Skip to main content
Log in

World-Line Path Integral for the Propagator Expressed as an Ordinary Integral: Concept and Applications

  • Published:
Foundations of Physics Aims and scope Submit manuscript

Abstract

The (Feynman) propagator \(G(x_2,x_1)\) encodes the entire dynamics of a massive, free scalar field propagating in an arbitrary curved spacetime. The usual procedures for computing the propagator—either as a time ordered correlator or from a partition function defined through a path integral—requires introduction of a field \(\phi (x)\) and its action functional \(A[\phi (x)]\). An alternative, more geometrical, procedure is to define a propagator in terms of the world-line path integral which only uses curves, \(x^i(s)\), defined on the manifold. I show how the world-line path integral can be reinterpreted as an ordinary integral by introducing the concept of effective number of quantum paths of a given length. Several manipulations of the world-line path integral becomes algebraically tractable in this approach. In particular I derive an explicit expression for the propagator \(G_\mathrm{QG}(x_2,x_1)\), which incorporates the quantum structure of spacetime through a zero-point-length, in terms of the standard propagator \(G_\mathrm{std}(x_2,x_1)\), in an arbitrary curved spacetime. This approach also helps to clarify the interplay between the path integral amplitude and the path integral measure in determining the form of the propagator. This is illustrated with several explicit examples.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Similar content being viewed by others

Notes

  1. As we shall see, there is some algebraic advantage in using \(\mathcal {G}_\mathrm{std}\) rather than \(G_{std}\). Of course, both contain the same amount of information in the case of a massive field, which I will be focusing on; the massless case can be treated by a limiting procedure and I will comment on it when relevant.

  2. In a time dependent background, there is the usual ambiguity of choice of positive/negative frequency mode functions, inequivalent vacuua etc. These are not relevant to the main thrust of the current discussion. Any one choice of mode functions and vacuum state is good enough for my purpose.

  3. I will work in the Euclidean space(time) for mathematical convenience and will assume that the results in the pseudo-Riemannian spacetime arise through analytic continuation. This is particularly useful in the path integral representation discussed below.

  4. This works best in Euclidean sector because path lengths \(d\sigma =\sqrt{g_{ab}dx^adx^b}\) are real. In an earlier work I have tried to do it with Lorentzian signature but it leads to ambiguities.

  5. Of course, the actual number of geometrical paths, of a given length connecting any two points in the Euclidean space, is either zero or infinity. But the effective number of paths \(N_{std}(\ell ;x_2,x_1)\), formally defined as the inverse Laplace transform of \(\mathcal {G}_{std}(x_2,x_1;m)\) (see Eq. (4)), will be a finite quantity.

  6. Notation: I will use the subscript ‘std’ for functions pertaining to a classical gravitational background, not necessarily a flat spacetime; for corresponding expressions evaluated in the flat spacetime, I will use the subscript ‘free’. In the later discussion the subscript ‘QG’ will give the corresponding functions with quantum gravitational corrections.

  7. From our definition, it follows that \(N(x,\ell )\) has the dimensions of \(L^{-D}\) in a D dimensional space(time) which is the same as that of space(time) number density. Its Fourier transform \(N(p,\ell )\) is dimensionless in all D.

  8. I want to work with a descriptor of the field dynamics which is robust enough to survive (and be useful) at mesoscopic scales. The propagator, described in terms of \(N(\ell )\), is a good choice for such a description.

  9. What happens to the classical relativistic particle if the action is modified from \(m\ell \) to another function \(A[\ell ]\) monotonic in \(\ell \)? Since \(\delta A=A'[\ell ]\delta \ell \), the equations of motion does not change. This implies that, at least in the classical case, the dispersion relation \(\omega ^2=\varvec{p}^2+m^2\) does not change by the addition of zero-point-length. In fact, it turns out that the dispersion relations for the excitations does not change even when the propagator \(\mathcal {G}_{QG}\) is obtained from a QFT but the discussion of this feature goes beyond the scope of this paper [26].

