Download as pdf or txt
Download as pdf or txt
You are on page 1of 28

Superconductor in a weak static gravitational field

Giovanni Alberto Ummarino


Politecnico di Torino, Dipartimento DISAT
National Research Nuclear University MEPhI - Moscow Engineering Physics Institute
giovanni.ummarino@polito.it

Antonio Gallerati
Politecnico di Torino, Dipartimento DISAT
Istituto Nazionale di Fisica Nucleare (INFN), Sezione di Torino, Italy
antonio.gallerati@polito.it
arXiv:1710.01267v1 [gr-qc] 2 Oct 2017

Abstract

We provide the detailed calculation of a general form for Maxwell and Lon-
don equations that takes into account gravitational corrections in linear
approximation. We determine the possible alteration of a static gravita-
tional field in a superconductor making use of the time-dependent Ginzburg-
Landau equations, providing also an analytic solution in the weak field con-
dition. Finally, we compare the behavior of a high-Tc superconductor with
a classical low-Tc superconductor, analyzing the values of the parameters
that can enhance the reduction of the gravitational field.

Contents
1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2

2 Weak field approximation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4


2.1 Generalizing Maxwell equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
2.2 Generalizing London equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9

3 Isotropic superconductor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
3.1 Time-dependent Ginzburg-Landau equations . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
3.2 Solving dimensionless TDGL . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
3.3 YBCO vs. Pb . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18

4 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22

A Sign convention . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25

1
1 Introduction
There is no doubt that the interplay between the theory of the gravitational
field and superconductivity is a very intriguing field of research, whose theoreti-
cal study has been involving many researchers for a long time [1–19]. Podkletnov
and Nieminem declared the achievement of experimental evidence for a gravi-
tational shielding in a high-Tc superconductor (HTCS) [20, 21]. After their an-
nouncement, other groups tried to repeat the experiment obtaining controversial
results [22–24], so that the question is still open.
In 1996, G. Modanese interpreted the results by Podkletnov and Nieminem in
the frame of the quantum theory of General Relativity [25, 26] but the complexity
of the formalism makes it very difficult to extract quantitative predictions. Af-
terwards, Agop et al. wrote generalized Maxwell equations that simultaneously
treat weak gravitational and electromagnetic fields [27, 28].

Superfluid coupled to gravity. It is well known that, in general, the gravita-


tional force is not influenced by any dielectric-type effect involving the medium.
In the classical case, this is due to the absence of a relevant number of charges
having opposite sign which, redistributing inside the medium, might counteract
the applied field. On the other side, if we regard the medium as a quantum sys-
tem, the probability of a graviton excitation of a medium particle is suppressed,
due to the smallness of gravitational coupling. This means that any kind of
shielding due to the presence of the medium can only be the result of an inter-
action with a different state of matter, like a Bose condensate or a more general
superfluid.
The nature of the involved field is also relevant for the physical process. If
the gravitational field itself is considered as classical, it is readily realized that no
experimental device – like the massive superconducting disk of the Podkletnov
experiment [20, 21] – can influence the local geometry so much as to modify
the measured sample weight. This means that the hypothetical shielding effect
should consist of some kind of modification (or “absorption”) of the field in the
superconducting disk.
Since the classical picture is excluded, we need a quantum field description for
the gravitational interaction [25, 26]. In perturbation theory the metric gµν (x)
is expanded in the standard way [29]

gµν (x) = ηµν + hµν (x) (1)

as the sum of the flat background ηµν plus small fluctuations encoded in the
hµν (x) component. The Cooper pairs inside the superconducting sample com-
pose the Bose condensate, described by a bosonic field φ with non-vanishing
vacuum expectation value φ0 = h0|φ|0i.

2
The Einstein-Hilbert Lagrangian has the standard form1

1
Leh = (R − 2 Λ) , (2)
8πGn

where R is the Ricci scalar and Λ is the cosmological constant. The part of the
Lagrangian describing the bosonic field φ coupled to gravity has the form:

1 µν 1
Lφ = − g ∂µ φ∗ ∂ν φ + m2 φ∗ φ (3)
2 2
where m is the mass of the Cooper pair [25].
If we expand the bosonic field as φ = φ0 + φ̄, one can consider the v.e.v.
φ0 as an external source, related to the structure of the sample and external
electromagnetic fields, while the φ̄ component can be included in the integra-
tion variables. The terms including the φ̄ components are related to graviton
emission-absorption processes (which we know to be irrelevant) and can safely be
neglected in Lφ . Perturbatively, the interaction processes involving the metric
and the condensate are of the form

Lint ∝ hµν ∂µ φ0 ∗ ∂ν φ0 (4)

and give rise to (gravitational) propagator corrections, which are again irrelevant.
The total Lagrangian L = Leh + Lφ contains a further coupling between
gµν and φ0 , which turns out to be a contribution to the so-called intrinsic cosmo-
logical term given by Λ. Explicitly, the total Lagrangian can in fact be rewritten
as
1
L = Leh + Lφ = (R − 2 Λ) + Lint + L0 + L¯φ̄ , (5)
8πGn
where L¯φ̄ are the negligible contributions having at least one field φ̄ and where

1 1
L0 = − ∂µ φ0 ∗ ∂ µ φ0 + m2 |φ0 |2 , (6)
2 2
that is, a Bose condensate contribution to the total effective cosmological term.
This may produce slightly localized “instabilities” and thus an observable effect,
in spite of the smallness of the gravitational coupling (4).
The above instabilities can be found in the superconductor regions where the
condensate density is larger: in these regions, the gravitational field would tend to
assume fixed values due to some physical cutoff, that prevents arbitrary growth.
The mechanism is similar to classical electrostatics in perfect conductors, where
the electric field is constrained to be globally zero within the sample. In the
latter case, the physical constraint’s origin is different (and is due to a charge
redistribution), but in both cases the effect on field propagation and on static
potential turns out to be a kind of partial shielding.
1
we work in the “mostly plus” framework, η = diag(−1, +1, +1, +1), and set c = ~ = 1

3
In accordance with the framework previously exposed, the superfluid density
φ0 (x) is determined not only by the internal microscopic structure of the sample,
but also by the same magnetic fields responsible for the Meissner effect and the
currents in the superconductor. The high-frequency components of the magnetic
field can also provide energy for the above gravitational field modification [25].
The previous calculation shows how Modanese was able to demonstrate, in
principle, how a superfluid can determine a gravitational shielding effect. In
Sect. 3 we will quantify this effect by following a different approach, as the
Ginzburg-Landau theory for a superfluid in an external gravitational field.

