Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Physics Letters B 533 (2002) 138–145

www.elsevier.com/locate/npe

Weyl invariant gonihedric strings


G.K. Savvidy, R. Manvelyan 1
National Research Center Demokritos, Ag. Paraskevi, GR-15310 Athens, Hellenic Republic, Greece
Received 4 December 2001; accepted 7 March 2002
Editor: L. Alvarez-Gaumé

Abstract
We study quantum corrections in the earlier proposed string theory, which is based on Weyl invariant purely extrinsic
curvature action. At one-loop level it remains Weyl invariant irrespective of the dimension D of the embedding spacetime.
To some extent the counterterms are reminiscent of the ones in pure quantum gravity. At classical level the string tension is
equal to zero and quarks viewed as open ends of the surface are propagating freely without interaction. We demonstrate that
quantum fluctuations generate nonzero area term (string tension).  2002 Elsevier Science B.V. All rights reserved.

1. Introduction

In [1,2] authors suggested so-called gonihedric model of random surfaces, which is based on the concept of
extrinsic curvature. It differs in two essential points from the models considered in the previous studies [3–6]: first,
it claims that extrinsic curvature term alone should be considered as the fundamental action of the theory
 

S = m · L = m d 2 ζ g Kaia Kbib , (1)

here m has dimension of mass, Kab i is a second fundamental form (extrinsic curvature) and there is no area term

in the action. Secondly, it is required that the dependence on the extrinsic curvature should be such that the action
will have dimension of length L ∝ length, that is proportional to the linear size of the surface similar to the path
integral.2 When the surface degenerates into a single world line the functional integral over surfaces will naturally
transform into the Feynman path integral for point-like relativistic particle

S = mL → m ds. (2)

At the classical level the string tension in this theory is equal to zero and quarks viewed as open ends of the surface
are propagating freely without interaction Tclassical = 0 because the action (1) is equal to the perimeter of the flat

E-mail addresses: savvidy@mail.demokritos.gr (G.K. Savvidy), manvel@moon.yerphi.am (R. Manvelyan).


1 Permanent address: Yerevan Physics Institute.
2 This is in contrast with the previous proposals when the extrinsic curvature term is a dimensionless functional S(extrinsic curvature) ∝ 1,
invariant only under rigid scale transformations.

0370-2693/02/$ – see front matter  2002 Elsevier Science B.V. All rights reserved.
PII: S 0 3 7 0 - 2 6 9 3 ( 0 2 ) 0 1 5 5 7 - 5
G.K. Savvidy, R. Manvelyan / Physics Letters B 533 (2002) 138–145 139

Wilson loop S → m(R + T ). It was demonstrated in [1] that quantum fluctuations generate the area term A in the
effective action
Seff = mL + Tq A + · · · , (3)
with the nonzero string tension Tq = aD2 (1 − ln D
β ), here
D is the dimension of the spacetime, a is a scaling
parameter. In the scaling limit β → βc = D/e the string
tension has a finite limit. Therefore, at the tree level
the theory describes free quarks with string tension equal to zero, instead quantum fluctuations generate nonzero
string tension and, as a result, quark confinement. The theory may consistently describe asymptotic freedom and
confinement as it is expected to be the case in QCD.
Dynamical string tension Tq has been found in the discrete formulation of the theory when the action (1)
is written for the triangulated surfaces. Our aim now is to show that this theory is well defined in one-loop
approximation and that similar generation of nonzero string tension takes place in a continuum formulation of
the theory and therefore may lead to a nontrivial string theory. Here we shall treat quantum fluctuations in two
different ways following the works of Polyakov [3] and Kleinert [4].3

2. Weyl invariance

We shall represent the gonihedric action (1) in a Weyl invariant form


 

