Appendix A

A short course on path integrals

Path integrals are a powerful means of representing quantum theories, especially on surfaces of nontrivial topology. The introduction we present here is similar to that which the reader will find in any modern field theory text. We include it in order to emphasize certain ideas that we will need, such as the relation between the path integral and Hilbert space formalisms and the use of operator equations inside the path integral.

A.1    Bosonic fields

Consider first a quantum mechanics problem, one degree of freedom with Hamiltonian Image, where

Image

Throughout the appendix operators are indicated by hats. A basic quantity of interest is the amplitude to evolve from one q eigenstate to another in a time T,

Image

In field theory it is generally convenient to use the Heisenberg representation, where operators have the time dependence

Image

The state |q, t Image is an eigenstate of Image(t),

Image

In terms of the t = 0 Schrödinger eigenstates this is

Image

In this notation the transition amplitude (A.1.2) is Imageqf, T \qi, 0Image.

By inserting a complete set of states, we can write the transition amplitude as a coherent sum over all states q through which the system might pass at some intermediate time t :

Image

Divide the time interval further into N steps as in figure A.1,

Image

Fig. A.1.   Transition amplitude broken down into time steps. The dashed line suggests a piecewise linear path, which is one of many ways to define path integration.

Image

and corresponding to each intermediate time insert a complete set of states

Image

Translate back to Schrödinger formalism and introduce an integral over intermediate momenta,

Image

By commuting, we can always write Image with all Image to the left and all Image to the right, so that

Image

Use this to evaluate the matrix element in eq. (A.1.9). From expanding the exponential there will be terms of order Image2 and higher that have Image to the left of Image; drop these. Then to this order Image becomes

Image

Thus

Image

In the last line we have taken a formal Image → 0 limit, so the integral runs over all paths p(t), q(t) with given q(T) and q(0). This is the Hamiltonian path integral (or functional integral), the sum over phase space paths weighted by exp(iS/Image) with S the action in Hamiltonian form.

Let us integrate over p(t), making first the approximation that the integral is dominated by the point of stationary phase,

Image

Solving for p in terms of q and Image and recalling that L = pImageH, the stationary phase approximation gives

Image

Often the momentum integral is Gaussian, Image, so that the stationary phase approximation is exact up to a normalization that can be absorbed in the measure [dq]. In fact, we can do this generally: because the relation (A.1.13) between p and Image is local, the corrections to the stationary phase approximation are local and so can be absorbed into the definition of [dq] without spoiling the essential property that [dq] is a product over independent integrals at each intermediate time. The result (A.1.14) is the Lagrangian path integral, the integral over all paths in configuration space weighted by exp(iS/Image), where now S is the Lagrangian action.

We have kept Image so as to discuss the classical limit. As Image → 0, the integral is dominated by the paths of stationary phase, ImageS/Imageq(t) = 0, which is indeed the classical equation of motion. Henceforth, Image = 1.

The derivation above clearly generalizes to a multicomponent qi. One can always write a quantum field theory in terms of an infinite number of qs, either by making the spatial coordinate x discrete and taking a limit, or by expanding the x-dependence in terms of Fourier modes. Thus the above derivation of the path integral representation applies to bosonic field theory as well.

The derivation of the path integral formula treated the continuum limit Image → 0 rather heuristically. Renormalization theory is the study of this limiting process, and is rather well understood. The limit exists provided the Lagrangian satisfies certain conditions (renormalizability) and provided the coeffcients in the Lagrangian (bare couplings) are taken to vary appropriately as the limit is taken. The result is independent of various choices that must be made, including the operator ordering in the Hamiltonian, the precise definition of the measure, and the differences between regulators (the limit of piecewise linear paths suggested by the figure versus a frequency cutoff on the Fourier modes, for example).1 Being local effects these can all be absorbed into the bare couplings. Actually, we will evaluate explicitly only a few Gaussian path integrals. We use the path integral primarily to derive some general results, such as operator equations and sewing relations.