  10. One can carry out the differentiation in this expression and obtain two terms, one involving a Dirac delta function \(\delta (m_0-m)\) and the other involving a product \(\Theta ( m_0-m)J_1[ L (m_0^2 - m^2)^{1/2}]\). It turns out to be often more convenient not to carry out this differentiation and work with the expression in Eq. (19).

  11. Of course if you change \(\ell \) to some other function \(f(\ell )\) both in the measure and in the amplitude, you will change nothing in a definite integral.

References

  1. Padmanabhan, T.: Quantum Field Theory: The Why What and How. Springer, Heidelberg (2016)

    Book  Google Scholar 

  2. Padmanabhan, T.: Duality and zero-point length of spacetime. Phys. Rev. Letts. 78, 1854 (1997). [hep-th-9608182]

    Article  ADS  Google Scholar 

  3. Padmanabhan, T.: The hypothesis of path integral duality I: quantum gravitational corrections to the propagator. Phys. Rev. D 57, 6206 (1998)

    Article  ADS  MathSciNet  Google Scholar 

  4. Deser, S.: General relativity and the divergence problem in quantum field theory. Rev. Mod. Phys. 29, 417 (1957)

    Article  ADS  MathSciNet  Google Scholar 

  5. DeWitt, B.S.: Gravity: a universal regulator? Phys. Rev. Lett. 13, 114 (1964)

    Article  ADS  MathSciNet  Google Scholar 

  6. Padmanabhan, T.: Planck length is the lower bound to all physical length scales. Gen. Rel. Grav. 17, 215 (1985)

    Article  ADS  Google Scholar 

  7. Padmanabhan, T.: Physical significance of Planck length. Ann. Phys. 165, 38 (1985)

    Article  ADS  MathSciNet  Google Scholar 

  8. Padmanabhan, T.: Limitations on the operational definition of spacetime events and quantum gravity, Class. Quan. Grav. 4, L107 (1987)

    Article  ADS  Google Scholar 

  9. Mead, C.A.: Possible connection between gravitation and fundamental length. Phys. Rev. 135, B849 (1964)

    Article  ADS  MathSciNet  Google Scholar 

  10. Amati, D., Ciafaloni, M., Veneziano, G.: Can space-time be probed below the string size? Phys. Lett. B 216, 41 (1989)

    Article  ADS  Google Scholar 

  11. Yoneya, T.: On the interpretation of minimal length in string theories. Mod. Phys. Lett. A 4, 1587 (1989)

    Article  ADS  MathSciNet  Google Scholar 

  12. Konishi, K., Paffuti, G., Provero, P.: Minimum physical length and the generalized uncertainty principle in string theory. Phys. Lett. B 234, 276 (1990)

    Article  ADS  MathSciNet  Google Scholar 

  13. Greensite, J.: Is there a minimum length in D = 4 lattice quantum gravity? Phys. Lett. B 255, 375 (1991)

    Article  ADS  Google Scholar 

  14. Maggiore, M.: A generalized uncertainty principle in quantum gravity. Phys. Lett. B 304, 65 (1993). [hep-th/9301067]

    Article  ADS  Google Scholar 

  15. Srinivasan, K., Sriramkumar, L., Padmanabhan, T.: The hypothesis of path integral duality II: corrections to quantum field theoretic results. Phys. Rev. D 58, 044009 (1998). [gr-qc-9710104]

    ADS  MathSciNet  Google Scholar 

  16. Shankaranarayanan, S., Padmanabhan, T.: Hypothesis of path integral duality: applications to QED, Int. J. Mod. Phys 10(3), 351–365 (2001). [gr-qc-0003058]

    Article  ADS  Google Scholar 

  17. Kothawala, D., Sriramkumar, L., Shankaranarayanan, S., Padmanabhan, T.: Path integral duality modified propagators in spacetimes with constant curvature. Phys. Rev. 79, 104020 (2009)

    Google Scholar 

  18. Kothawala, D., Padmanabhan, T.: Entropy density of spacetime as a relic from quantum gravity. Phys. Rev. D 90, 124060 (2014)