S = m d 2 ζ g (∆(g)Xµ )2 , (4)
√ √
here gab = ∂a Xµ ∂b Xµ is induced metric, ∆(g) = 1/ g ∂a g g ab ∂b is a Laplace operator and Kaia Kbib =
(∆(g)Xµ )2 . The second fundamental form K is defined through the relations:
i i
Kab nµ = ∂a ∂b Xµ − Γab
c
∂c Xµ = ∇a ∂b Xµ , (5)
niµ njµ = δij , niµ ∂a Xµ = 0, (6)
where niµ are D − 2 normals and a, b = 1, 2; µ = 0, 1, 2, . . . , D − 1; i, j = 1, 2, . . . , D − 2.
We can introduce independent metric coordinates gab using standard Lagrange multipliers λab and then fix
conformal gauge gab = ρδab using reparametrization invariance. After fixing conformal gauge we shall have
  2  
S=m d ζ 2
∂ 2 Xµ + λab ∂a Xµ ∂b Xµ − ρδab . (7)

As one can see the first term is ρ independent and therefore is explicitly Weyl invariant (local scale
transformations). Note that extrinsic curvature part of the Polyakov–Kleinert action
 
1 2 √ 1  2
SPK = d ζ g Ka Kb =
ia ib
d 2 ζ ρ −1 ∂ 2 Xµ
e e
is invariant only under rigid scale transformations. This action leads to the following equation of motion:
∂ 2 Xµ  
∂2  − ∂a λab ∂b Xµ = 0, ∂a Xµ ∂b Xµ − ρδab = 0, λaa = 0. (8)
(∂ 2 Xν )2

In the light cone coordinates ζ ± = (ζ 0 ± ζ 1 )/ 2 the conformal gauge looks like:
g++ = g−− = 0, g+− = ρ, (9)

3 See also subsequent publications [13–18].


140 G.K. Savvidy, R. Manvelyan / Physics Letters B 533 (2002) 138–145

+ −
the connection has only two nonzero components Γ++ = ∂+ ln ρ, Γ−− = ∂− ln ρ and metric variation is:
δg±± = ∇± ε± = ρ∂± ε∓ . (10)
In these coordinates the partition function takes the following form:

Z = exp ı{S} det{∇+ } det{∇− }DXµ Dλab Dρ, (11)

where det{∇+ }, det{∇− } are Faddeev–Popov determinants corresponding to the conformal gauge (9).
To obtain quantum correction to the classical action we have to expand our action around some classical
µ
1 , ρ1 ), thus S = S + S2 + Sint , where S =
solution (Xµ , λ̄ab , ρ̄ ) up to second order on small fluctuations (X1 , λab
 2 √
m d ζ n̄2 ,
 
m µ 4 n̄µ n̄ν  ab a µ  2
S2 = √ X1 ∂ η − √ µν
X1 d ζ + m
ν 2
2λ1 ēµ ∂b X1 − λaa
1 ρ1 d ζ,
2 n̄ 2 n̄2
 ab µ µ µ µ 
Sint = m d ξ λ̄ ∂a X1 ∂b X1 + λab
2
1 ∂a X1 ∂b X1 + · · · , (12)

where we have introduced convenient notations


n̄µ = ∂ 2 Xµ , ēµa = ∂a Xµ .
Here we admit slow behavior of the first and second derivatives of the classical solution Xµ and natural separation
of interaction with external field λ̄ab .
µ
One can see from (12) that in one-loop approximation we have factorization of tangential modes of X1
and cancellation of their contribution with the corresponding ghost determinant. To see that, we shall expand
µ
X1 into tangential and normal fields φa , ξ i :
X1µ = φa ēµa + ξ i n̄iµ , (13)
n̄iµ ēµa = 0, ēµa n̄µ = 0. (14)
The last relation, together with the following ones n̄µ n̄iµ = Kaia ≡ K i , n̄µ n̄µ = Kaia Kbib ≡ K 2 , can be easily seen
in conformal gauge (9). We also have
 
S = m d 2ζ K 2. (15)

We have to substitute expansion (13) into (12) and take into account slow behavior of classical fields compared
with quantum fields:

m KiKj j  
S2 = d 2 ζ √ ξ i ∂ 4 δ ij − 2
ξ + λ±±
1 ∇± φ± + λ+−
1 ρ0 ∂+ φ + + ∂− φ − − λ+−
1 ρ1 ,
2 K2 K