Relation to the Hilbert space formalism

Starting from the path integral one may cut it open to recover the Hilbert space of the theory. In equation (A.1.6), if one writes each of the amplitudes as a path integral one finds

Image

That is, the path integral on [0, T] breaks up into separate path integrals on [0,t] and [t, T], plus an ordinary integral over q(t). The boundary conditions on paths are denoted by super/subscripts. Note the restriction that Tt ≥ 0. One can also replace the intermediate integral with a sum over any complete set of states. This cutting (or, inversely, sewing) principle is particularly useful on the surfaces of nontrivial topology that arise in string theory. There one may cut the surface open along many different closed curves; different choices give different Hilbert space representations of a given amplitude. This is known as world-sheet duality.

Consider now a path integral with the additional insertion of a factor of q(t), where 0 < t < T. Then, using the cutting relation (A.1.15),

Image

Thus, q(t) in the functional integral translates into Image(t) in the matrix element. For a product of two insertions q(t)q(t) (which is equal to q(t)q(t), since these are simply variables of integration), where both t and t are in the range [0, T], one finds by the same method that

Image

Here, T denotes the time-ordered product

Image

Because of the way the path integral is built out of successive infinitesimal time slices, two or more insertions in the path integral will always correspond to the time-ordered product of operators in the matrix element.

The equation of motion is Image Let us see how this works inside the path integral. Consider this equation in a path integral with other insertions Image,

Image

We have integrated by parts in the second line.2 If none of the fields in Image are at time t, the right-hand side vanishes and the equation of motion holds in the path integral. We therefore say that the equation of motion holds as an operator equation. This is a technique that we will use extensively: an operator equation is one which holds when inserted into any path integral that has no other fields coincident with those involved in the equation. Translating to Hilbert space language, we can use the other insertions Image to set up arbitrary initial and final states, so this corresponds to the ordinary notion that an operator equation is one that should hold for general matrix elements.

When additional insertions are present at t, operator equations inside the path integral hold only up to additional delta-function source terms. For example, let Image = q(t)…, where t is near t and the other fields in ‘…’ are far away. Then eq. (A.1.19) becomes

Image

This is an example of the Schwinger–Dyson equation, the equation of motion for the expectation value of a product of fields. The same result written as an operator equation is

Image

As an explicit example, take the harmonic oscillator,

Image

Then the operator equations we have derived are

Image

and

Image

In Hilbert space language these become

Image

The first is the Heisenberg equation for the field operator, while the delta function in the second comes from the time derivative acting on the step functions in the time-ordered product,

Image

Euclidean path integrals

In Schrödinger language, the time-ordered product (A.1.17) is

Image

where t> is the greater of t and t, and t< is the lesser. By the nature of the path integral, which builds amplitudes from successive time steps, the time differences multiplying H are always positive in amplitudes such as (A.1.27): one only encounters expImage for positive Δt. Now continue analytically by uniformly rotating all times in the complex plane, t → eiImaget for any phase 0 ≤ ImageImage. The real part of the exponent is −E sinImageΔt, where E is the energy of the intermediate state. Because the Hamiltonian is bounded below for a stable system, the sum over intermediate states is safely convergent and no singularities are encountered — the continued amplitude is well defined.

Taking the times to be purely imaginary, t = −iu for real u, defines the Euclidean amplitudes, as opposed to the Minkowski amplitudes studied above:

Image

is defined to be the continuation of

Image

to t = −iu. One can repeat the derivation of the path integral representation for the Euclidean case, but the result is extremely simple: just replace t in the action with −iu,

Image

The weight in the Euclidean path integral is conventionally defined to be exp(−SE), so the Euclidean action contains a minus sign,

Image

Euclidean path integrals are usually the more well-defined, because the exponential of the action is damped rather than oscillatory. The Minkowski matrix elements can then be obtained by analytic continuation. In the text our main concern is the Euclidean path integral, which we use to represent the string S-matrix.

