetic field lines (field-aligned
structure). Now comes the question: whether the values of
on different field lines are equal to each other?
To answer this question, we choose a direction that is different from
on the magnetic surface and examine whether
is constant along this direction, i.e, whether
equals zero, where
is the
chosen direction. For axsiymmetric magnetic surfaces, then
is a direction in the magnetic surface and it is usually
not identical with
. Then we
obtain
![]() |
(A.91) |
If is non-axisymmetric, then Eq. (A.91)
indicates that
, i.e., the
values of
on different magnetic field lines on
the same magnetic surface are not equal to each other. This is the case
for the
coordinates discussed in Sec. 15.1.
Realistic perturbations in tokamak plasmas only have approximate
field-aligned structure, i.e., .
Put it in other words,
.
This reasoning is for the case of axsiymmetric magnetic surfaces. It is
ready to do the same reasoning for non-axisymmetrica magnetic surface
after we find a convenient direction on the
magnetic surface.
(In practice, I choose the positive direction of
and
along the direction of toroidal and poloidal
magnetic field (i.e.,
and
are always positive in
coordinates system).
Then, the
defined by Eq. (8.81) is
always positive. It follows that
should be also
positive. Next, let us examine whether this property is correctly
preserved by Eq. (8.87). Case 1: The radial coordinate
is chosen as
.
Then the factor before the integration in Eq. (8.87) is
negative. We can further verify that
is always
negative for either the case that
is pointing
inward or outward. Therefore the .r.h.s. of Eq. (8.87) is
always positive for this case. Case 2: The radial coordinate
is chosen as
.
Then the factor before the integration in Eq. (8.87) is
positive. We can further verify that
is always
positive for either the case that
is pointing
inward or outward. The above two cases include all possibilities.
Therefore, the positivity of
is always
guaranteed)
We know that the toroidal flux ,
safety factor
, and the
in the GS equation are related by the following
equations:
![]() |
(A.92) |
![]() |
(A.93) |
Define:
![]() |
(A.94) |
(In the Toray_ga code, the radial coordinate is
defined as
![]() |
(A.95) |
where is a constant factor.
defined this way is of length dimension, which is an effective geometry
radius obtained by approximating the flux surface as circular.)
I use Eq. (A.94) to define .
Then we have
![]() |
(A.96) |
![]() |
(A.97) |
![]() |
(A.98) |
![]() |
(A.99) |
![]() |
(A.100) |
Eq. (A.100) is used to transform between
and
.
If are magnetic surface coordinates, i.e.,
, then the toroidal elliptic
operator in Eq. (13.3) is reduced to
![]() |
(A.101) |
and the GS equation, Eq. (13.5), is reduced to
![]() |
(A.102) |
Define coordinates by
![]() |
(A.103) |
![]() |
(A.104) |
where are the cylindrical coordinates and
is a constant. The above transformation is shown
graphically in Fig. A.9.
The Jacobian of coordinates can be calculated
using the definition. Using
,
, and
, the Jacobian (with respect to the Cartesian
coordinates
) is written as
![]() |
![]() |
![]() |
|
![]() |
![]() |
||
![]() |
![]() |
||
![]() |
![]() |
||
![]() |
![]() |
||
![]() |
![]() |
Next, we transform the GS equation from
coordinates to
coordinates. Using the relations
and
,
we have
![]() |
(A.106) |
![]() |
(A.107) |
![]() |
(A.108) |
![]() |
(A.109) |
The GS equation in coordinates is given by
![]() |
(A.110) |
The term is written as
Using Eq. (A.111), is written as
![]() |
(A.113) |
![]() |
(A.114) |
![]() |
(A.115) |
Summing the the right-hand-side of Eq. (A.112) and the expression on line (A.116) yields
![]() |
(A.118) |
Using these, the GS equation is written as
which can be arranged in the form
![]() |
(A.119) |
which agrees with Eq. (3.6.2) in Wessson's book[29], where
is defined by
,
which is different from
by a
factor.