 
Sint = m d 2 ζ λ̄ab ∂a ξ i ∂b ξ i + ∂a φ c ∂b φc . (16)

Using the last expression for S2 in partition function (11) and integrating it over ρ1 and λ±±1 we shall get delta
functions δ(λ+−
1 )δ(∇ ± φ ± ). Then integrating over λ+−
1 and longitudinal components φ ± we shall get determinants

det−1 ∇± . We observe now that in the one-loop approximation there is a cancellation of these determinants with
the ghost determinants in (11) and therefore absence of conformal anomalies. This should be verified in the next
µ µ
order where we have to consider third order interactions of the form λab
1 ∂a X1 ∂a X1 .
Finally we have the following one-loop partition function
  
    
Z1 = exp i S exp iS2 K, ξ i + iSint λ̄, ξ i Dξ i , (17)
G.K. Savvidy, R. Manvelyan / Physics Letters B 533 (2002) 138–145 141

where

m KiKj j
S2 = d 2 ζ √ ξ i ∂ 4 δ ij − ξ , (18)
2 K2
 2 K
Sint = m d 2 ζ λ̄ab ∂a ξ i ∂b ξ i . (19)

From quadratic part (18) we can deduce the propagator



 K 2 Π ij KiKj
ξ (p)ξ (−p) =
i j
2 2
, Π ij = δ ij − , Π ii = D − 3 (20)
m (p ) K2
and calculate first correction to the classical action (15). For that we have to contract ξ i fields in (19) using (20)
  
 D−3  
W1 = m d ζ λ̄ ∂a ξ ∂b ξ =
2 ab i i
log Λ/Λ̃ d 2 ζ λ̄aa K 2 , (21)

thus m1 = m + λ̄aa D−32π log(Λ/Λ̃). On the classical trajectory (8) we have λ̄
aa = 0 and, therefore, there is no

quantum corrections on-mass shell S-matrix elements and the theory is one-loop ultraviolet finite. The absence
of quantum corrections at one-loop level is very similar to the pure quantum gravity in four dimensions [7].
One can expect that summation of all one-loop diagrams in external field λ̄ab may generate nontrivial solution
and condensation of Lagrange multiplier λ̄aa = 0 on quantum level due to the delicate mixture of ultraviolet and
infrared divergences [8–11].

3. One-loop effective action

Our aim here is to sum up all one-loop diagrams in the λ̄ab background. For that we have to keep all one-loop
diagrams which are induced by the vertex λ̄ab ∂a ξ i ∂b ξ i . Integrating (17) together with the interaction term one can
get [9,10]
i
W1 = Tr ln H , (22)
2
where
   
ij = Πij
H
2
∂ 2 − 2 K 2 λ̄ab ∂a ∂b . (23)

We are looking for the solution in the form



λ̄ab = λ g g ab = λδ ab , (24)
where λ is a constant field. Then
   
Hij = Πij ∂ 2 2 − 2 K 2 λ∂ 2 . (25)

 = H ∗ H0 ,
It is convenient to introduce the notation Pα = i∂α and we shall factor this operator into two pieces H
where
  
Hij = −Πij P 2 + 2λ K 2 , H0ij = −Πij P 2 . (26)

The effective action takes the form


 
i ds i ds
L1 = − Tr U (s) − Tr U0 (s), (27)
2 s 2 s
142 G.K. Savvidy, R. Manvelyan / Physics Letters B 533 (2002) 138–145

where trace Tr is over Lorentz and world sheet coordinates ζ and

U (s) = exp(−iH s), U0 (s) = exp(−iH0s). (28)


Taking the trace over Lorentz indexes and using the matrix element
        iP 2 s      2
ζ (s)ζ (0) = ζ e ζ = 1 exp −i (ζ − ζ )
4πs 4s
one can get

i ds D − 3 2iλs √K 2 D − 3
L1 = − e + (29)
2 s 4πs 4πs
and after rotation of the contour by s → −is and substraction of vacuum contribution we shall get