    Article  ADS  Google Scholar 

  19. Kothawala, D.: Minimal length and small scale structure of spacetime. Phys. Rev. D 88, 104029 (2013)

    Article  ADS  Google Scholar 

  20. Padmanabhan, T.: Distribution function of the atoms of spacetime and the nature of gravity. Entropy 17, 7420 (2015)

    Article  ADS  Google Scholar 

  21. Stargen, D.J., Kothawala, D.: Small scale structure of spacetime: the van Vleck determinant and equigeodesic surfaces. Phys. Rev. 92, 024046 (2015)

    ADS  MathSciNet  Google Scholar 

  22. Kan, Nahomi, et al.: UV modified Green’s function, and higher derivative theories, [arXiv:2007.00220]; Vacuum expectation values in non-trivial background space from three types of UV improved Green’s functions, [arXiv:2007.00220]; [arXiv:2004.07527]

  23. Curiel, Erik, Finster, Felix, Isidro, J.M.: Summing over spacetime dimensions in quantum gravity. Symmetry 12(1), 138 (2020)

    Article  Google Scholar 

  24. Fontanini, Michele, Spallucci, Euro, Padmanabhan, T.: Zero-point length from string fluctuations. Phys. Lett. B633, 627–630 (2006). [hep-th/0509090]

    Article  ADS  MathSciNet  Google Scholar 

  25. Rovelli, C., Speziale, S.: Reconcile planck-scale discreteness and the Lorentz–Fitzgerald contraction. Phys. Rev. 67, 064019 (2003)

    MathSciNet  MATH  Google Scholar 

  26. Padmanabhan, T.: A class of QFTs with higher derivative field equations leading to standard dispersion relation for the particle excitations. Phys. Letts. B811, 135912 (2020)

    Article  MathSciNet  Google Scholar 

  27. Erdelyi, A. (ed.): Tables of integral transforms (Vol. I). McGraw-Hill, New York (1954)

  28. Gradshteyn, I.S., Ryzhik, I.M.: Table of Integrals. Series and Products, 7th edn. Academic Press, Cambridge (2007)

    MATH  Google Scholar 

  29. Gibbons, G.W., et al.: Higher derivative scalar quantum field theory in curved spacetime. Phys. Rev. D100, 105008 (2019)

    ADS  MathSciNet  Google Scholar 

Download references

Acknowledgements

My research is partially supported by the J.C.Bose Fellowship of Department of Science and Technology, Government of India.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to T. Padmanabhan.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendix: Calculational Details

Appendix: Calculational Details

1.1 Path Measure in Real Space

Given the path measure in the momentum space, \(N_\mathrm{free}(\ell , p) = \cos p\ell \), we can find the measure \(N_\mathrm{free}(\ell , x)\) in real space by evaluating the D-dimensional Fourier transform of \(\cos p\ell \). To do this, we start with the standard result for the Fourier transform of spherically symmetric function. If

$$\begin{aligned} F(k) = \int d^D\varvec{x}\ f(|\varvec{x}|) \, e^{-i\varvec{k\cdot x}} \end{aligned}$$
(42)

then we can write

$$\begin{aligned} k^{\frac{D-2}{2}} F(k) = (2\pi )^{D/2} \int _0^\infty J_{\frac{D-2}{2}} (kr) \, r^{\frac{D-2}{2}}\ f(r)\, r \, dr \end{aligned}$$
(43)

This allows us to write the relevant Fourier transform as:

$$\begin{aligned} N_{free}(\ell ,x) = \int _0^\infty \frac{d^Dp}{(2\pi )^D}\ e^{ip.x} \cos p\ell = \frac{1}{(2\pi )^{D/2}} \frac{1}{x^\alpha } \int _0^\infty dp \ p^{\alpha + 1} J_\alpha (p x) \cos \ell p \end{aligned}$$
(44)

where \(\alpha = (D/2) - 1\). We next evaluate this integral using the standard cosine transform (see, for e.g., page 45, 1.12 (12) of [27])

$$\begin{aligned} \int _0^\infty dx\ (\cos xy)\, x^{\nu + 1} J_\nu (ax) = \frac{2^{\nu + 1} \sqrt{\pi }\, a^\nu y \Theta [y -a]}{\Gamma \left( -\frac{1}{2} - \nu \right) (y^2 - a^2)^{\nu + \frac{3}{2}}} \end{aligned}$$
(45)

