7 Polarization density expressed as local Fourier expansion

Since δΦ is independent of the velocity in the particle coaordinates, the first term (adiabatic term) in expression (224) is trivial and the velocity integration can be readily performed (assume F0 is Maxwellian), giving

         ∫
δnad  =     q(δΦ )∂F0dv.
            m  ∫  ∂(𝜀     )
      =  -q(δΦ)    − m-fM dv.
         m           T
      =  − qδΦn ,                                 (226)
            T  0
which is called adiabatic response. Next, let us perform the gyro-averaging and the velocity integration of the second term in expression (224), i.e.,
  ∫
−   -q⟨δΦ⟩α∂F0-dv.
    m       ∂𝜀
(227)

7.1 Gyro-averaging of δΦ in guiding-center coordinates

In order to perform the gyro-averaging of δΦ, we Fourier expand δΦ in space as

       ∫
δΦ(x) =   δΦ exp(ik ⋅x)-dk--,
            k         (2π)3
(228)

and then express x in terms of the guiding center variables (X,v) since the gyro-averaging is taken by holding X rather than x constant. The guiding-center transformation gives

                        e∥(X-)-
x = X + ρ(x,v ) ≈ X − v× Ω(X ).
(229)

Using expressions (228) and (229), the gyro-average of δΦ is written as

          ⟨ ∫                  ⟩
⟨δΦ⟩   =      δΦ  exp(ik⋅x )-dk--
    α           k          (2π)3  α
          ⟨ ∫        (   (        e∥(X )))  dk  ⟩
       =      δΦk exp  ik⋅  X − v× -Ω(X-)   (2π)3
          ∫               ⟨   (              ) ⟩ α
       =     δΦk exp(ik ⋅X ) exp  − ik⋅v × e∥(X)    -dk--.       (230)
                                        Ω (X)   α (2π)3
When doing the gyro-averaging, X is hold constant and thus e(X) is also constant. Then it is straightforward to define the gyro-angle α. Let k define one of the perpendicular direction ˆe1, i.e., k = kˆe1. Then another perpendicular basis vector is defined by ˆe2 = e׈e1. Then v is written as v = v(ˆe1 cosα + ˆe2 sinα), which defines the gyro-angle α. Then the blue expression in Eq. (230) is written as
− ik⋅v × e∥(X) =   − ik⋅v⊥ (ˆe1cosα + ˆe2sin α)× e∥(X-)
        Ω (X)                                Ω(X)
               =   − ik⋅-v⊥-(− ˆe2cosα+ ˆe1sinα)
                        Ω(X)
               =   − ik⊥v⊥-sin α.                            (231)
                       Ω
Then the gyro-averaging in expression (230) is written as
⟨   (              ) ⟩       ⟨   (            )⟩
 exp  − ik ⋅v× e∥(X-)    =    exp  − ik⊥v⊥-sin α
               Ω(X)   α                Ω         α
                              1 ∫ 2π   (   k⊥v⊥     )
                         =   2π-    exp  − i-Ω---sinα  dα
                               ( 0   )
                         =   J0  k⊥v⊥- .                       (232)
                                  Ω
where use has been made of the definition of the zeroth Bessel function of the first kind. Then δΦα in expression (230) is written as
       ∫                (      )
⟨δΦ ⟩α =   δΦk exp(ik ⋅X)J0  k⊥v⊥- --dk-.
                           Ω    (2π)3
(233)

7.2 Gryo-angle integration in particle coordinates

Next, we need to perform the integration in velocity space, which is done by holding x (rather than X) constant. Therefore, it is convenient to transform back to particle coordinates. Using X = x + v ×e∥(x)
 Ω(x), expression (233) is written as

        ∫                (k⊥v⊥ )    (       e∥)  dk
⟨δΦ⟩α =   δΦk exp(ik⋅x )J0  --Ω--  exp  ik⋅v × Ω-  (2π-)3.
(234)

Then the velocity integration is written as

∫
  ⟨δΦ ⟩α ∂F0dv
  ∫    ∂𝜀       [∫    (     )                      ]
                        k⊥v⊥-    (       e∥) ∂F0-   -dk--
=   δΦkexp(ik⋅x)    J0   Ω    exp ik ⋅v×  Ω   ∂𝜀 dv (2π)3.    (235)
Similar to Eq. (231), except for now at x rather than X, ik v ×e∥
Ω is written as
       e∥   k⊥v⊥-
ik ⋅v×  Ω = i  Ω  sin α.
(236)

Since this is at x rather than X, k, v, and Ω are different from those appearing in expression (231). However, since this difference is due to the variation of the equilibrium quantity eΩ in a Larmor radius, and thus is small and is ignored in the following.