One is often interested in traces, such as

Image

where the sum runs over all eigenvectors of Image and the Ei are the corresponding eigenvalues. In the q basis the trace becomes

Image

In the last line, [dq]P denotes the integral over all periodic paths on [0, U]. An interesting variation on this is the integral over all antiperiodic paths:

Image

Here Image is the reflection operator, taking q to −q. We see that twisting the boundary conditions — for example, replacing periodic with antiperiodic — is equivalent to inserting an operator into the trace. Since Image2 = 1, its eigenvalues are ±1. If we take the average of the periodic and antiperiodic path integrals, this inserts Image(1 + Image), which projects onto the Image = +1 space:

Image

In the same way, Image(periodic − antiperiodic) gives the trace on the Image = −1 space.

A minor but sometimes confusing point: if Image is a Hermitean operator, the Euclidean Heisenberg operator Image(u) = expImage exp(−Imageu) satisfies

Image

and so is not in general Hermitean. For this reason it is useful to define a Euclidean adjoint,

Image

A Hermitean operator in the Schrödinger picture becomes Euclidean self-adjoint in the Euclidean Heisenberg picture.

Diagrams and determinants

The diagrammatic expansion is easily obtained from the path integral. This subject is well treated in modern field theory texts. We will only derive here a few results used in the text. For notational clarity we switch to field theory, replacing q(t) with a real scalar field Image(t, x). Begin with the free action

Image

where Δ is some differential operator. Consider the path integral with a source J(images), taking the Euclidean case to be specific,

Image

This is the generating functional for Euclidean Green’s functions. Taking functional derivatives with respect to J(images) pulls down factors of iImage(images) into the integral. Shifting Image to Image + iΔ−1J completes the square, leaving

Image

In the second line, each Image derivative becomes iJ. Any functional Image of Image can be written

Image

where Image [J] is the (functional) Fourier transform of Image. By linearity one obtains

Image

Expanding the exponential generates the diagrammatic expansion, all ways of contracting pairs of fields in Image. Also, when Image is a product of factors, Image, we can separate

Image

where Image/ImageImagei acts only on Imagei. The self-contractions are then the terms

Image

Consider now the overall factor Z [0]. We assume that Δ is self-adjoint and so we expand Image(images) in a complete set of eigenfunctions Imagei(images):

Image

where

Image

Define the path integral measure by

Image

Then

Image

The product of eigenvalues must be suitably regulated and renormalized. The value of Image1Image is usually given simply as (detΔ)−1/2, because a rescaling of Δ is equivalent to a rescaling of Image, the Jacobian for which can be absorbed into the bare couplings. (In free field theory, the only ‘coupling’ that would be renormalized would be a c-number constant in the Lagrangian density; we will see this appearing in the example.)

All of the above discussion extends readily to multicomponent qi or Imagei. One case of interest is a complex scalar field. One can break this up into two real scalars, Image = 2−1/2(Image1 + iImage2), and the measure is [dImage dImage*] ≡ [dImage1 dImage2]. For complex fields, (A.1.48) becomes

Image

One similarly finds for two real fields

Image

This is a functional version of the delta-function integral

Image

An example

As an example, we study the Euclidean transition amplitude for the harmonic oscillator,

Image

where Sct is a cutoff-dependent piece to be determined later. To evaluate this, separate out the ‘classical’ part of q,

Image

where

Image

The boundary conditions on q are then

Image

and the action separates

Image

Here

Image

The amplitude becomes

Image

Expand in a complete set of eigenfunctions of Image subject to the boundary condition (A.1.55),

Image

The infinite product diverges. The Pauli–Villars regulator is a simple way to define it. Divide by the amplitude for a regulator oscillator of very high frequency Ω,

Image

For j small, the terms in (A.1.60) are the same as those in (A.1.59b) up to a multiplicative constant, while for j large the product converges. The value of the product (A.1.60) is

Image

This follows from a standard infinite product representation for the sinh. One can guess the result (A.1.61) by noting that it has the right poles and zeros in U, and the right value as U goes to zero.