Consider the case that the boundary flux surface is circular with radius
and the center of the cirle at
. Consider the case
. Expanding
in the small
parameter
,
![]() |
(A.120) |
where ,
. Substituting Eq. (A.120) into
Eq. (A.119), we obtain
Multiplying the above equation by ,
we obtain
![]() |
(A.121) |
Further assume the following orderings (why?)
![]() |
(A.122) |
and
![]() |
(A.123) |
Using these orderings, the order of the terms in Eq. (A.121) can be estimated as
![]() |
(A.124) |
![]() |
(A.125) |
![]() |
(A.126) |
![]() |
(A.127) |
![]() |
(A.128) |
![]() |
(A.129) |
![]() |
(A.130) |
![]() |
(A.131) |
![]() |
(A.132) |
The leading order ( order) balance is given by
the following equation:
![]() |
(A.133) |
It is reasonable to assume that is independent
of
since
corresponds to
the limit
. (The limit
can have two cases, one is
, another is
.
In the former case,
must be independent of
since
should be
single-valued. The latter case corresponds to a cylinder, for which it
is reasonable (really?) to assume that
is
independent of
.) Then Eq.
(A.133) is written
![]() |
(A.134) |
(My remarks: The leading order equation (A.134) does not
corresponds strictly to a cylinder equilibrium because the magnetic
field depends on
.)
The next order (
order) equation is
![]() |
(A.135) |
![]() |
(A.136) |
It is obvious that the simple poloidal dependence of
will satisfy the above equation. Therefore, we consider
of the form
![]() |
(A.137) |
where is a new function to be determined.
Substitute this into the Eq. (), we obtain an equation for
,
![]() |
(A.138) |
![]() |
(A.139) |
![]() |
(A.140) |
![]() |
(A.141) |
Using the identity
equation () is written as
![]() |
(A.142) |
Using the leading order equation (), we know that the second and fourth term on the l.h.s of the above equation cancel each other, giving
![]() |
(A.143) |
![]() |
(A.144) |
Using the identity
![]() |
(A.145) |
equation (A.144) is written
![]() |
(A.146) |
![]() |
(A.147) |
Using
![]() |
(A.148) |
equation (A.147) is written
![]() |
(A.149) |
![]() |
(A.150) |
which agrees with equation (3.6.7) in Wessson's book[29].
The normalized pressure gradient, ,
which appears frequently in tokamak literature, is defined by[3]
![]() |
(A.151) |
which can be further written
![]() |
(A.152) |
where . Equation (A.152)
can be further written as
![]() |
(A.153) |
where ,
, and
is the minor
radius of the boundary flux surface. (Why is there a
factor in the definition of
?)
The global magnetic shear is defined by
![]() |
(A.154) |
which can be written
![]() |
(A.155) |
In the case of large aspect ratio and circular flux surface, the leading
order equation of the Grad-Shafranov equation in
coordinates is written
![]() |
(A.156) |
which gives concentric circular flux surfaces centered at . Assume that
is
uniform distributed, i.e.,
,
where
is the total current within the flux
surface
. Further assume the
current is in the opposite direction of
,
then
. Using this, Eq. (A.156) can be solved to give
![]() |
(A.157) |
Then it follows that the normalized radial coordinate
relates to
by
(I check
this numerically for the case of EAST discharge #38300). Sine in my
code, the radial coordinate is
,
I need to transform the derivative with respect to
to one with respect to
,
which gives
![]() |
(A.158) |
![]() |
(A.159) |
The necessary condition for the existence of TAEs with frequency near the upper tip of the gap is given by[3]
![]() |
(A.160) |
which is used in my paper on Alfvén eigenmodes on EAST tokamak[14]. Equations (A.158) and (A.159)
are used in the GTAW code to calculate and
.
Magnetic field (e.g., equilibrium poloidal field, RMP field, ripple field) and vector potential generated by coils can be calculated in the following way.
