Becker, Becker, Schwarz: String theory and M-theory
Tong: String Theory (online lecture notes)
Green, Schwarz, Witten: Superstring Theory (Volume 1)
Polchinski: Joe's Little Book of String (Polchinski's online lecture notes)
Srednicki: Quantum Field Theory
This section is based on Zwiebach Chapter 1. The Standard Model describes four fundamental forces (strong, weak, electromagnetic, gravity). The first three give consistent quantum theories, but quantum Einstein gravity is plagued by divergences and is believed to break down at sufficiently small distances/high energies given by
where G=G= Newton's constant, ℏ=\hbar= Planck's constant, c=c= speed of light. Planck length l_(p)=l_{p}= is the unique length that can be constructed using only powers of G,ℏ,cG, \hbar, c. Similarly Planck mass m_(p)=m_{p}= is unique mass that can be constructed in this way.
String theory provides a mathematically consistent theory of quantum gravity (no UV divergences). It also unifies gravity with the other forces. Roughly speaking, gravity comes from fluctuations of closed strings, while gauge theories (like electromagnetism) come from fluctuations of open strings. Since open strings can close to form closed strings, it's not possible to have a theory with only open strings. Hence, gravity is required by string theory.
String theory is also important theoretically because it lead to the discovery of supersymmety and the AdS/CFT correspondence. The latter relates quantum gravity to a lower-dimensional non-gravitational quantum theory. On the other hand making experimentally testable predictions remains a challenge for string theory. This may require a better understanding of its conceptual foundations. The goal of this course is to understand the emergence gravity, its unification with gauge theory, and its good high energy behaviour in the context of bosonic string theory. Supersymmetry will be described next term.
2 Basics
In this section we will review some basic concepts:
Special relativity
Field theory: Klein-Gordon, Maxwell
General relativity
2.1 Special Relativity
The principle of Relativity states that equations of motion should be the same in all frames. Frames are coordinate systems that observers use to measure locations of events in space and time, which are tied to how the observer is moving. There is a special set of frames in which objects follow linear trajectories (at least locally). These are known as inertial frames. Physically, these arise in the absence of gravity and acceleration. Even in the presence of gravity, it is possible for an observer to have a locally inertial frame by going to a state of free-fall. Special Relativity describes motion in inertial frames. General Relativity describes motion in general frames and explains how gravity arises from the curvature of spacetime.
where x^(0)x^{0} is time and x^(1,2,3)x^{1,2,3} are spatial directions. Suppose an inertial observer measures two events at x^(mu)x^{\mu} and x^(mu)+Deltax^(mu)x^{\mu}+\Delta x^{\mu}. The distance in spacetime between these two events is given by
Now suppose that another inertial observer measures the events to be at x^('mu)x^{\prime \mu} and x^('mu)+Deltax^('mu)x^{\prime \mu}+\Delta x^{\prime \mu}. The spacetime distance they measure is the same:
Hence DeltaS^(2)\Delta S^{2} is an invariant interval.
Deltas^(2) > 0\Delta s^{2}>0 time-like
Deltas^(2) < 0quad\Delta s^{2}<0 \quad spacelike
Deltas^(2)=0\Delta s^{2}=0
If Deltas^(2)=0\Delta s^{2}=0, then the two events can be separated by a light ray. Since the distance between the two events is invariant, the speed of light is the same in all frames.