Keeping only terms that are nonvanishing as Ω → ∞,

Image

To get a finite answer as Ω → ∞, we need first to include a term ImageΩ in the Lagrangian Lct, canceling the linear divergence in (A.1.62). That is, there is a linearly divergent bare coupling for the operator ‘1’. It may seem strange that we need to renormalize in a quantum mechanics problem, but the power counting is completely uniform with quantum field theory. The logarithmic divergence is a wavefunction renormalization. To determine the finite normalization, compare the path integral expression (A.1.62) to the canonical result as U → ∞. The U → ∞ amplitude is easily evaluated by inserting a complete set of states, since only the ground state |0Image contributes in the limit:

Image

On the other hand, the path integral expression (A.1.62) becomes

Image

Thus

Image

and

Image

which is the correct result.

A.2    Fermionic fields

Consider the basic fermionic quantum system, two states |↑Image and |Image with raising operator Image and lowering operator Image:

Image

Amplitudes will be expressed as path integrals over the classical analogs of Image and Image. These are anticommuting c-numbers or Grassmann variables, elements Imagem and Imagem of a Grassmann algebra

Image

There is a natural notion of integration over Grassmann variables, the Berezin integral. Notice that if Image is a Grassmann variable then Image2 = 0, and so the most general function of Image is of the form a + Imageb. Define

Image

Integrals of more general expressions are determined by requiring the integral to be linear and by defining dImage to anticommute with all Grassmann variables. The significance of the definition (A.2.3) is that the integral of a total derivative is zero:

Image

For multiple Grassmann variables the integral is

Image

where c is the coefficient of the highest term in the Taylor expansion of Image,

Image

Looking back on the discussion of the bosonic path integral, one sees that all of the properties of ordinary integrals that were used are also true for the Berezin integral: linearity, invariance (except for a sign) under change of order of integration, and integration by parts.3

In order to follow the bosonic discussion as closely as possible, it is useful to define states that are formally eigenstates of Image:

Image

Define ImageImage| to satisfy

Image

The right-hand side is the Grassmann version of the Dirac delta function. It has the same properties as the ordinary delta function,

Image

and

Image

for arbitrary Image (Image). The state ImageImage| is a left eigenstate with opposite sign,

Image

as one verifies by taking the inner product of both sides with a general state |ImageImage. Note also that

Image

One can verify this by taking the matrix element between arbitrary states ImageImage′| and |ImageImage and evaluating the Grassmann integral.

The delta function has an integral representation

Image

By expanding the inner product (A.2.8) in powers of Image one obtains

Image

and therefore

Image

Using the further property that |ImageImage is an eigenstate of Image, one obtains

Image

This holds for any function Image with the prescription that it is written with the Image on the right.

We can now emulate the bosonic discussion, using the completeness relation (A.2.12) to join together successive time steps, with the result

Image

The inconvenient sign in the integral representation has been offset by reversing the order of Grassmann integrations. A typical fermionic Hamiltonian is Image = mImageImage. The corresponding Lagrangian is L = iImageImageH. For such first order Lagrangians the Hamiltonian and Lagrangian formulations coincide, and the result (A.2.17) is also the Lagrangian path integral.

There are two useful inner products for this system,

Image

Inner product (A) is positive definite and gives Image. It is the inner product for an ordinary physical oscillator. Inner product (B) is not positive, and it makes Image and Image Hermitean. It is the appropriate inner product for the Faddeev–Popov ghost system, as discussed in chapter 4. The derivation of the path integral was independent of the choice of inner product, but the relations (A.2.14) determine the expansion of ImageImage| in basis states in terms of the inner product.

All of the discussion of the bosonic case applies to the fermionic path integral as well. The derivation is readily extended to multiple Fermi oscillators and to field theory. Renormalization theory works in the same way as for the bosonic path integral. Fermionic path integrals again satisfy cutting relations, which simply amount to taking one of the iterated integrals (A.2.17) and saving it for the end. Insertion of Image(t) and Image(t) in the path integral again gives the corresponding operator in the matrix element, and multiple insertions are again time-ordered. The path integral insertion

Image

becomes the operator

Image

Note that a minus sign appears in the definition of time-ordering. The equations of motion and the Euclidean continuation work as in the bosonic case, but there is one important change. Consider a general operator Image,

Image

where i, j run over ↓ and ↑. Then by expanding Image|ImageImage in terms of the basis states one finds

Image

where the fermion number Image has been defined to be even for |Image and odd for |↑Image. In particular,

Image

In contrast to the bosonic results (A.1.33) and (A.1.34), the periodic path integral gives the trace weighted by (–1)Image and the antiperiodic path integral gives the simple trace.