The Biot-Savart law for a zero-thinkness wire is given by
![]() |
where is a line element along the wire (the
freedom of choosing which one of the two possible directions along the
wire is up to users),
is the current in the wire
(
is positive if the current in the same
direction of
, otherwise,
negative).
Given a current source , the
vector potential can be calculated using
![]() |
(B.2) |
where
![]() |
(B.3) |
is called the retarded time. For a steady-state source, . Then Eq. (B.2) is simplified as
![]() |
(B.4) |
For a current flowing in a thin wire, the above equation is written as
![]() |
(B.5) |
where is a surface element perpendicular to the
wire and
is a line element along the wire. Using
, the above eqaution is
written as
![]() |
![]() |
![]() |
|
![]() |
![]() |
In cylindrical coordinates, the toroidal component of ,
,
is written as
![]() |
(B.7) |
( is needed in several applications. In solving
the free-boundy equilibrium problem, we need to calculate
generated by the PF coils. In studying the effects of RMP
on particles, we need to calculate
,
the canonical toroidal angular momentum, whose defintion involves
. So we need to calculate
generated by the RMP coils.)
In terms of the cylindrical basis, the line element
of a toroidal curve is given by
.
Using Cartesian coordinate basis vectors,
is
written as
![]() |
(B.8) |
and the position vector is written as
![]() |
(B.9) |
where a global Cartesian coordinates with axis
in
plane is assumed.
Let us consider a full toroidal coil at with
current
(a filamentary current loop). Since the
problem is axisymmetric, the magnetic field/potential is independent of
. For calculation ease, we
select
plane and calculate
and
values in this plane. Then
![]() |
(B.10) |
![]() |
(B.11) |
Using these, we obtain
![]() |
(B.12) |
and Eq. (B.7) is written as
![]() |
(B.13) |
Because is axsymmetric, the above formula
applies to all values of
.
Multiply
by
and set
, we get Green's function for
:
This expression can be directly used in numerical codes to compute the
Green function table. From this formula, we can see that the source
location and observation location are symmetric, i.e., .
One can also write the above integration to the elipitic integrals and
use library to compute these inegreals. Let us write the above expresion
in this form. Define , then
. Using this Eq. (B.14)
is written as
![]() |
|||
![]() |
![]() |
||
![]() |
![]() |
||
![]() |
![]() |
Define
![]() |
(B.16) |
Then
![]() |
|||
![]() |
![]() |
||
![]() |
![]() |
||
![]() |
![]() |
||
![]() |
![]() |
||
![]() |
![]() |
where
are the eliptic integrals of the first and second kind.
Similarly, we can calculate the magnetic field:
where the cross product is written as
The and
components of
the above result are written as
Using these in Eq. (B.18), and
are written as
Dropping in the above expressions gives the
corresponding Green's function:
![]() |
![]() |
One can also try to write the above formula to eliplicity integral form.
The ressult is not compact as the formula for . So I directly use Eqs. (B.18) and (B.19) in HEQ code. Alternatively, one can directly take the
finite difference of
to get
and
in a numerical code.
For a curve in the poloidal plane, the line element
in terms of the cylindrical basis is written as
. Using Cartesian coordinate basis vectors,
is written as
![]() |
(B.21) |
Using Eq. (B.21) and (), we obtain
![]() |
(B.22) |
This document initially contains notes I took when learning tokamak equilibrium theory. As a result of the on-going learning, the document is evolving. I enjoy seeing the continuous improvement of this document over a long time scale (10 years). Most of the content is elementary but I still found I got something totally wrong in the first version. Let me know if you spot something specious.
This document was written by using TeXmacs[27], a
structured wysiwyw (what you see is what you want) editor for
scientists. The HTML version is generated by first converting the
TeXmacs file to TeX format and then using htlatex
to
convert the TeX to HTML.
R Aymar, P Barabaschi, and Y Shimomura. The ITER design. Plasma Physics and Controlled Fusion, 44(5):519–565, apr 2002.
M. A. Beer, S. C. Cowley, and G. W. Hammett. Field-aligned coordinates for nonlinear simulations of tokamak turbulence. Phys. Plasmas (1994-present), 2(7):2687–2700, 1995.