This gives the result

$$\begin{aligned} N_{free}(\ell ,x) = \frac{1}{\pi ^{\frac{D-1}{2}}} \ \frac{\Theta [\ell ^2-x^2]}{\Gamma \left( -\frac{D-1}{2}\right) }\ \frac{\ell }{[\ell ^2 - x^2]^{\frac{D+1}{2}}} \end{aligned}$$
(46)

which reduces to the expression quoted in the text when \(D=4\). (Strictly speaking the integral in Eq. (45) is defined only for \(-1<\mathrm{Re}\nu <-{1/2}\) but can be analytically continues for other \(\nu \), as often done in dimensional regularization.) Note that \(N_\mathrm{free}(\ell , x)\) vanishes for \(\ell < x\); paths, with lengths less than the geometrical length between the two points, do not contribute to the path integral which is rather nice feature.

One can directly verify that this expression leads to the correct massive propagator in real space which, of course, is obvious from the fact that \(N_\mathrm{free}(\ell , x)\) is the Fourier transform of \(N_\mathrm{free} (\ell ,p)\) and we know that the latter gives the correct propagator in the momentum space. To verify this directly we need the integral

$$\begin{aligned} \int _1^\infty dt\ \frac{t \ e^{-zt}}{(t^2-1)^{\frac{1}{2} - \nu }} = \frac{\Gamma \left( \nu + \frac{1}{2}\right) }{\sqrt{\pi }}\, \left( \frac{2}{z}\right) ^\nu \, K_{\nu +1}(z) \end{aligned}$$
(47)

which can be obtained from a standard result (see, 8.432 (3) of [28]) by differentiating with respect to z and using the recursion relation for \(K_\nu (z)\) (see page 929 (13) of [28]). Using this integral and the expression for \(N_{free}(\ell ,x)\) in Eq. (46). we find that (with \(\mu \equiv (1/2) (D+1)\))

$$\begin{aligned} \int _0^\infty d\ell \ N_\mathrm{free}(\ell ,x)\ e^{-m\ell }&= \int _x^\infty d\ell \frac{1}{\pi ^{\mu -1}} \, \frac{1}{\Gamma (1-\mu )} \, \frac{\ell \ e^{-m\ell }}{(\ell ^2 - x^2)^\mu }\nonumber \\&= \frac{x^{2(1-\mu )}}{\pi ^{\mu -1}\sqrt{\pi }} \, \left( \frac{2}{mx}\right) ^{\frac{1}{2} - \mu } K_{\mu - \frac{3}{2}} (m x) = \mathcal {G}_D(mx) \end{aligned}$$
(48)

which is indeed m times the standard massive propagator \(\mathcal {G}_D(mx) =mG_D(mx)\) in D-dimensions. In the case of \(D=4\), this reduces to

$$\begin{aligned} \mathcal {G}(x) = m \, G(x) = \frac{m}{4\pi ^2} \left( \frac{m}{x}\right) \, K_1(mx) \end{aligned}$$
(49)

which is the familiar expression.

1.2 Relation Between \(\mathcal {G}_{QG}\) and \(\mathcal {G}_{std}\)

I will now sketch the proof of Eq. (18) and Eq. (16) which involves slightly nontrivial manipulations of integrals over Bessel functions. I begin with two easily provable identities:

$$\begin{aligned} \frac{e^{-m\sqrt{\ell ^2 + L^2}}}{\sqrt{\ell ^2 + L^2}} = \int _m^\infty dm_0\ e^{-m_0\ell } \, J_0\left[ L \sqrt{m_0^2 - m^2}\right] \end{aligned}$$
(50)

and

$$\begin{aligned} \frac{1}{2s} e^{-m^2s - \frac{L^2}{4s}} = \int _m^\infty dm_0\ m_0 e^{-m_0^2s} J_0 \left[ L \sqrt{m_0^2-m^2}\right] \end{aligned}$$
(51)