2 Weak field approximation


Now we consider a nearly flat spacetime configuration, i.e. an approximation
where the gravitational field is weak and where we shall assume eq. (1), that is,
the metric gµν can be expanded as:

gµν ' ηµν + hµν , (7)

where the symmetric tensor hµν is a small perturbation of the flat Minkowski
metric in the mostly plus convention2 , ηµν = diag(−1, +1, +1, +1). The inverse
metric in the linear approximation is given by

g µν ' η µν − hµν . (8)

2.1 Generalizing Maxwell equations


If we consider an inertial coordinate system, to linear order in hµν the connection
is written as
1
Γλµν ' η λρ (∂µ hνρ + ∂ν hρµ − ∂ρ hµν ) , (9)
2
The Ricci tensor (Appendix A) is given by the contraction of the Riemann tensor

Rµν = Rσµσν . (10)

and, to linear order in hµν , it reads

Rµν ' ∂λ Γλµν + ∂µ Γλλν +  ΓΓ − 


Γ Γ =
1 1 1
= (∂µ ∂ ρ hνρ + ∂ν ∂ ρ hµρ ) − ∂ρ ∂ ρ hµν − ∂µ ∂ν h = (11)
2 2 2
ρ 1 2 1
= ∂ ∂(µ hν)ρ − ∂ hµν − ∂µ ∂ν h ,
2 2

where we have used eq. (9) and where h = hσσ .


2
see Appendix A for definitions and sign conventions

4
The Einstein equations have the form [29, 30]:

(E) 1
Gµν = Rµν − gµν R = 8πGn Tµν , (12)
2
and the term with the Ricci scalar R = g µν Rµν can be rewritten, in first-order
approximation and using eq. (11), as

1 1 1  
gµν R ' ηµν η ρσ Rρσ = ηµν ∂ ρ ∂ σ hρσ − ∂ 2 h , (13)
2 2 2
so that the l.h.s. of the Einstein equations in weak field approximation reads

(E) 1
Gµν = Rµν − gµν R '
2 (14)
1 1 1  
' ∂ ρ ∂(µ hν)ρ − ∂ 2 hµν − ∂µ ∂ν h − ηµν ∂ ρ ∂ σ hρσ − ∂ 2 h .
2 2 2
If one introduces the symmetric tensor

1
h̄µν = hµν − ηµν h , (15)
2
the above expression can be rewritten as

(E) 1 2 1
Gµν ' ∂ ρ ∂(µ h̄ν)ρ − ∂ h̄µν − ηµν ∂ ρ ∂ σ h̄ρσ =
2 2
= ∂ ρ ∂[ν h̄ρ]µ + ∂ ρ ∂ σ ηµ[σ h̄ν]ρ = (16)
 
= ∂ ρ ∂[ν h̄ρ]µ + ∂ σ ηµ[ρ h̄ν]σ ,

where we have exchanged dumb indices in the last term of the second line.
If we now define the tensor

Gµνρ ≡ ∂[ν h̄ρ]µ + ∂ σ ηµ[ρ h̄ν]σ , (17)

whose structure implies the property

Gµνρ = − Gµρν , (18)

the Einstein equations can be rewritten in the compact form:

(E)
Gµν = ∂ ρ Gµνρ = 8πGn Tµν . (19)

We can impose a gauge fixing making use of the harmonic coordinate condition,
expressed by the relation [29]:

−g g µν = 0 2xµ = 0 ,

∂µ ⇔ (20)

5
where g ≡ det [gµν ], and that can be rewritten in the form

g µν Γλµν = 0 , (21)

also known as De Donder gauge.


Imposing the above condition and using eqs. (7) and (9), in first-order ap-
proximation we find:

1 µν λρ 1
0 ' η η (∂µ hνρ + ∂ν hρµ − ∂ρ hµν ) = ∂µ hµν − ∂ ν h , (22)
2 2
that is, we have the condition

1 ν 1
∂µ hµν ' ∂ h ⇔ ∂ µ hµν ' ∂ν h . (23)
2 2
Now, one also has

1 1
 
µ µ
∂ hµν = ∂ h̄µν + ηµν h = ∂ µ h̄µν + ∂ν h , (24)
2 2

and, using eq. (23), we find the so-called Lorenz gauge condition:

∂ µ h̄µν ' 0 . (25)

The above relation further simplifies expression (17) for Gµνρ , which takes the
very simple form
Gµνρ ' ∂[ν h̄ρ]µ , (26)

and verifies also the relation

∂[λ| G0|µν] = 0 ⇒ G0µν ∝ ∂µ Aν − ∂ν Aµ , (27)

which implies the existence of a potential.

Gravito-Maxwell equations. Now, let us define the fields3 [27]


1 1
Eg ≡ Ei = − G00i = − ∂[0 h̄i]0 , (28.i)
2 2
1
Ag ≡ Ai = h̄0i , (28.ii)
4
1
Bg ≡ Bi = εi jk G0jk , (28.iii)
4

where obviously i = 1, 2, 3 and

1 
G0ij = ∂[i h̄j]0 = ∂i h̄j0 − ∂j h̄i0 = 4 ∂[i Aj] . (29)
2
3
for the sake of simplicity, we initially set the physical charge e = m = 1

6
One can immediately see that

1 jk
Bg = εi 4 ∂[j Ak] = εi jk ∂j Ak = ∇ × Ag ,
4
(30)
=⇒ ∇ · Bg = 0 .