H. L. Berk, J. W. Van Dam, D. Borba, J. Candy, G. T. A. Huysmans, and S. Sharapov. More on core localized toroidal alfvén eigenmodes. Phys. Plasmas, 2(9):3401–3406, 1995.
R. Gallix, C. Jong, J. Knaster, N. Mitchell C. Sborchia, Y. Fu. Design and specifications of the iter tf coils. IEEE TRANSACTIONS ON APPLIED SUPERCONDUCTIVITY,, 18(2):463, 2008.
Yang Chen and Scott E. Parker. A delta-f particle method for gyrokinetic simulations with kinetic electrons and electromagnetic perturbations. Journal of Computational Physics, 189(2):463–475, 2003.
Yang Chen and Scott E. Parker. Electromagnetic
gyrokinetic f
particle-in-cell turbulence simulation with realistic
equilibrium profiles and geometry. Journal of Computational
Physics, 220(2):839–855, 2007.
Yang Chen, Scott E. Parker, J. Lang, and G.-Y. Fu. Linear gyrokinetic simulation of high-n toroidal alfvén eigenmodes in a burning plasma. Phys. Plasmas, 17(10):102504, 2010.
Fu Peng, Gao Ge Chen Yuan-Yang, Bao Xiao-Hua. Plasma shape optimization for east tokamak using orthogonal method. Chinese Physics B, 28(1):15201, 2019.
C.Z Cheng and M.S Chance. Nova: a nonvariational code for solving the mhd stability of axisymmetric toroidal plasmas. J. of Comput. Phys., 71(1):124–146, 1987.
J DeLucia, S.C Jardin, and A.M.M Todd. An iterative metric method for solving the inverse tokamak equilibrium problem. Journal of Computational Physics, 37(2):183–204, 1980.
A. M. Dimits. Fluid simulations of tokamak turbulence in quasiballooning coordinates. Phys. Rev. E, 48:4070–4079, Nov 1993.
T. Görler, N. Tronko, W. A. Hornsby, A. Bottino, R. Kleiber, C. Norscini, V. Grandgirard, F. Jenko, and E. Sonnendrücker. Intercode comparison of gyrokinetic global electromagnetic modes. Physics of Plasmas, 23(7):72503, 2016.
J. van der Hoeven et al. GNU TeXmacs. https://www.texmacs.org, 1998.
Youjun Hu, Guoqiang Li, N. N. Gorelenkov, Huishan Cai, Wenjun Yang, Deng Zhou, and Qilong Ren. Numerical study of alfvén eigenmodes in the experimental advanced superconducting tokamak. Phys. Plasmas, 21(5):0, 2014.
Stephen C. Jardin. Computational methods in plasma physics. CRC Press, 2010.
Young Mu Jeon. Development of a Free-boundary Tokamak Equilibrium Solver for Advanced Study of Tokamak Equilibria. Journal of the Korean Physical Society, 67(5):843–853, 09 2015.
L.L. Lao, H. St. John, R.D. Stambaugh, A.G. Kellman, and W. Pfeiffer. Reconstruction of current profile parameters and plasma shapes in tokamaks. Nucl. Fusion, 25(11):1611, 1985.