The first identity in Eq. (50) can be proved by changing the integration variable on the right hand side to \(x\equiv m_0/m\) and using a standard integral (see, 6.616 (2) of [28])

$$\begin{aligned} \int _1^\infty dx\ e^{-\alpha x} J_0\left[ \beta \sqrt{x^2-1}\right] = \frac{e^{-\sqrt{\alpha ^2 + \beta ^2}}}{\sqrt{\alpha ^2 + \beta ^2}} \end{aligned}$$
(52)

The second identity in Eq. (51) can again be proved by changing the integration variable on the right hand side to \(x=m_0/m\) and using a result derivable from 6.614 (1) of [28]:

$$\begin{aligned} \int _0^\infty 2k dk \ J_0(kL) e^{-sk^2} = \frac{1}{s} \exp \left( -\frac{L^2}{4s}\right) \end{aligned}$$
(53)

Differentiating both sides of Eq. (50) with respect to m, we obtain

$$\begin{aligned} e^{-m\sqrt{\ell ^2 + L^2}}&= -\frac{\partial }{\partial m} \int _m^\infty dm_0\ e^{-m_0 \ell } J_0\left[ L \sqrt{m_0^2-m^2}\right] \nonumber \\&= \int _0^\infty dm_0 \ e^{-m_0\ell }\, (-1) \frac{\partial }{\partial m} \left\{ \Theta (m_0-m) J_0\left[ L \sqrt{m_0^2 - m^2}\right] \right\} \end{aligned}$$
(54)

This gives

$$\begin{aligned} e^{-m\sqrt{\ell ^2 + L^2}} = \int _0^\infty dm_0\ e^{-m_0\ell } \mathcal {P}[m_0; m, L] \end{aligned}$$
(55)

with \(\mathcal {P}\) defined by Eq. (19). Similarly, differentiating both sides of Eq. (51) with respect to m and manipulating as before we get

$$\begin{aligned} m e^{-m^2 s - \frac{L^2}{4s}} = \int _0^\infty dm_0 \ m_0 e^{-m_0^2 s } \mathcal {P} [m_0; m,L] \end{aligned}$$
(56)

Notice that the right hand sides of Eqs. (56) and  (55) have very similar structures with \(e^{-m_0\ell }\) replaced by \(m_0e^{-m_0^2 s}\).

The results in Eqs. (18) and  (17), which we need to prove, can now be obtained in a straightforward manner as follows: Multiply both sides of Eq. (55) by \(N(\ell )\) and integrate over \(\ell \) to get:

$$\begin{aligned} \mathcal {G}_\mathrm{QG}&= \int _0^\infty d\ell \ N(\ell ) \, e^{-m\sqrt{\ell ^2 + L^2}} = \int _0^\infty dm_0\ \mathcal {P} [ m_0; m, L] \int _0^\infty d\ell \ N(\ell ) \, e^{-m_0 \ell }\nonumber \\&= \int _0^\infty dm_0\ \mathcal {P} [ m_0; m, L]\mathcal {G}_\mathrm{std} (m_0) \end{aligned}$$
(57)

where \(\mathcal {P} [ m_0; m, L]\) is defined by Eq. (19), reproduced here for ready reference:

$$\begin{aligned} \mathcal {P}[m_0; m,L] = - \frac{\partial }{\partial m} \left\{ \Theta ( m_0-m) J_0\left[ L \sqrt{m_0^2 - m^2}\right] \right\} \end{aligned}$$
(58)

This gives Eq. (18). To obtain Eq. (17), we write \(\mathcal {G}_\mathrm{std}\) in terms of \(K_\mathrm{std}(s)\). This leads to

$$\begin{aligned} \mathcal {G}_\mathrm{QG}&= \int _0^\infty dm_0\ \mathcal {P} [ m_0; m, L]\int _0^\infty ds\ e^{-m_0^2s} K_{std}(s) m_0 \nonumber \\&= \int _0^\infty ds\ K_{std}(s)\int _0^\infty dm_0\ m_0 \, e^{-m_0^2s}\mathcal {P} [ m_0; m, L]\nonumber \\&= \int _0^\infty ds\ K_{std}(s) m e^{-m^2 s - \frac{L^2}{4s}} \end{aligned}$$
(59)

In arriving at the last equality we have used the result in Eq. (56). This proves Eq. (17).