Then one also has


G00i T00
∇ · Eg = ∂ i Ei = − ∂ i = − 8πGn = 4πGn ρg , (31)
2 2
using eq. (19) and having defined ρg ≡ −T00 .

If we consider the curl of Eg , we obtain

G00k 1
∇ × Eg = εi jk ∂j Ek = − εi jk ∂j = − εi jk ∂j ∂[0 h̄k]0 =
2 2
(32)
1 ∂Bg
= − 4 ∂0 εi jk ∂j Ak = − ∂0 Bi = − .
4 ∂t
Finally, one finds for the curl of Bg

1 jk `m
∇ × Bg = εi jk ∂j Bk = εi εk ∂j G0`m =
4
1  ` jm  1
= δi δ − δi m δ j` ∂j G0`m = ∂ j G0ij =
4 2
1 µ 1
= (∂ G0iµ + ∂0 G0i0 ) = (∂ µ G0iµ − ∂0 G00i ) = (33)
2 2
1 ∂Ei
= (8πGn T0i − ∂0 G00i ) = 4πGn ji + =
2 ∂t
∂Eg
= 4πGn jg + ,
∂t

using again eq. (19) and having defined jg ≡ ji ≡ T0i .


Summarizing, once defined the fields of (28) and having restored physical
units, one gets the field equations:

m2 ρg
∇ · Eg = 4πGn 2
ρg = ;
e εg

∇ · Bg = 0 ;
(34)
∂Bg
∇ × Eg =− ;
∂t
m2 1 ∂Eg 1 ∂Eg
∇ × Bg = 4πGn 2 2
jg + 2 = µg jg + 2 ,
c e c ∂t c ∂t
formally equivalent to Maxwell equations, where Eg and Bg are the gravito-

7
electric and gravitomagnetic field, respectively, and where we have defined the
vacuum gravitational permittivity

1 e2
εg = , (35)
4πGn m2

and the vacuum gravitational permeability

m2
µg = 4πGn . (36)
c2 e2
For example, on the Earth surface, Eg is simply the Newtonian gravitational
acceleration and the Bg field is related to angular momentum interactions [27,
28, 31–33]. The mass current density vector jg can also be expressed as:

jg = ρg v , (37)

where v is the velocity and ρg is the mass density.

Gravito-Lorentz force. Let us consider the geodesic equation for a particle


in the field of a weakly gravitating object:

d2 xλ λ dxµ dxν
+ Γ µν = 0. (38)
ds2 ds ds
If we consider a particle in non-relativistic motion, the velocity of the particle
i vi v j
becomes vci ' dx dt . If we also neglect terms in the form c2 and limit ourselves
to static fields (∂t gµν = 0), it can easily be verified that a geodesic equation for
a particle in non-relativistic motion can be written as:[34, 35]

dv
= E g + v × Bg , (39)
dt
which shows that the free fall of the particle is driven by the analogous of a
Lorentz force produced by the gravito-Maxwell fields.

Generalized Maxwell equations. It is possible to define the generalized elec-


tric/magnetic field, scalar and vector potentials containing both electromagnetic
and gravitational term as
m m m m
E = Ee + Eg ; B = Be + Bg ; φ = φe + φg ; A = Ae + Ag , (40)
e e e e
where m and e are the mass and electronic charge, respectively, and the sub-
scripts identify the electromagnetic and gravitational contributions.

8
The generalized Maxwell equations then become
!
1 1
∇·E = + ρ;
εg ε 0

∇·B = 0;
(41)
∂B
∇×E =− ;
∂t
1 ∂E
∇ × B = (µg + µ0 ) j + ,
c2 ∂t
where we have set
e
ρg = ρ,
m
(42)
e
jg = j,
m
and where ε0 and µ0 are the electric permittivity and magnetic permeability in
the vacuum.

2.2 Generalizing London equations


The London equations for a superfluid in stationary state read [36–38]:

m ∂j
Ee = ; (43.i)
ns e2 ∂t
m
Be =− ∇×j. (43.ii)
ns e2

where j = ns e vs is the supercurrent and ns is the superelectron density. If


we also consider Ampère’s law for a superconductor in stationary state (no dis-
placement current)
∇ × Be = µ0 j , (44)

from (43.ii) and using vector calculus identities, we obtain

) − ∇2 B = µ ∇ × j = −µ ns e 2
∇ × ∇ × Be = ∇ (
∇·B

e e 0 0 Be , (45)
m
that is,
1
∇2 Be = Be , (46)
λ2e
where we have introduced the penetration depth

m
r
λe = . (47)
µ0 ns e2

9
Using the vector potential Ae , the two London equations (43) can be summarized
in the (not gauge-invariant) form

1 
j =− Ae Be = ∇ × A e . (48)
µ0 λ2e

Generalized London equations. If we now take into account gravitational


corrections, we should consider for the fields and the vector potential the gener-
alized form of definition (40):
m m
B = Be + Bg , A = Ae + Ag , B=∇×A. (49)
e e
If A is minimally coupled to the wave function

ψ = ψ0 eiϕ , ψ02 ≡ |ψ|2 = ns , (50)

the second London equation can be derived from a quantum mechanical current
density
i  ∗˜ 
˜ ∗ ,
j =− ψ ∇ψ − ψ ∇ψ (51)
2m
where ∇˜ is the covariant derivative for the minimal coupling:

˜ = ∇ − i g̃ A ,
∇ (52)

so that one has for the current


i g̃ 1
j = − (ψ ∗ ∇ψ − ψ∇ψ ∗ ) − A |ψ|2 = |ψ|2 (∇ϕ − g̃ A) . (53)
2m m m
If we now take the curl of the previous equation, we find

m 1
B = − 2
∇×j = − ∇×j, (54)
g̃ |ψ| ζ

which is the generalized form of the second London equation (43.ii).