For events that are infinitesimally close together, we denote the invariant interval as
where LL is a Lorentz transformation and TT is a translation. This change of coordinatess leaves the metric invariant:
{:[ds^(2)=n_(alpha beta)dx^(alpha)dx^(beta)],[=n_(mu nu)dx^(mu)dx^(nu)],[=n_(mu nu)(delx^('mu))/(delx^(alpha))(delx^(nu))/(delx^(beta))dx^(alpha)dx^(beta)],[=eta_(mu nu)L_(alpha)^(mu)L_(beta)^(nu)dx^(alpha)dx^(beta)],[ rarreta_(mu nu)L_(alpha)^(mu)L_(beta)^(nu)=eta_(alpha beta)]:}\begin{aligned}
& d s^{2}=n_{\alpha \beta} d x^{\alpha} d x^{\beta} \\
& =n_{\mu \nu} d x^{\mu} d x^{\nu} \\
& =n_{\mu \nu} \frac{\partial x^{\prime \mu}}{\partial x^{\alpha}} \frac{\partial x^{\nu}}{\partial x^{\beta}} d x^{\alpha} d x^{\beta} \\
& =\eta_{\mu \nu} L_{\alpha}^{\mu} L_{\beta}^{\nu} d x^{\alpha} d x^{\beta} \\
& \rightarrow \eta_{\mu \nu} L_{\alpha}^{\mu} L_{\beta}^{\nu}=\eta_{\alpha \beta}
\end{aligned}
In terms of matrices, this can be written as L^(T)eta L=etaL^{T} \eta L=\eta. Lorentz transformations can be thought of as rotations in spacetime. For example a rotation in the x^(1)-x^(2)x^{1}-x^{2} plane takes in the form
In this case, ss is called the proper time. It corresponds to the time elapsed in the rest frame of the particle; where dx^(1)=dx^(2)=dx^(3)=0d x^{1}=d x^{2}=d x^{3}=0 so, ds^(2)=(dx^(0))^(2)d s^{2}=\left(d x^{0}\right)^{2}. We then define the 4 -velocity as
Since dsd s is Lorentz -invariant and dx^(mu)d x^{\mu} is a Lorentz Vector, u^(mu)u^{\mu} also transforms like a vector under coordinate transformations. More explicitly, it can be written as
p^(0)=m+(1)/(2)mbeta^(2)+dotsp^{0}=m+\frac{1}{2} m \beta^{2}+\ldots
which we recognise as the rest energy plus the the kinetic energy. Hence p^(0)=Ep^{0}=E (energy). In general, the energy is related to the mass an spatial momentum as:
This will be based on Zwiebach 10.2-10.4, Srednicki Chapter 3. First we consider scalar (or spin-0) fields. This is described by the Klein-Gordon equation:
where eta^(mu nu)del_(mu)del_(nu)phi=(-del_(0)^(2)+ vec(grad)^(2))phi=del^(2)phi\eta^{\mu \nu} \partial_{\mu} \partial_{\nu} \phi=\left(-\partial_{0}^{2}+\vec{\nabla}^{2}\right) \phi=\partial^{2} \phi We previously showed that the Laplacian is Lorentz invariant we see that the equation of motion is the same in all frames.
Let us Fourier transform expand the scalar field:
phi(x)=int(d^(D)p)/((2pi)^(D))e^(ip*x)phi(p)\phi(x)=\int \frac{d^{D} p}{(2 \pi)^{D}} e^{i p \cdot x} \phi(p)
where DD is spacetime dimension, and we recall that
If phi(x)\phi(x) is real, it has one propagating degree of freedom. If phi(x)\phi(x) is complex, it has two real propagating degrees of freedom. Let's take it to be real. Taking the complex conjugate,
when we took vec(p)rarr- vec(p)\vec{p} \rightarrow-\vec{p} in the second line. From this, we see that phi(x)=\phi(x)=phi^(**)(x)longleftrightarrow b( vec(p))=a^(**)(- vec(p))\phi^{*}(x) \longleftrightarrow b(\vec{p})=a^{*}(-\vec{p}). Plugging this back in finally gives
This can be derived using a procedure called "canonical quantization". We wont go into the details, but physically one can think of the quatum field phi(x)\phi(x) as an infinite set of harmonic oscillations, one for each Fourier mode. For more details, see see section 10.4 of Zwieback and chapter 3 of Srednicki.
vacuum state satifies a( vec(p))|0:)=0a(\vec{p})|0\rangle=0.
a^(†)( vec(p))|0:)a^{\dagger}(\vec{p})|0\rangle is a single particle state with momentum vec(p)\vec{p} and energy E_(p)E_{p}.
a^(†)( vec(p)_(1))dotsa^(†)( vec(p)_(n))|0:)a^{\dagger}\left(\vec{p}_{1}\right) \ldots a^{\dagger}\left(\vec{p}_{n}\right)|0\rangle is an nn-particle state. The particles have momenta vec(p)_(1),dots, vec(p)_(n)\vec{p}_{1}, \ldots, \vec{p}_{n} and energies E_(p_(1),dots,),E_(p_(n))E_{p_{1}, \ldots,}, E_{p_{n}}.