G.S. Lee, M. Kwon, C.J. Doh, B.G. Hong, K. Kim, M.H. Cho, W. Namkung, C.S. Chang, Y.C. Kim, J.Y. Kim, H.G. Jhang, D.K. Lee, K.I. You, J.H. Han, M.C. Kyum, J.W. Choi, J. Hong, W.C. Kim, B.C. Kim, J.H. Choi, S.H. Seo, H.K. Na, H.G. Lee, S.G. Lee, S.J. Yoo, B.J. Lee, Y.S. Jung, J.G. Bak, H.L. Yang, S.Y. Cho, K.H. Im, N.I. Hur, I.K. Yoo, J.W. Sa, K.H. Hong, G.H. Kim, B.J. Yoo, H.C. Ri, Y.K. Oh, Y.S. Kim, C.H. Choi, D.L. Kim, Y.M. Park, K.W. Cho, T.H. Ha, S.M. Hwang, Y.J. Kim, S. Baang, S.I. Lee, H.Y. Chang, W. Choe, S.G. Jeong, S.S. Oh, H.J. Lee, B.H. Oh, B.H. Choi, C.K. Hwang, S.R. In, S.H. Jeong, I.S. Ko, Y.S. Bae, H.S. Kang, J.B. Kim, H.J. Ahn, D.S. Kim, C.H. Choi, J.H. Lee, Y.W. Lee, Y.S. Hwang, S.H. Hong, K.-H. Chung, D.-I. Choi, and KSTAR Team. Design and construction of the KSTAR tokamak. Nuclear Fusion, 41(10):1515–1523, oct 2001.
G Q Li, Q L Ren, J P Qian, L L Lao, S Y Ding, Y J Chen, Z X Liu, B Lu, and Q Zang. Kinetic equilibrium reconstruction on east tokamak. Plasma Phys. Controlled Fusion, 55(12):125008, 2013.
J.L. Luxon. A design retrospective of the DIII-D tokamak. Nuclear Fusion, 42(5):614–633, may 2002.
R. L. Miller, M. S. Chu, J. M. Greene, Y. R. Lin-Liu, and R. E. Waltz. Noncircular, finite aspect ratio, local equilibrium model. Phys. Plasmas, 5(4):973–978, 1998.
M. Ottaviani. An alternative approach to field-aligned coordinates for plasma turbulence simulations. Physics Letters A, 375(15):1677–1685, 2011.
Y. Ou, T.C. Luce, E. Schuster, J.R. Ferron, M.L. Walker, C. Xu, and D.A. Humphreys. Towards model-based current profile control at diii-d. Fusion Engineering and Design, 82(5):1153–1160, 2007. Proceedings of the 24th Symposium on Fusion Technology.
S.A. Sabbagh, A.C. Sontag, J.M. Bialek, D.A. Gates, A.H. Glasser, J.E. Menard, W. Zhu, M.G. Bell, R.E. Bell, A. Bondeson, C.E. Bush, J.D. Callen, M.S. Chu, C.C. Hegna, S.M. Kaye, L.L. Lao, B.P. LeBlanc, Y.Q. Liu, R. Maingi, D. Mueller, K.C. Shaing, D. Stutman, K. Tritz, and C. Zhang. Resistive wall stabilized operation in rotating high beta nstx plasmas. Nucl. Fusion, 46(5):635, 2006.
I. Song, C. Choi, and M. Cho. Quench protection system for the superconducting coil of the kstar tokamak. IEEE Transactions on Applied Superconductivity, 17(1):1–6, March 2007.
Youwen Sun. Tokamak geometry and equilibrium. http://theory.ipp.ac.cn/~yj/other/TOKAMAK_geometry_and_equilibrium_Youwen_Sun_201208.pdf, 2012. [Online; Institute of Plasma Physics, Chinese Academy of Sciences].
Joris van der Hoeven. Gnu texmacs, a structured wysiwyw (what you see is what you want) editor for scientists. http://www.texmacs.org/, 2007. [Online].
W. X. Wang, Z. Lin, W. M. Tang, W. W. Lee, S. Ethier, J. L. V. Lewandowski, G. Rewoldt, T. S. Hahm, and J. Manickam. Gyro-kinetic simulation of global turbulent transport properties in tokamak experiments. Physics of Plasmas, 13(9):92505, 2006.
R. B. White and M. S. Chance. Hamiltonian guiding center drift orbit calculation for plasmas of arbitrary cross section. The Physics of Fluids, 27(10):2455–2467, 1984.
B J Xiao, D A Humphreys, and M L Walker. East plasma control system. Fusion Engineering and Design, 83:181, 2008.
B. J. Xiao. The first diverted plasma on east tokamak. 34th EPS Conference on Plasma Phys, 2007.