To find an explicit expression for ζ, we consider the case Bg = 0 obtaining

m 1
B = Be + g =−
B
Z ∇×j, (55)
e Z ζ

and, using (43.ii), (47) and (50), we find

1
g̃ = e2 , = µ0 λ2e . (56)
ζ

Then we consider the case Be = 0, so that we have


m m
B = Ze+
B Bg = −µ0 λ2e ∇ × j = −µ0 λ2e ∇ × jg , (57)
Z e e

10
together with gravito-Ampère’s law (34) in stationary state,

∇ × Bg = µg jg , (58)

so that, taking the curl of the above equation, we find

1 1
∇ × ∇ × Bg = −∇2 Bg = µg ∇ × jg = − µg Bg = − 2 Bg , (59)
µ0 λ2e λg

where we have introduced the penetration depth


s s
µ0 λ2e c2
λg = = . (60)
µg 4πGn m ns

Finally, using the stationary generalized Ampère’s law from (41) and using eq.
(60) we find !
λ2e
∇ × B = (µ0 + µg ) j = µ0 1 + 2 j , (61)
λg
and taking the curl we obtain the general form
! !
λ2 1 λ2
2
∇ B = −µ0 1 + 2e ∇ × j = µ0 1 + e2 B =
λg µ0 λ2e λg
! (62)
1 1 1
= 2
+ 2 B = 2 B,
λe λg λ

where we have defined a generalized penetration depth λ :

λg λe λg
 
λ = q ' λe ' 1021 . (63)
λ2g + λ2e λe

The general form of eq. (48) is



j = −ζ A B = ∇×A . (64)

and, since charge-conservation requires the condition ∇ · j = 0, we obtain for the


vector potential
∇·A=0,

that is, the so-called Coulomb gauge (or London gauge).

11
3 Isotropic superconductor
In Sect. 1 we have shown how Modanese was able to theoretically describe the
gravitational shielding effect due to the presence of a superfluid. Now we are
going to study the same problem with a different approach.
Modanese has solved gravitational field equation where the contribution of
the superfluid was encoded in the energy-momentum tensor. In the following,
we are going to solve the Ginzburg-Landau equation for the superfluid order
parameter in an external gravitational field.
Let us restrict ourselves to the case of an isotropic superconductor in the
gravitational field of the earth and in absence of an electromagnetic field, we can
take Ee = 0 and Be = 0. Moreover, Bg in the solar system is very small [39, 40],
therefore E = m e Eg and B = 0. Finally, we also have the relations φ = e φg
m
m
and A = e Ag , so we can write down our set of conditions:
m
Ee = 0 , Be = 0 , Bg = 0 =⇒ E= Eg , B=0; (65.i)
e

together with
m m
φ= φg , A= Ag . (65.ii)
e e

The situation is not the same as the Meissner effect but, rather, as the case of a
superconductor in an electric field.

3.1 Time-dependent Ginzburg-Landau equations


Since the gravitoelectric field is formally analogous to an electric field we can use
the time-dependent Ginzburg-Landau equations (TDGL) which, in the Coulomb
gauge ∇ · A = 0 are written in the form [41–47]:
2
~2 ∂ 2ie 1 2e
  
2
+ φ ψ − a ψ + b |ψ| ψ + i~∇ + A ψ = 0,
2mD ∂t ~ 2m c
(66.i)
!
4π σ ∂A i~ e ∗ 4 e2
∇×∇×A−∇×H =− + σ ∇φ + (ψ ∇ψ − ψ∇ψ ∗ ) + |ψ|2 A ,
c c ∂t m mc
(66.ii)
where D is the diffusion coefficient, σ is the conductivity in the normal phase, H
is the applied field and the vector field A is minimally coupled to ψ. The above
TDGL equations for the variables ψ, A are derived minimizing the total Gibbs
free energy of the system [36–38].

12
The coefficients a and b in (66.i) have the following form:

a = a(T ) = a0 (T − Tc ) ,
(67)
b = b(T ) ≡ b(Tc ) ,

a0 , b being positive constants and Tc the critical temperature of the supercon-


ductor. The boundary and initial conditions are

2e
  
i~ ∇ψ + Aψ · n = 0 

c

 
ψ(x, 0) = ψ0 (x) 


∇ × A · n = H · n  on ∂Ω × (0, t) ; on Ω


 A(x, 0) = A0 (x) 

A·n=0 
(68)

where ∂Ω is the boundary of a smooth and simply connected domain in RN .

Dimensionless TDGL. In order to write eqs. (66) in a dimensionless form,


the following quantities can be introduced:
s
2 |a(T )| h b m c2
Ψ (T ) = , ξ(T ) = p , λ(T ) = ,
b 2 m |a(T )| 4π |a(T )| e2
(69.i)
s
4π µ0 |a(T )|2 h
Hc (T ) = = √ , (69.ii)
b 4 e 2π λ(T ) ξ(T )

λ(T ) λ2 (T ) 4π σ D
κ = , τ (T ) = , η = , (69.iii)
ξ(T ) D ε0 c2

where λ(T ), ξ(T ) and Hc (T ) are the penetration depth, coherence length and
thermodynamic field, respectively. The dimensionless quantities are then defined
as:
x t ψ
x0 = , t0 = , ψ0 = , (70.i)
λ τ Ψ

and the dimensionless fields are written

Aκ φκ Hκ
A0 = √ , φ0 = √ , H0 = √ . (70.ii)
2 Hc λ 2 Hc D 2 Hc

Inserting eqs. (70) in eqs. (66) and dropping the prime gives the dimensionless
TDGL equations in a bounded, smooth and simply connected domain in RN
[41, 42]:

∂ψ  
+ i φ ψ + κ2 |ψ|2 − 1 ψ + (i ∇ + A)2 ψ = 0 , (71.i)
∂t

13
∂A i
 
∇ × ∇ × A − ∇ × H = −η + ∇φ − (ψ ∗ ∇ψ − ψ∇ψ ∗ ) − |ψ|2 A ,
∂t 2
(71.ii)

and the boundary and initial conditions (68) become, in the dimensionless form

(i ∇ψ + A ψ) · n = 0 
 
ψ(x, 0) = ψ0 (x) 
 
∇×A·n=H·n on ∂Ω × (0, t) ; on Ω .