Plugging expression (236) into expression (235) and using dv = vdvdv, we get

∫
  ⟨δΦ ⟩α∂F0dv
  ∫    ∂ 𝜀       [∫   (     )    (           )              ]
                        k⊥v⊥-       k⊥v⊥-      ∂F0-           -dk--
=   δΦk exp (ik ⋅x )   J0   Ω    exp  i Ω   sinα   ∂𝜀 v⊥dv⊥dv∥dα  (2π)3.(237)
Note that ∂F0∕∂𝜀 is independent of the gyro-angle α in terms of guiding-center variables. When transformed back to particle coordinates, X contained in ∂F0∕∂𝜀 will introduce α dependence via X = x + v ×e∥-
Ω. This dependence on α is weak since the equilibrium quantities can be considered constant over a Larmor radius distance evaluated at the thermal velocity. Therefore this dependence can be ignored when performing the integration over α, i.e., in terms of particle coordinates, ∂F0∕∂𝜀 is approximately independent of the gyro-angle α. Then the integration over α in Eq. (237) can be performed, yielding
    ∫
      ⟨δΦ ⟩α ∂F0dv
           ∂𝜀
    ∫              ∫ ∫   ( k⊥v⊥-)[∫ 2π    ( k⊥v⊥-    )   ] ∂F0        -dk--
=     δΦk exp(ik⋅x )    J0   Ω      0  exp  i Ω   sin α dα   ∂𝜀 v⊥dv⊥dv∥(2π)3
    ∫              [∫ ∫   ( k v  )     (k  v ) ∂F         ]  dk
=     δΦk exp(ik⋅x )     J0  -⊥-⊥- 2πJ0  -⊥-⊥-  --0v⊥dv⊥dv∥  ---3,       (238)
                             Ω            Ω    ∂𝜀           (2π )
where again use has been made of the definition of the Bessel function.

7.2.1 The remaining velocity integration can be performed analytically if F0 is Maxwellian

In order to perform the remaining velocity integration in expression (238), we assume that F0 is a Maxwellian distribution given by

                              (      )
F   =   f  = ----n0(X)-----exp  − mv2-                 (239)
 0       M   (2πT(X )∕m )3∕2      2T(X)
           n0      ( − v2)
    =   (2π)3∕2v3exp  2v2- ,                            (240)
               t       t
where vt = ∘ ----
  T∕m, then
∂F0     m
-∂𝜀-= − T-fM .
(241)

Again we will ignore the weak dependence of n0(X) and T(X) on v introduced by X = x + v × eΩ when transformed back to particle coordinates. (For sufficiently large velocity, the corresponding Larmor radius will be large enough to make the equilibrium undergo substantial variation. Since the velocity integration limit is to infinite, this will definitely occur. However, F0 is exponentially decreasing with velocity, making those particles with velocity much larger than the thermal velocity negligibly few and thus can be neglected.)

Parallel integration Using Eq. (241), the expression in the square brackets of Eq. (238) is written as

  ∫  ∫   (      )
2π     J2  k⊥v⊥-  ∂F0v dv dv
        0   Ω     ∂𝜀  ⊥  ⊥  ∥
    m   n   ∫  ∫   ( k v  ) 1    (  v2∥ + v⊥2)
= − -----01∕2     J20  -⊥-⊥-  -3exp  −----2--  v⊥dv⊥dv∥       (242)
    T (2π)            Ω     vt  (      2vt)
    m   n0  ∫  ∫  2( k⊥v⊥ )       v2∥ + v2⊥ -   -  -
= − T-(2π)1∕2     J0  -Ω--- exp  − ---2--- v⊥d v⊥dv∥,        (243)
where v = v∕vt, v = v∕vt. Using
∫ ∞    (    2)     √ ---
    exp  − x- dx =   2π,
 −∞        2
(244)

the integration over v in expression (243) can be performed, yielding

      ∫ ∞    (     )    (  -2 )
− m-n0    J20  k⊥v⊥-  exp − v⊥- v⊥dv ⊥
  T    0        Ω           2
(245)

Perpendicular integration Using (I verified this by using Sympy)

∫            (    )
  ∞  2          x2              2    2
 0  J0(ax)exp  − 2  xdx = exp(− a )I0(a ),
(246)

where I0(a) is the zeroth modified Bessel function of the first kind, expression (245) is written

− m-n0exp(− b)I0(b)
  T
(247)

where b = k2vt2Ω2 = k2ρt2. Then the corresponding density (227) is written as

    ∫                  ∫
−-q   ⟨δΦ ⟩α ∂F0dv = qn0   δΦkexp(ik⋅x)exp(− b)I0(b)-dk-.
 m         ∂𝜀       T                           (2π)3
(248)

7.2.2 Final form of polarization density

In Fourier space, the adiabatic term in expression (226) is written as

∫                              ∫
  -q(δΦ)∂F0-dv = − qn0δΦ = − qn0  δΦ  exp(ik ⋅x)-dk--.
  m      ∂𝜀        T         T      k         (2π)3
(249)

Plugging expression (248) and (249) into expression (224), the polarization density np is written as

            ∫
np  =  − qn0   δΦkexp(ik ⋅x)[1 − exp(− b)I0(b)]-dk--.           (250)
          T                               (2π)3
Define
Γ 0 = exp(− b)I0(b),
(251)

then Eq. (250) is written as

         ∫
n = − qn0   δΦ  exp(ik ⋅x)[1 − Γ ]-dk-,
 p     T      k              0(2π)3
(252)

Expression (252) agrees with the result given in Yang Chen’s notes. Note that the dependence on species mass enters the formula through the Larmor radius ρt in Γ0.