D − 3 ds  −2λs √K 2 
L1 = 2
e − 1 + Cλ. (30)
8π s
The counterterm is equal to Cλ and we can fix it by normalization condition [8,9]:
 
∂L1 
= K 2, (31)
∂λ  λ=m
after which we can get ultraviolet and infrared finite effective action in the form
∞
D−3 ds  −λ̃s 
L1 = λ̃/2 + e − 1 + λ̃se−m̃s , (32)
8π s2
0
√ √
where λ̃ = 2λ K 2 and m̃ = 2m K 2 . This expression can be integrated out and we shall get
  
D−3  2 λ
L1 = λ K 2 + λ K ln −1 (33)
4π m
with its new minimum at the point λ = m exp(− D−3

). The dynamical string tension is equal, therefore, to
Tq = m exp(− D−3 ) because
2 4π



Seff = m2 e− D−3 d 2 ζ g g ab ∂a Xµ ∂b Xµ ,

(34)

as one can see from (7).

4. Consideration in the physical gauge

To study quantum effects from a different perspective we shall consider only normal, physical perturbation of
the world sheet Xµ (ζ ) [4,12]
µ = Xµ + ξ i niµ ,
X (35)
µ = ∂a Xµ + ∂a ξ i niµ + ξ i ∂a niµ and the metric is equal to
then the first derivative is ∂a X

g̃ab = gab + ξ i ∂a Xµ ∂b niµ + ξ i ∂b Xµ ∂a niµ + ∂a ξ i ξ j niµ ∂b njµ + ξ i ∂b ξ j ∂a niµ njµ


+ ∂a ξ i ∂b ξ i + ξ i ξ j ∂a niµ ∂b niµ , (36)
G.K. Savvidy, R. Manvelyan / Physics Letters B 533 (2002) 138–145 143

then it follows that

g̃ ab = g ab + 2K iab ξ i − ∇ a ξ i ∇ b ξ i + 3Kcia K j cb (37)


and that
 
1 1
g̃ 1/2 = g 1/2 1 − ξ i Kaia + R ij ξ i ξ j + g ab ∇a ξ i ∇b ξ j , (38)
2 2
where
jb jb
i
Kab = ∂a ∂b Xµ niµ , R ij = Kaia Kb − Kbia Ka . (39)
For the second derivative we have
µ = ∂a ∂b Xµ + ∂a ∂b ξ i niµ + ∂b ξ i ∂a niµ + ∂a ξ i ∂b niµ + ξ i ∂a ∂b niµ .
∂a ∂b X
In order to compute the variation of the extrinsic curvature we have to find the perturbation of the normals
µ = 0, ñiµ ñjµ = δ ij . From this it follows that
δniµ = ñiµ − niµ , where ñiµ ∂a X
 
∂a ξ i + δniµ ∂a Xµ + δniµ ∂a njµ ξ j + njµ ∂a ξ j = 0, niµ δnjµ + njµ δniµ + δniµ δnjµ = 0 (40)
j
and we can always represent δniµ in the form δniµ = Aia ∂a Xµ + B ij nµ . The solution of (40) is
 jb  1
δniµ = −∂ a Xµ ∂a ξ i + Ka ξ j ∂b ξ i − njµ ∂ a ξ j ∂a ξ i . (41)
2
With this formulas we can get the variation of the extrinsic curvature up to the second order
  
i = ∂a ∂b Xµ + ∂a ∂b ξ i niµ + ∂b ξ i ∂a niµ + ∂a ξ i ∂b niµ + ξ i ∂a ∂b niµ niµ + δniµ ,
K (42)
ab

i − K i
i =K
substituting (41) one can get for the variation δKab ab ab

j jc  j j  1 j
δKabi
= ∇a ∇b ξ i − Kaic Kcb ξ j + ∇b Ka ξ j ∇c ξ i + Kbc ∇a ξ j + Kac ∇b ξ j ∇ c ξ i − Kab ∇ c ξ j ∇c ξ i . (43)
2
Using (37) we can get the trace of the extrinsic curvature
ja j j
δKaia = ∇ a ∇a ξ i + Kaic Kc ξ j + ∇ a Kab ξ j ∇ b ξ i + 2Kab ∇ a ξ j ∇ b ξ i − Kab
i
∇a ξ j ∇bξ j
j 1 ja j
+ Kab Kcib K nca ξ j ξ n − Ka ∇ c ξ j ∇c ξ i + 2Kab ξ j ∇ a ∇ b ξ i . (44)
2
Using the last expression it is easy to compute the first and the second variations of the action (1):