 A(x, 0) = A0 (x) 
A·n=0 

(72)

3.2 Solving dimensionless TDGL


If the superconductor is on the Earth’s surface, the gravitational field is very
weak and approximately constant. This means that one can write

φ = −g? x , (73)

with
λ(T ) κ m g
g? = √  1, (74)
2 e Hc (T ) D

g being the acceleration of gravity. The corrections to φ in the superconductor


are of second order in g? and therefore they are not considered here.

Now we search for a solution of the form

ψ(x, t) = ψ0 (x, t) + g? γ(x, t) ,


A(x, t) = g? β(x, t) , (75)
φ(x) = −g? x .

At order zero in g? , eq. (71.i) gives

∂ψ0 (x, t)   ∂ 2 ψ0 (x, t)


+ κ2 |ψ0 (x, t)|2 − 1 ψ0 (x, t) − = 0, (76)
∂t ∂x2
with the conditions

ψ0 (x, 0) = 0 ,
ψ0 (0, t) = 0 , (77)
ψ0 (L, t) = 0 ,

where L is the length of the superconductor, here in units of λ, and t = 0 is the


instant in which the material undergoes the transition to the superconducting
state.

14
The static classical solution of eq. (76) is

κx κ (x − L)
   
ψ0 (x, t) ≡ ψ0 (x) = tanh √ tanh √ , (78)
2 2

and, from (71.i), one obtains

∂γ(x, t) ∂ 2 γ(x, t) 2

2

− + κ 3 |ψ 0 (x)| − 1 γ(x, t) = i x ψ0 (x) (79)
∂t ∂x2
at first-order in g? , with the conditions

γ(x, 0) = 0 ,
γ(0, t) = 0 , (80)
γ(L, t) = 0 .

The first-order equation for the vector potential is written

∂β(x, t)
η + |ψ0 (x)|2 β(x, t) + J(x, t) − η = 0 , (81)
∂t
with the constraint
β(x, 0) = 0 . (82)

The second-order spatial derivative of β does not appear in eq. (81): this is due
to the fact that, in one dimension, one has


∇2 A = ∇·A, (83)
∂x
and therefore, in the Coulomb gauge

∇ × ∇ × A = ∇ (∇ · A) − ∇2 A = 0 . (84)

The quantity J(x, t) that appears in eq. (81) is given by

1 ∂ ∂
 
J(x, t) = ψ0 (x) Im [γ(x, t)] − Im [γ(x, t)] ψ0 , (85)
2 ∂x ∂x

and the solution of eq. (81) is

e−P(x) t
Z t
1  
β(x, t) = 1 − e−P(x) t − dt J(x, t) eP(x) t , (86)
P(x) η 0

with

|ψ0 (x)|2
P(x) = . (87)
η

Now, we have the form (78) for ψ0 (x, t) and also the above (86) for β(x, t) as
a function of γ(x, t) through the definition of J(x, t): the latter can be used in

15
(75) to obtain both ψ(x, t) and A(x, t) as functions of γ(x, t).
The gravitoelectric field can be found using the relation

∂A
Eg = − ∇φ − , (88)
∂t
and its explicit form reads
!
e−P(x) t
Z t
1 ∂
Eg (x, t) = 1 − e−P(x) t − dt J(x, t) e P(x) t
. (89)
g? ∂t η 0

The above formula shows that, for maximizing the effect of the reduction of the
gravitational field in a superconductor, it is necessary to reduce η and have large
spatial derivatives of ψ0 (x) and γ(x, t). The condition for a small value of η
is a large normal-state resistivity for the superconductor and a small diffusion
coefficient
vf `
D ∼ , (90)
3
where vf is the Fermi velocity (which is small in HTCS) and ` is the mean free
path: this means that the effect is enhanced in "bad" samples with impurities,
not in single crystals.
If we consider the case J(x, t) = 0, given by the condition

ψ0 (x) = Im [γ(x, t)] ≡ Im [γ(x)] , (91)

we obtain the simplified equation

∂β(x, t)
η + |ψ0 (x)|2 β(x, t) − η = 0 , (92)
∂t
which is solved, together with the constraint (82), by the function
2
− |ψ0 (x)|
!
η η t
β(x, t) = 1−e . (93)
|ψ0 (x)|2

Using then eqs. (88) and (75) we find


2
Eg − |ψ0 (x)|
η t
= 1−e . (94)
g?

The above equation shows that, unlike the general case, in the absence of the
contribution of J(x, t) the effect is bigger than in the case of single crystal low-
Tc superconductor, where η is large.

3.2.1 Approximate solution


From the experimental viewpoint, the greater are the length and time scales
over which there is a variation of Eg , the easier is the observation of this effect.

16
Actually, we started from dimensionless equations and therefore the length and
time scales are determined by λ(T ) and τ (T ) of eqs. (69), which should therefore
be as large as possible. In this sense, materials having very large λ(T ) could
be interesting for the study of this effect [48]. Moreover, eq. (89) shows the
dependence of relaxation with respect to |ψ0 (x)|2 through the definition of P(x):
one can see that |ψ0 (x)| must be as small as possible and this implies that also
κ must be small, see eq. (78). This also means that λ(T ) and ξ(T ) must both
be large.
Up to now we have dealt with the expression of β(x, t) as a function of
γ(x, t). If we want to obtain an explicit expression for Eg , we have to solve
the equation (79) for γ(x, t): this is a difficult task which can be undertaken
only in a numerical way. Nevertheless, if one puts ψ0 (x) ≈ 1, which is a good
approximation in the case of YBa2 Cu3 O7 (YBCO) in which κ = 94.4, one can
find the simple approximate solution:

2
Qn sin (ωn x) e−Cn t ,
X
γ(x, t) = i γ0 (x) + i (95)
n=1

with
 !
x α α α
 
γ0 (x) = 1 − cosh − x sech , (96.i)
2 κ2 2 L 2

(1) (2)
!
1 L
(−1)n 1 Q +Q
Z
Qn = dx γ0 (x) sin(ωn x) = − ωn n 2 n , (96.ii)
L 0 2 κ2 ωn Cn
and