To obtain the relation between \(G_\mathrm{QG} = \mathcal {G}_\mathrm{QG} / m\) and \(G_\mathrm{std} = \mathcal {G}_\mathrm{std} / m\) we can proceed as follows. We pull the derivative with respect to m (arising from the expression in Eq. (58)) out of the integral sign in Eq. (57) and change the variable of integration to k with \(k^2 \equiv m_0^2 - m^2\). This leads to the relation:

$$\begin{aligned} \mathcal {G}_\mathrm{QG} (m) = - \frac{\partial }{\partial m} \int _0^\infty \frac{kdk}{\sqrt{k^2+m^2}}\ J_0(Lk)\ \mathcal {G}_\mathrm{std}(m^2+k^2) \end{aligned}$$
(60)

where \(\mathcal {G}_\mathrm{std}(m^2+k^2)\) is the standard propagator with \(m^2\) replaced by \(m^2+k^2\). Using the fact that \(\mathcal {G}(M) = MG(M)\) for any mass parameter M, this is equivalent to

$$\begin{aligned} G_\mathrm{QG} (m)&= - \frac{1}{m}\frac{\partial }{\partial m} \int _0^\infty \frac{kdk}{\sqrt{k^2+m^2}}\ J_0(Lk)\ \sqrt{k^2+m^2} \, G_\mathrm{std}(m^2+k^2)\nonumber \\&= - \frac{\partial }{\partial m^2} \int _0^\infty 2k\, dk\ J_0(Lk)\ G_\mathrm{std} ( k^2 +m^2) \end{aligned}$$
(61)

This is the relation quoted in the main text; see Eq. (22). Further, for any function f(k) which depends only on the magnitude of the 2-dimensional vector \(\varvec{k}\), we have the identity

$$\begin{aligned} \int \frac{d^2k}{(2\pi )^2} \, e^{i\varvec{k\cdot L}}\, f(k) = \frac{1}{2\pi }\int _0^\infty k\,dk\ J_0(kL)\, f(k) \end{aligned}$$
(62)

Using this Eq. (61) can be expressed in the form

$$\begin{aligned} G_\mathrm{QG}(m) = - 4\pi \frac{\partial }{\partial m^2} \int \frac{d^2k}{(2\pi )^2}\, e^{i\varvec{k\cdot L}}\, G_\mathrm{std} ( k^2 +m^2) \end{aligned}$$
(63)

where \(\varvec{L}\) is a 2-dimensional vector with magnitude equal to the zero-point-length. From this it is possible to obtain a higher dimensional interpretation of \(G_\mathrm{QG}\). However, the result is probably more transparent when obtained from first principles and I will provide such a derivation:

I will now work in D dimensional space(time) and express \(G_\mathrm{QG}^D\) in D-dimensions in terms of the standard propagator \(G_\mathrm{std}^N\) in \(N=D+2\) dimensions. To do this, let us consider a fictitious \(N=D+2\), Euclidean curved space(time) with the metric

$$\begin{aligned} dS_N^2 = \left( g_{ab} dx^a dx^b\right) _D + \delta _{AB}\ dX^A dX^B \qquad \qquad\quad (A,B = 1,2) \end{aligned}$$
(64)

where we have added two “flat” directions, \(X^A\) with \(A=1,2\). (The metric \(g_{ab}\) in D dimensions, of course depends only on the D coordinates \(x^a\).) The N-dimensional Schwinger kernel for \(m=0\) (ZMSK) now factorizes and we can write

$$\begin{aligned} K^N_{m=0} \equiv {\langle x,\varvec{L}|e^{s\Box _N}|y,\varvec{0}\rangle } =\left( \frac{1}{4\pi s}\right) e^{-L^2/4s} \, K^D_{m=0}(x,y; s) \, ;\qquad L^2 \equiv L_AL^A \end{aligned}$$
(65)