√ i  ij  Ka ja
ja
δ1 AGonihedric = m d 2ζ g ξ δ ∆ + Kaib Kb − δ ij K 2 √ , (45)
K2

m √ 1
δ2 AGonihedric = g√
d 2ζ
2 K2
KiKj ja
× ∆ξ i ∆ξ i − ∆ξ i ∆ξ i + K i K j ∇ c ξ i ∇c ξ j + K i K i ∇ c ξ j ∇c ξ j − 2Kaib Kb ∇ c ξ i ∇c ξ j
K2
2 ja
− 2K i Kbci
∇ b ξ j ∇ c ξ j + 2 K n Kanb Kb K i ∇ e ξ i ∇e ξ j , (46)
K
144 G.K. Savvidy, R. Manvelyan / Physics Letters B 533 (2002) 138–145

where K i = Kaia and K 2 = Kaia Kbib = K i K i . The equation of motion is


 ij  Kj
δ ∆ − R ij √ = 0, (47)
K2
ja
where R ij = K i K j − Kaib Kb . It is easy to see that

 ij K 1
2  
∇a ξ ∇b ξ = δab Π
i j
log Λ/Λ̃ , (48)
m 2π
where Π ij = δ ij − K i K j /K 2 . Using (48) one can find one-loop contribution

m ij K 1
2  
√ K K 2Π i j
log Λ/Λ̃ (49)
2 K2 m 2π

m K2 1  
+ √ K 2(D − 3)2
log Λ/Λ̃ (50)
2 K 2 m 2π

m ib j a ij K 1
2  
+ √ (−2)Ka Kb 2Π log Λ/Λ̃ (51)
2 K2 m 2π

m K2 1  
+ √ (−2)K 2 (D − 3) log Λ/Λ̃ (52)
2 K 2 m 2π

m 2 n nb j a i ij K 1
2  
+ √ 2
K K a K b K 2Π log Λ/Λ̃ . (53)
2 K 2 K m 2π
The first and the last terms are equal to zero because Π ij K j = 0, the second and the fourth terms cancel each other
and we see that in this theory the counterterms do not depend on the dimension of the embedding spacetime. Only
the third therm is non-zero
ja 1     1   1  
−2Kaib Kb Π ij log Λ/Λ̃ = 2 K i K j − R ij Π ij log Λ/Λ̃ = −2R log Λ/Λ̃ , (54)
2π 2π 2π
here we have used equation of motion R ij K j = 0. Thus it is proportional to the Euler characteristic of the surface
and can be neglected. This is completely consistent with our previous result that the theory is one-loop ultraviolet
finite.
It is instructive to compare this result with the quantum corrections in the theory with dimensionless action
[3,4]. The second variation of the Polykov–Kleinert action is

1 √ jb 1
δAPK = d 2 u g ∆ξ i ∆ξ i + Kaia Kb ∇ c ξ i ∇c ξ j + Kaia Kbib ∇ c ξ j ∇c ξ j
2e0 2
ja
− 2Kaib Kb ∇ c ξ i ∇c ξ j − 2Kaia Kbc
i
∇bξ j ∇cξ j , (55)

and the one-loop correction renormalizes the coupling constant


1 1 D  
= − log Λ/Λ̃ . (56)
e e0 4π

Acknowledgement

One of the authors (R.M.) is indebted to the National Research Center Demokritos for kind hospitality. The work
of G. Savvidy was supported in part by the by EEC Grant no. HPRN-CT-1999-00161. The work of R. Manvelyan
G.K. Savvidy, R. Manvelyan / Physics Letters B 533 (2002) 138–145 145

was supported by the Hellenic Ministry of National Economy Fellowship Nato Grant and partially by Volkswagen
Foundation of Germany.