Cn2 = ωn2 + 2 κ2 , ωn = n π/L , α = 2κL , (96.iii)

(1) 2 α ωn2
Qn = (−1)n − cosh α + sinh α , (96.iv)
L Cn2
!
(2) 2 α (−1)n − cosh α
Qn = (cosh α − 1) 1 + 2 2 . (96.v)
L Cn sinh α

Taking into account eq. (78) and inserting eq. (95) in eq. (85) and then in eq. (86),
we can find a new expression for the gravitoelectric field Eg :

1 J0 (x) 1 X
 
s Eg (x, t) = 1 − e−P(x) t 1− + Qn Rn (x) Sn (x, t) , (97)
g? η η n=1

where
1 ∂ ∂
 
J0 (x) = ψ0 (x) γ0 (x) − γ0 (x) ψ0 (x) , (98.i)
2 κ2 ∂x ∂x

17

Rn (x) = ωn ψ0 (x) cos(ωn x) − sin(ωn x) ψ0 (x) , (98.ii)
∂x
2
C 2 e−Cn t − P(x) e−P(x) t
Sn (x, t) = n . (98.iii)
P(x) − Cn2

By making the approximation

ix
γ(x) ' , (99)
2 κ2
one finds the result
1 J00 (x)
 
Eg (x, t) = 1 − e−P(x) t 1− , (100)
g? η

where
1 ∂
 
J00 (x) = ψ0 (x) − x ψ0 (x) . (101)
2 κ2 ∂x

In spite of its crudeness, in the case of YBCO the above approximate solu-
tion (100) gives the same results of the solution (97). Moreover, nothing changes
significantly if one neglects the finite size of the superconductor and uses
 √ 
ψ0 (x) = tanh κx/ 2 (102)

instead of eq. (78).

3.3 YBCO vs. Pb


In the case of YBCO, the variation of the gravitoelectric field Eg in time and
space is shown in Figs. 1 and 2 . It is easily seen that this effect is almost
independent on the spatial coordinate.
The results in the case of Pb are reported in Figs. 3 and 4 , which clearly
show that, due to the very small value of κ, the reduction is greater near the
surface. Moreover, in this particular case, some approximations made in the
case of YBCO are no longer allowed: for example, the simplified relation (99) is
not valid for small values of L. In fact, when κ is small, the length L plays an
important role and, in particular, if L is small the effect is remarkably enhanced,
as shown in Fig. 5 . In the same condition, a maximum of the effect (and therefore
a minimum of Eg ) can occur at t 6= 0, as can be seen in the same figure. In the
extreme case L = 6 λ, we found that the system returns to the unperturbed value
after a time t0 ' 105 τ .
Table 1 reports the values of the parameters of YBCO and Pb, calculated at
a temperature T? such that the quantity T?T−T c
c
is the same in the two materials.
In Tables 2 .i and 2 .ii are shown the calculated values of λ, τ and g? at different
temperatures.

18
Eg
g⋆

PSfrag repla ements


t/τ
x/λ

Fig. 1 : The gravitational field Eg /g? as a function of the normalized time and space for
YBCO at T = 77 K

x=λ

x = 10−2 λ
Eg
g⋆

x = 5 · 10−3 λ

PSfrag repla ements

t/τ

Fig. 2 : The gravitational field as a function of the normalized time for increasing values
of the x variable for YBCO

19
Eg
g⋆

PSfrag repla ements


104 t/τ
x/λ

Fig. 3 : The gravitational field Eg /g? as a function of the normalized time and space for
Pb at T = 6.3 K

x = 10 λ

x = 2λ
Eg
g⋆

x=λ

PSfrag repla ements

105 t/τ

Fig. 4 : The gravitational field as a function of the normalized time for increasing values
of the x variable for Pb

20
YBCO Pb

Tc 89 K 7.2 K

T? 77 K 6.3 K

ξ(T? ) 3.6 · 10−9 m 1.7 · 10−7 m

λ(T? ) 3.3 · 10−7 m 7.8 · 10−8 m

σ −1 4 · 10−7 Ω m 2.5 · 10−9 Ω m


[T = 90 K ] [T = 15 K ]

Hc (T? ) 0.2 Tesla 0.018 Tesla

κ 94.4 0.48

τ (T? ) 3.4 · 10−10 s 6.1 · 10−15 s

η 1.27 · 10−2 6.6 · 103

D 3.2 · 10−4 m2 /s 1 m2 /s

` 6 · 10−9 m 1.7 · 10−6 m

vf 1.6 · 105 m/s 1.83 · 106 m/s

Table 1 : YBCO vs. Pb

YBCO λ τ g? Pb λ τ g?

T =0K 1.7 · 10−7 m 9.03 · 10−11 s 2.6 · 10−12 T =0K 3.90 · 10−8 m 1.5 · 10−15 s 1 · 10−17
T = 70 K 2.6 · 10−7 m 2.1 · 10−10 s 9.8 · 10−12 T = 4.20 K 4.3 · 10−8 m 1.8 · 10−15 s 1.4 · 10−17
T = 77 K 3.3 · 10−7 m 3.4 · 10−10 s 2 · 10−11 T = 6.26 K 7.8 · 10−8 m 6.1 · 10−15 s 8.2 · 10−17
T = 87 K 8 · 10−7 m 2 · 10−9 s 2.8 · 10−7 T = 7.10 K 2.3 · 10−7 m 5.3 · 10−14 s 2.2 · 10−15

Table 2 .i : YBCO Table 2 .ii : Pb

21
Eg
10−4
g⋆
L = 10 λ

L = 8λ

L = 6λ

PSfrag repla ements

t/τ

Fig. 5 : The gravitational field Eg /g? as a function of the normalized time in the case of
Pb, for different values of L and x = 4 λ. The maximum of the shielding effect
is evident.