Therefore, the corresponding N-dimensional, massive, Schwinger kernel becomes

$$\begin{aligned} K^N(x,\varvec{L}; y,\varvec{0}; s) = \left( \frac{1}{4\pi s}\right) e^{-(L^2/4s) - m^2s} \, K^D_{m=0}(x,y; s) \end{aligned}$$
(66)

The corresponding N-dimensional massive propagator is obtained by integrating this kernel over s in the range 0 to \(\infty \). This is almost the same as \(G_\mathrm{QG}\) but for the extra factor \((1/4\pi s)\). This factor can be taken care of by differentiating the propagator with respect to \(m^2\). We then find that

$$\begin{aligned} -4\pi \frac{\partial }{\partial m^2} \, G_\mathrm{std}^N(x,\varvec{L}; y,\varvec{0}) = \int _0^\infty ds\ e^{-(L^2/4s) - m^2s} \, K^D_{m=0}(x,y; s) = G_\mathrm{QG}^D (x,y) \end{aligned}$$
(67)

So, we have related the quantum corrected propagator \(G_\mathrm{QG}^D(x,y)\) in D dimensions to the standard Klein-Gordan propagator in the fictitious \(N=D+2\) space with the metric in Eq. (64), through the relation:

$$\begin{aligned} G_\mathrm{QG}^D(x,y) = - 4\pi \frac{\partial }{\partial m^2} G^N_\mathrm{std}(x,\varvec{L}; y, \varvec{0}) \bigg |_{\mathbf {L}^2=L^2} \end{aligned}$$
(68)

The \(N=D+2\) dimensional propagator, \(G^N_\mathrm{std}(x,\varvec{L}; y, \varvec{0})\), of course has a standard QFT interpretation in the curved spacetime. The zero-point-length in D dimensions arises as the magnitude of the (fictitious) propagation distance in the extra dimensions. This proves Eq. (24) in the main text.

The derivative of standard Klein-Gordan propagator with respect to \(m^2\)—which appears in the right hand side of Eq. (68) —can be related to the ‘transitivity integral’ for the propagator \(G^N_\mathrm{std}\). To see this, consider the following integral (with the notation \(d^N\bar{z} \equiv d^Dz\, d^2Z\)):

$$\begin{aligned}&\int d^N\bar{z} \, G^N_\mathrm{std}(x,\varvec{L}; z,\varvec{Z}) \, G^N_\mathrm{std}(z,\varvec{Z}; y,\varvec{0}) \nonumber \\&\quad = \int d^N\bar{z} \ {\langle x,\varvec{L}|(-\Box ^N + m^2)^{-1}|z,Z\rangle } {\langle z,Z|(-\Box ^N + m^2)^{-1}|y,\varvec{0}\rangle }\nonumber \\&\quad = {\langle x,\varvec{L}|(-\Box ^N + m^2)^{-2}|y,\varvec{0}\rangle } \end{aligned}$$
(69)

We can, however, write:

$$\begin{aligned} {\langle x,\varvec{L}|(-\Box ^N + m^2)^{-2}|y,\varvec{0}\rangle } = - \frac{\partial }{\partial m^2} \, G^N_\mathrm{std}(x,\varvec{L}; y,\varvec{0}) \end{aligned}$$
(70)

Combining this result with Eq. (68) we find that

$$\begin{aligned} G^D_{QG} = (4\pi ) {\langle x,\varvec{L}|(-\Box ^N + m^2)^{-2}|y,\varvec{0}\rangle } \bigg |_{\mathbf {L}^2=L^2} \end{aligned}$$
(71)