7.3 Pade approximation

Γ0 defined in Eq. (251) can be approximated by the Pade approximation as

     -1---
Γ 0 ≈ 1+ b.
(253)

The comparison between the exact value of Γ0 and the above Pade approximation is shown in Fig. 1.


pict

Figure 1: Comparison between the exact value of Γ0 = exp((kρ)2)I0((kρ)2) and the corresponding Pade approximation 1(1 + (kρ)2).

Using the Pade approximation (253), the polarization density np in expression (252) can be written as

             ∫
n   ≈   − qn0  δΦ  exp(ik ⋅x)-k2⊥-ρ2---dk-.               (254)
  p       T      k          1+ k2⊥ρ2 (2π)3
(Pade approximate is the “best” approximation of a function by a rational function of given order – under this technique, the approximant’s power series agrees with the power series of the function it is approximating.)
7.3.1 Long wavelength approximation of the polarization density

In the long wavelength limit, kρ 1, expression (254) can be further approximated as

             ∫
n   ≈  − qn0   δΦ exp(ik⋅x)k2ρ2 -dk-,
 p        T      k          ⊥   (2π)3
    =   qn0ρ2∇2 δΦ.                                   (255)
        T     ⊥
Then the corresponding term in the Poisson equation is written as
-qn   =   q2n0ρ2∇2 δΦ
𝜀0  p     𝜀0T    ⊥
          ρ2- 2
      =   λ2 ∇⊥ δΦ,                            (256)
           D
where λD is the Debye length defined by λD2 = T𝜀0(n0q2). For typical tokamak plasmas, the thermal ion gyroradius ρi is much larger than λD. Therefore the term in expression (256) for ions is much larger than the space charge term 2δΦ ≡∇2δΦ + 2δΦ ≈∇2δΦ in the Poisson equation. Therefore the space charge term can be neglected in the long wavelength limit.

Equation (256) also shows that electron polarization density is smaller than the ion polarization density by a factor of ρe∕ρi 160. Note that this conclusion is drawn in the long wavelength limit. For short wavelength, the electron polarization and ion polarization density can be of similar magnitude (to be discussed later).

7.3.2 Polarization density expressed in terms of Laplacian operator

The polarization density expression (255) is for the long wavelength limit, which partially neglects FLR effect. Let us go back to the more general expression (254). The Poisson equation is written

     2
− 𝜀0∇⊥δΦ = qiδni + qeδne.
(257)

Write δni = npi + δni, where δnpi is the ion polarization density, then the above expression is written

− 𝜀0∇2⊥δΦ − qinpi = qiδn′i + qeδne.
(258)

Fourier transforming in space, the above equation is written

− 𝜀0k2⊥δˆΦ − qiˆnpi = qiδˆn′i + qeδˆne,
(259)

where ˆnpi is the Fourier transformation (in space) of the polarization density npi and similar meanings for δˆΦ, δˆni, and δˆne. Expression (254) implies that ˆnpi is given by

                2 2
ˆnpi = − qini0δˆΦ-k⊥ρi2-2.
        Ti   1 + k⊥ρi
(260)

Using this, equation (259) is written

            (         2  2    )
− 𝜀0k2⊥ δˆΦ− qi − qini0--k⊥-ρ2i2δˆΦ  =  qiδˆn′i + qeδˆne,
                Ti  1+ k⊥ρi
(261)

Multiplying both sides by (1 + k2ρi2)∕𝜀0, the above equation is written

                    (              )
− (1 + k2⊥ρ2i)k2⊥ δˆΦ−  qi − qini0(k2⊥ρ2i)δˆΦ  = -1(1+ k2⊥ ρ2i)(qiδˆn′i + qeδˆne).
                  𝜀0     Ti             𝜀0
(262)

Next, transforming the above equation back to the real space, we obtain

                     (            )
− (1− ρ2i∇2⊥ )∇2⊥δΦ − qi  qini0ρ2i∇2⊥δΦ  = -1(1− ρ2i∇2⊥)(qiδn′i + qeδne).
                   𝜀0   Ti            𝜀0
(263)

Neglecting the Debye shielding term, the above equation is written

  (  2      )
−   ρi2-∇2⊥δΦ  = -1(1− ρ2i∇2⊥)(qiδn′i + qeδne),
    λDi         𝜀0
(264)

which is the equation actually solved in many gyrokinetic codes, where λDi2 = 𝜀0Ti(qi2ni0).