4 Conclusions
It is clearly seen that λ and τ grow with the temperature, so that one could think
that the effect is maximum when the temperature is very close to the critical
temperature Tc . However, this is true only for low-Tc superconductors (LTSC)
because in high-Tc superconductors fluctuations are of primary importance for
some Kelvin degree around Tc . The presence of these opposite contributions
makes it possible that a temperature Tmax ≤ Tc exists, at which the effect is
maximum. In all cases, the time constant tint is very small, and this makes the
experimental observation rather difficult.
Here we suggest to use pulsed magnetic fields to destroy and restore the
superconductivity within a time interval of the order of tint . The main conclusion
of this work is that the reduction of the gravitational field in a superconductor,
if it exists, is a transient phenomenon and depends strongly on the parameters
that characterize the superconductor.

22
A Sign convention
We work in the “mostly plus” convention, where

η = diag(−1, +1, +1, +1) . (A.1)

We define the Riemann tensor as:

Rσµλν = ∂λ Γσµν − ∂ν Γσµλ + Γσρλ Γρνµ − Γσρν Γρλµ =


(A.2)
= 2 ∂[λ Γσν]µ + 2 Γσρ[λ Γρν]µ ,
where
Γλνρ = g λµ Γµνρ ,
1 (A.3)
Γµνρ = (∂ρ gµν + ∂ν gµρ − ∂µ gνρ ) .
2

The Ricci tensor is defined as a contraction of the Riemann tensor

Rµν = Rσµσν , (A.4)

the Ricci scalar is given by


R = g µν Rµν , (A.5)
(E)
and the so-called Einstein tensor Gµν has the form

(E) 1
Gµν ≡ Rµν − gµν R . (A.6)
2
The Einstein equations are written

(E) 1
Gµν ≡ Rµν − gµν R = 8πGn Tµν , (A.7)
2
where Tµν is the total energy-momentum tensor. The cosmological constant con-
tribution can be pointed out splitting Tµν tensor in the matter and Λ component:

(M) (Λ) (M) Λ


Tµν = Tµν + Tµν = Tµν − gµν , (A.8)
8πGn

so that the Einstein equation can be rewritten as

1  (M) (Λ)

Rµν − gµν R = 8πGn Tµν + Tµν , (A.9)
2
or, equivalently,
1 (M)
Rµν − gµν R + Λ gµν = 8πGn Tµν . (A.10)
2

23
Acknoledgments
This work was supported by the Competitiveness Program of the National Re-
search Nuclear University MEPhI of Moscow for the contribution of prof. G. A.
Ummarino. We also thank Fondazione CRT that partially supported this
work for dott. A. Gallerati.

24
References
[1] B. S. DeWitt, “Superconductors and gravitational drag”, Phys. Rev. Lett.
16 (1966) 1092–1093. [p. 2]

[2] G. Papini, “Detection of inertial effects with superconducting


interferometers”, Physics Letters A 24 (1967), n. 1, 32–33. [p. 2]

[3] S. B. Felch, J. Tate, B. Cabrera, and J. T. Anderson, “Precise


determination of h/me using a rotating, superconducting ring”, Phys. Rev.
B 31 (Jun, 1985) 7006–7011. [p. 2]

[4] J. Anandan, “Relativistic gravitation and superconductors”, Classical and


Quantum Gravity 11 (1994), n. 6A, A23. [p. 2]

[5] J. Anandan, “Nuovo Cimento Soc. Ital. Fis. A 53A, 221 (1979)”, Phys.
Rev. D 15 (1977) 1448. [p. 2]

[6] J. Anandan, “Relativistic thermoelectromagnetic gravitational effects in


normal conductors and superconductors”, Physics Letters A 105 (1984),
n. 6, 280–284. [p. 2]

[7] D. Ross, “The london equations for superconductors in a gravitational


field”, Journal of Physics A: Mathematical and General 16 (1983), n. 6,
1331. [p. 2]

[8] H. Hirakawa, “Superconductors in gravitational field”, Physics Letters A


53 (1975), n. 5, 395–396. [p. 2]

[9] R. Rystephanick, “On the london moment in rotating superconducting


cylinders”, Canadian Journal of Physics 51 (1973), n. 8, 789–794. [p. 2]

[10] H. Peng, G. Lind, and Y. Chin, “Interaction between gravity and moving
superconductors”, General relativity and gravitation 23 (1991), n. 11,
1231–1250. [p. 2]

[11] C. Ciubotariu and M. Agop, “Absence of a gravitational analog to the


meissner effect”, General Relativity and Gravitation 28 (1996), n. 4,
405–412. [p. 2]

[12] M. Agop, C. G. Buzea, V. Griga, C. Ciubotariu, C. Buzea, C. Stan, and


D. Jatomir, “Gravitational paramagnetism, diamagnetism and
gravitational superconductivity”, Australian journal of physics 49 (1996),
n. 6, 1063–1074. [p. 2]

[13] O. Y. Dinariev and A. Mosolov, “A relativistic effect in the theory of


superconductivity”, in Soviet Physics Doklady, vol. 32, p. 576, 1987. [p. 2]

25
[14] I. Minasyan, “Londons equations in riemannian space”, Doklady Akademii
Nauk SSSR 228 (1976), n. 3, 576–578. [p. 2]

[15] F. Rothen, “Application de la theorie relativiste des phenomenes


irreversible a la phenomenologie de la supraconductivite”, Helv. Phys. Acta
41 (1968) 591. [p. 2]

[16] N. Li and D. Torr, “Effects of a gravitomagnetic field on pure


superconductors”, Physical Review D 43 (1991), n. 2, 457. [p. 2]

[17] H. Peng, D. Torr, E. Hu, and B. Peng, “Electrodynamics of moving


superconductors and superconductors under the influence of external
forces”, Physical Review B 43 (1991), n. 4, 2700. [p. 2]

[18] N. Li and D. G. Torr, “Gravitational effects on the magnetic attenuation of


superconductors”, Physical Review B 46 (1992), n. 9, 5489. [p. 2]