References

[1] G.K. Savvidy, K.G. Savvidy, Mod. Phys. Lett. A 8 (1993) 2963;
G.K. Savvidy, JHEP 0009 (2000) 044.
[2] R.V. Ambartzumian, G.K. Savvidy, K.G. Savvidy, G.S. Sukiasian, Phys. Lett. B 275 (1992) 99;
G.K. Savvidy, K.G. Savvidy, J. Mod. Phys. A 8 (1993) 3993;
B. Durhuus, T. Jonsson, Phys. Lett. B 297 (1992) 271.
[3] A. Polyakov, Nucl. Phys. B 268 (1986) 406.
[4] H. Kleinert, Phys. Lett. B 174 (1986) 335.
[5] W. Helfrich, Z. Naturforsch C 28 (1973) 693;
W. Helfrich, J. Phys. (Paris) 46 (1985) 1263;
L. Peliti, S. Leibler, Phys. Rev. Lett. 54 (1985) 1690;
D. Forster, Phys. Lett. A 114 (1986) 115;
T.L. Curtright et al., Phys. Rev. Lett. 57 (1986) 799;
T.L. Curtright et al., Phys. Rev. D 34 (1986) 3811;
F. David, Europhys. Lett. 2 (1986) 577;
P.O. Mazur, V.P. Nair, Nucl. Phys. B 284 (1987) 146;
E. Braaten, C.K. Zachos, Phys. Rev. D 35 (1987) 1512;
E. Braaten, R.D. Pisarski, S.M. Tye, Phys. Rev. Lett. 58 (1987) 93;
P. Olesen, S.K. Yang, Nucl. Phys. B 283 (1987) 73;
R.D. Pisarski, Phys. Rev. Lett. 58 (1987) 1300.
[6] D. Weingarten, Nucl. Phys. B 210 (1982) 229;
A. Maritan, C. Omero, Phys. Lett. B 109 (1982) 51;
T. Sterling, J. Greensite, Phys. Lett. B 121 (1983) 345;
B. Durhuus, J. Fröhlich, T. Jonsson, Nucl. Phys. B 225 (1983) 183;
J. Ambjørn, B. Durhuus, J. Fröhlich, T. Jonsson, Nucl. Phys. B 290 (1987) 480;
T. Hofsäss, H. Kleinert, Phys. Lett. A 102 (1984) 420;
M. Karowski, H.J. Thun, Phys. Rev. Lett. 54 (1985) 2556;
F. David, Europhys. Lett. 9 (1989) 575.
[7] G. ’t Hooft, M.J. Veltman, Ann. Poincaré Phys. Theor. A 20 (1974) 69.
[8] S. Coleman, E. Weinberg, Phys. Rev. D 7 (1973) 1888.
[9] G.K. Savvidy, Phys. Lett. B 71 (1977) 133.
[10] I.A. Batalin, S.G. Matinyan, G.K. Savvidy, Yad. Fiz. 25 (1977) 6.
[11] A.M. Polyakov, Gauge Fields and Strings, Harwood Academic, Chur, 1987.
[12] W. Blaschke, Vorlesungen über Differentialgeometrie, Springer, Berlin, 1930.
[13] J. Wheater, Phys. Lett. B 208 (1988) 388;
J. Pawelczyk, Nucl. Phys. B 491 (1997) 515.
[14] J. Polchinski, Z. Yang, Phys. Rev. D 46 (1992) 3667.
[15] H. Kleinert, A. Chervyakov, Phys. Lett. B 381 (1996) 286.
[16] M.C. Diamantini, C.A. Trugenberger, hep-th/0203053.
[17] R. Parthasarathy, K.S. Viswanathan, J. Geom. Phys. 38 (2001) 207.
[18] M.S. Plyushchay, Phys. Lett. B 253 (1991) 50;
Yu.A. Kuznetsov, M.S. Plyushchay, Nucl. Phys. B 389 (1993) 181.

You might also like