This result, in turn, can expressed in the form of either of the following differential equations for \(G^D_{QG}\), viz.:

$$\begin{aligned} (-\Box ^N + m^2)^{2} G^D_{QG} =4\pi \delta (x,y)\, \delta (\varvec{L}) \end{aligned}$$
(72)

and

$$\begin{aligned} (-\Box ^N + m^2) G^D_{QG} = 4\pi G^N_\mathrm{std}(x, \varvec{L}; y,\varvec{0}) \end{aligned}$$
(73)

These are valid any curved space(time). We can now write \(G^N_\mathrm{std}(x, \varvec{L}; y,\varvec{0})\) as the standard vacuum expectation value of time ordered products of the KG field operators in \(N=D+2\) space(time); then the QG corrected propagator in D dimensional space is given by a solution to Eq. (73). This proves Eqs. (25) and  (26).

We can, of course, verify that, in flat space(time) either of these equations lead to the correct \(G_\mathrm{QG}\). For example, if we Fourier transform either Eqs. (72) or  (73), we will find that \(G^D_\mathrm{QG}\) can be expressed as the integral

$$\begin{aligned} G^D_{QG} (x, \varvec{L}; 0,\varvec{0}) = 4\pi \int \frac{d^Dk \, d^2K}{(2\pi )^N} \ \frac{e^{ikx}\, e^{i\varvec{K\cdot L}}}{(k^2 + m^2 + K^2)^2} \end{aligned}$$
(74)

The square appearing in the denominator can be taken care of by the usual trick of differentiating the expression with respect to \(m^2\). Performing the 2-dimensional integral over the measure \(d^2K = K dK d\theta \) we get the result in terms of the Bessel function \(J_0(KL)\):

$$\begin{aligned} G^D_{QG} (x, \varvec{L}; 0,\varvec{0})&= -\frac{\partial }{\partial m^2} \int \frac{d^Dk}{(2\pi )^D} \int _0^\infty K dK\int _0^{2\pi }\frac{d\theta }{\pi } \, \frac{e^{ikx}\, e^{i{K L\cos \theta }}}{k^2 + m^2 + K^2} \nonumber \\&= -\frac{\partial }{\partial m^2} \int \frac{d^Dk}{(2\pi )^D}\, e^{ikx}\int _0^\infty 2K dK\ \frac{J_0(KL)}{k^2 + m^2 + K^2} \end{aligned}$$
(75)

Therefore the D-dimensional Fourier transform \(G^D_\mathrm{QG} (k,\varvec{L})\) of \(G^D_\mathrm{QG}(x, \varvec{L}; 0,\varvec{0})\) is given by

$$\begin{aligned} G^D_\mathrm{QG} (k,\varvec{L}) = - \frac{\partial }{\partial m^2} \int _0^\infty 2K dK \ \frac{J_0(KL)}{k^2 + m^2+K^2} \end{aligned}$$
(76)

To perform the integral over K we write the denominator in the exponential form, leading to:

$$\begin{aligned} -\frac{\partial }{\partial m^2} \int _0^\infty 2K dK&\ J_0(KL) \int _0^\infty ds\ e^{-s(k^2+m^2)} \, e^{-sK^2}\nonumber \\&= \int _0^\infty ds\ s e^{-s(k^2+m^2)} \times \ \int _0^\infty 2K \, dK\ J_0(KL) e^{-sK^2} \end{aligned}$$
(77)

Finally, we use the identity

$$\begin{aligned} \int _0^\infty 2K dK \ J_0(KL) e^{-sK^2} = \frac{1}{s} \exp \left( -\frac{L^2}{4s}\right) \end{aligned}$$
(78)

to recover the standard result

$$\begin{aligned} G^D_\mathrm{QG} (k,{L}) = \int _0^\infty ds\ e^{-s(k^2+m^2) - (L^2/4s)} \end{aligned}$$
(79)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Padmanabhan, T. World-Line Path Integral for the Propagator Expressed as an Ordinary Integral: Concept and Applications. Found Phys 51, 35 (2021). https://doi.org/10.1007/s10701-021-00447-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s10701-021-00447-8

Navigation