2.2.2 Maxwell
This will be based on Zwiebach 3.1,3.3,10.5. Maxwell's equations for electromagnetism:
so the equations have the same form in all inertial frames. Moreover, we see how vec(E)\vec{E} and vec(B)\vec{B} mix into eachother under Lorentz transformations.
The fields can be derived from a vector potential. To see this first note that
Hence, we see that electromagnetism is described by a spin-1 field A^(mu)A^{\mu}. From this form of F_(mu nu)F_{\mu \nu} it automatically follows that del_([mu)F_(nu lambda])=0\partial_{[\mu} F_{\nu \lambda]}=0.
Setting J^(mu)=0J^{\mu}=0, the remaining Maxwell equations can be written as
where I=2,dots,D-1I=2, \ldots, D-1 "transverse directions." Setting mu=I\mu=I, the equations of motion then reduce to
p^(2)A^(I)(p)=0p^{2} A^{I}(p)=0
which are just the equations of motion for (D-2)(D-2) massless scalars! Hence, we that the Maxwell field has D-2 independent degrees of freedom (one for each transverse degree of freedom).
A^(-)A^{-}is not an independent degree of freedom because it can be written in terms of A^(I)A^{I}. We may quantize A^(I)A^{I} as we did the scalar field, writing them in terms of creation and annihilation operators. The quanta of A^(I)A^{I} are called photons. A general single photon state is given by
where epsilon_(I)\epsilon_{I} is a polarisation vector.
2.3 General Relativity
This will mostly follow Zwiebach 3.6, 10.6. As mentioned previously, GR describes motion in general frames. Recall that if we start in an inertial frame and perform a coordinate transformation, the metric will become
where x^(mu)x^{\mu} are the coordinates of the inertial frame and x^('mu)x^{\prime \mu} are the coordinates of the new frame, respectively. If delx^(mu)//delx^('rho)\partial x^{\mu} / \partial x^{\prime \rho} is a Lorentz transformation, then g_(mu nu)=eta_(mu nu)g_{\mu \nu}=\eta_{\mu \nu}. For more general coordinate transformations, g_(mu nu)g_{\mu \nu} will not be the Minkowski metric. Physically, this can arise from going to an accelerating frame.
There is another way to have a non-trivial metric: turning on gravity. In this case it will not be possible to transform the metric to Minkowski (except locally) via coordinate transformation. This is because gravity arises from the curvature of spacetime, which can't be set to zero by a coordinate transformation. Minkowski has zero curvature and is therefore a flat spacetime.
Hence, the invariant interval is generally given by
-ds^(2)=g_(mu nu)(x)dx^(mu)dx^(nu)-d s^{2}=g_{\mu \nu}(x) d x^{\mu} d x^{\nu}
Under general coordinate transformation (or diffeomorphism)
You will show this in the homework. This is the gravitational analogue of a gauge transformation that we saw in Maxwell theory.
In the presence of gravity, the metric becomes dynamical and its equation of motion is the Einstein equation which states that the curvature of spacetime can be derived from the distribution of matter and energy. The Einstein equations are invariant under diffeomorphisms. We will not write them out in general but instead consider small fluctuations around Minkowski (i.e. weak gravity):
The equations of motion are invariant under this variation. To see this, let's Fourier transform to momentum space, which amounts to the replacement del_(mu)rarr\partial_{\mu} \rightarrow+ip_(mu)+i p_{\mu}. The equations of motion then become
Note that h^(I-),h^(--)h^{I-}, h^{--}are determined by h^(IJ)h^{I J} so are not independent degrees of freedom.