[19] D. G. Torr and N. Li, “Gravitoelectric-electric coupling via


superconductivity”, Foundations of Physics Letters 6 (1993), n. 4, 371–383.
[p. 2]

[20] E. Podkletnov and R. Nieminen, “A possibility of gravitational force


shielding by bulk YBa2 Cu3 O7−X superconductor”, Physica C:
Superconductivity 203 (1992), n. 3-4, 441–444. [p. 2]

[21] E. Podkletnov, “Weak gravitation shielding properties of composite bulk


YBa2 Cu3 O7−X superconductor below 70 k under em field”,
cond-mat/9701074 (1997). [p. 2]

[22] N. Li, D. Noever, T. Robertson, R. Koczor, and W. Brantley, “Static test


for a gravitational force coupled to type II YBCO superconductors”,
Physica C: Superconductivity 281 (1997), n. 2, 260–267. [p. 2]

[23] M. de Podesta and M. Bull, “Alternative explanation of "gravitational


screening" experiments”, Physica C: Superconductivity 253 (1995), n. 1-2,
199–200. [p. 2]

[24] C. Unnikrishnan, “Does a superconductor shield gravity?”, Physica C:


Superconductivity 266 (1996), n. 1-2, 133–137. [p. 2]

[25] G. Modanese, “Theoretical analysis of a reported


weak-gravitational-shielding effect”, EPL (Europhysics Letters) 35 (1996),
n. 6, 413. [p. 2, 3, 4]

[26] G. Modanese, “Role of a "local" cosmological constant in euclidean


quantum gravity”, Physical Review D 54 (1996), n. 8, 5002. [p. 2]

26
[27] M. Agop, C. G. Buzea, and P. Nica, “Local gravitoelectromagnetic effects
on a superconductor”, Physica C: Superconductivity 339 (2000), n. 2,
120–128. [p. 2, 6, 8]

[28] M. Agop, P. Ioannou, and F. Diaconu, “Some implications of gravitational


superconductivity”, Progress of Theoretical Physics 104 (2000), n. 4,
733–742. [p. 2, 8]

[29] R. M. Wald, “General Relativity”, Chicago, USA: Univ. Pr. (1984) 491.
[p. 2, 5]

[30] C. W. Misner, K. S. Thorne, and J. A. Wheeler, “Gravitation”. Macmillan,


1973. [p. 5]

[31] V. B. Braginsky, C. M. Caves, and K. S. Thorne, “Laboratory experiments


to test relativistic gravity”, Physical Review D 15 (1977), n. 8, 2047. [p. 8]

[32] P. Huei, “On calculation of magnetic-type gravitation and experiments”,


General relativity and gravitation 15 (1983), n. 8, 725–735. [p. 8]

[33] H. Peng, “A new approach to studying local gravitomagnetic effects on a


superconductor”, General Relativity and Gravitation 22 (1990), n. 6,
609–617. [p. 8]

[34] M. L. Ruggiero and A. Tartaglia, “Gravitomagnetic effects”, Il Nuovo


Cimento B 117 (2002) 743–768. [p. 8]

[35] B. Mashhoon, “Gravitoelectromagnetism: A Brief review”,


gr-qc/0311030. [p. 8]

[36] M. Tinkham, “Introduction to superconductivity”. Courier Corporation,


1996. [p. 9, 12]

[37] J. Ketterson and S. Song, “Superconductivity”. Cambridge University


Press, England, 1999. [p. 9, 12]

[38] P. G. De Gennes, “Superconductivity of metals and alloys”.


Addison-Wesley, 1989. [p. 9, 12]

[39] B. Mashhoon, H. J. Paik, and C. M. Will, “Detection of the


gravitomagnetic field using an orbiting superconducting gravity
gradiometer. theoretical principles”, Physical Review D 39 (1989), n. 10,
2825. [p. 12]

[40] A. Ljubičić and B. Logan, “A proposed test of the general validity of


Mach’s principle”, Physics Letters A 172 (1992), n. 1-2, 3–5. [p. 12]

[41] Q. Tang and S. Wang, “Time dependent ginzburg-landau equations of


superconductivity”, Physica D: Nonlinear Phenomena 88 (1995), n. 3,
139–166. [p. 12, 13]

27
[42] F.-H. Lin and Q. Du, “Ginzburg-landau vortices: dynamics, pinning, and
hysteresis”, SIAM Journal on Mathematical Analysis 28 (1997), n. 6,
1265–1293. [p. 12, 13]

[43] S. Ullah and A. T. Dorsey, “Effect of fluctuations on the transport


properties of type-ii superconductors in a magnetic field”, Physical Review
B 44 (1991), n. 1, 262. [p. 12]

[44] M. Ghinovker, I. Shapiro, and B. Y. Shapiro, “Explosive nucleation of


superconductivity in a magnetic field”, Physical Review B 59 (1999), n. 14,
9514. [p. 12]

[45] N. Kopnin and E. Thuneberg, “Time-dependent ginzburg-landau analysis


of inhomogeneous normal-superfluid transitions”, Physical review letters
83 (1999), n. 1, 116. [p. 12]

[46] J. Fleckinger-Pellé, H. G. Kaper, and P. Takáč, “Dynamics of the


ginzburg-landau equations of superconductivity”, Nonlinear Analysis:
Theory, Methods & Applications 32 (1998), n. 5, 647–665. [p. 12]

[47] Q. Du and P. Gray, “High-kappa limits of the time-dependent


ginzburg-landau model”, SIAM Journal on Applied Mathematics 56
(1996), n. 4, 1060–1093. [p. 12]

[48] H. Blackstead, J. D. Dow, D. Harshman, M. DeMarco, M. Wu, D. Chen,


F. Chien, D. Pulling, W. Kossler, A. Greer, et al., “Magnetism and
superconductivity in SrYRuCuO and magnetism in BaGdRuCuO”, The
European Physical Journal B-Condensed Matter and Complex Systems 15
(2000), n. 4, 649–656. [p. 17]

28

You might also like