Feynman diagrams are as in the bosonic case, but the determinants work out differently. Consider

Image

where we have again switched to the notation of field theory. Expand Image in eigenfunctions of Δ, and Image in eigenfunctions of ΔT,

Image

Define [dImagedImage] = Πi dImageidImagei. Then

Image

Similarly, one finds for a single field Image that

Image

where Δ must be antisymmetric. These fermionic path integrals are just the inverse of the corresponding bosonic integrals.

For finite-dimensional integrals we have to be careful about 2Images,

Image

Curiously, we can be more cavalier for functional determinants, where these can be eliminated by rescaling the fields. Bosonic path integrals actually give the absolute values of determinants, while fermionic path integrals give the determinant with a sign that depends on the precise order of Grassmann factors in the measure. We will fix the sign by hand.4

If Δ has zero eigenvalues, the path integral vanishes in the fermionic case and diverges in the bosonic case. When this occurs there is always a good reason, as we will see in practice, and the integrals of interest will have appropriate insertions to give a finite result.

A useful result about Berezin integrals: from the definition it follows that if Image′ = Imagea then dImage′ = a−1dImage. That is, the Jacobian is the inverse of the bosonic case. More generally, with bosonic variables imagesi and Grassmann variables Imagej, under a general change of variables to Image the infinitesimal transformation of the measure has an extra minus sign,

Image

Exercises

A.1 (a) Evaluate the path integral (A.1.33) for the harmonic oscillator by expanding [dq]P in a set of periodic eigenfunctions (not by inserting the result already obtained for the transition function into the left-hand side). Compare with an explicit evaluation of the sum over states (A.1.32).

(b) Do the same for the antiperiodic path integral (A.1.34).

A.2 Consider two harmonic oscillators of the same frequency, and form the complex coordinate q = 2−1/2(q1 + iq2). Evaluate the following path integral on [0, U] by using a suitable mode expansion:

Image

where the integral is to run over all paths such that q(U) = q(0)eiImage. Again compare with the answer obtained by summing over states: it is useful here to form complex combinations of oscillators, a1 ± ia2. This is a more exotic version of exercise A.1, and arises in various string theories such as orbifolds.

A.3 Consider a free scalar field X in two dimensions, with action

Image

Evaluate the path integral with both coordinates periodic,

Image

[Expanding the spatial dependence in Fourier modes, the answer takes the form of an infinite product of harmonic oscillator traces.]

A.4 Consider a free particle on a circle,

Image

By means of path integrals, evaluate Image. Be careful: there are topologically inequivalent paths, distinguished by the number of times they wind around the circle.

A.5 The fermionic version of exercise A.1: evaluate the fermionic path integral (A.2.23) and the corresponding antiperiodic path integral by using a complete set of periodic or antiperiodic functions, and compare with the sum over states.

__________

1  As another example, one can freely rescale fields. The Jacobian is formally a product of separate Jacobians at each point; its logarithm is therefore a sum of a local expression at each point. Regulated, this is just a shift of the bare couplings.

2  The functional derivative is defined as

Image

It can be regarded as the limit of the ordinary derivative

Image

In particular, one can integrate a functional derivative by parts, as for an ordinary derivative.

3  It is not useful to think of Image dImage as an integral in any ordinary sense: it is simply a linear mapping from functions of Image to the complex numbers. Since it shares so many of the properties of ordinary integrals, the integral notation is a useful mnemonic.

4  A fine point: when the world-sheet theory is parity-asymmetric, even the phase of the path integral is ambiguous. The point is that Δ is not Hermitian in this case, because the Euclidean adjoint includes a time-reversal (which is the same as a parity flip plus Euclidean rotation). One must instead take the determinant of the Hermitian operator ΔImageΔ, which determines that of Δ only up to a phase. We will encounter this in section 10.7, for example, where we will have to check that there is a consistent definition of the phase.