Hence, h^(IJ)h^{I J} can be treated like massless scalars, and h^(-I)h^{-I} and h^(--)h^{--}are determined from these. Since h^(IJ)h^{I J} is a symmetric traceless (D-2)xx(D-2)(D-2) \times(D-2) matrix, the number of degrees of freedom in the gravitational field is
For D=4D=4, we see that gravity has 2 degrees of freedom. After quantisation, we can once again write h^(IJ)h^{I J} in terms of creation and annihilation operators. The quanta of the gravitational field are called gravitons. One-graviton states are given by
where epsilon_(IJ)\epsilon_{I J} is a symmetric traceless tensor.
3 Relativistic Point Particle
This will primarily follow Tong Section 1.1. In non-relativistic quantum mechanics, spatial coordinates X^(i)X^{i} are operators white X^(0)X^{0} is a label. On the other hand, in relativity space and time are on equal footing so we have two choices: to demote spatial coordinates to labels and introduce fields giving QFT, or to promote time to an operator. The latter approach is less convenient in practice for the study of point particles, but generalises more straightforwardly to string theory.
Alter promoting X^(0)X^{0} to and operator, we must introduce a new parameter tau\tau to parameterize the world line. One then integrates over all paths between initial and final spacetime points. The action should be Lorentz-invariant and extremised by the classical trajectory. The natural candidate is the proper time of the particle's trajectory:
S=-m int dsS=-m \int d s
where
{:[ds^(2)=-dX^(mu)dX^(nu)eta_(mu nu)],[=-X^(˙)^(mu)X^(˙)^(nu)eta_(mu nu)dtau^(2)","quadX^(˙)^(mu)=del_(tau)X^(mu)]:}\begin{aligned}
d s^{2} & =-d X^{\mu} d X^{\nu} \eta_{\mu \nu} \\
& =-\dot{X}^{\mu} \dot{X}^{\nu} \eta_{\mu \nu} d \tau^{2}, \quad \dot{X}^{\mu}=\partial_{\tau} X^{\mu}
\end{aligned}
X^(mu)(tau)X^{\mu}(\tau), encodes the embedding of the world line into the "target space," which we take to be Minkowski space.
Hence the action for a relativistic point particle is
S=-m int d sigmasqrt(X^(˙)^(mu)X^(˙)^(v)eta_(mu v))S=-m \int d \sigma \sqrt{\dot{X}^{\mu} \dot{X}^{v} \eta_{\mu v}}
This is manifestly invariant under Poincare transformations of target space:
tau rarr tilde(tau)(tau)\tau \rightarrow \tilde{\tau}(\tau)
Indeed, we see that
{:[S=-m int d tilde(tau)sqrt(-del_( tilde(tau))X^(mu)del_( tilde(tau))X^(nu)eta_(mu nu))],[=-m int d tau(del( tilde(tau)))/(del tau)sqrt(-((del tau)/(del( tilde(tau)))del_(tau)X^(mu))((del tau)/(del( tilde(tau)))del_(tau)X^(nu))eta_(mu nu))],[=-m int d tausqrt(-del_(tau)X^(mu)del_(tau)X^(nu)eta_(mu nu))]:}\begin{aligned}
S & =-m \int d \tilde{\tau} \sqrt{-\partial_{\tilde{\tau}} X^{\mu} \partial_{\tilde{\tau}} X^{\nu} \eta_{\mu \nu}} \\
& =-m \int d \tau \frac{\partial \tilde{\tau}}{\partial \tau} \sqrt{-\left(\frac{\partial \tau}{\partial \tilde{\tau}} \partial_{\tau} X^{\mu}\right)\left(\frac{\partial \tau}{\partial \tilde{\tau}} \partial_{\tau} X^{\nu}\right) \eta_{\mu \nu}} \\
& =-m \int d \tau \sqrt{-\partial_{\tau} X^{\mu} \partial_{\tau} X^{\nu} \eta_{\mu \nu}}
\end{aligned}
We can use this symmetry to set tilde(tau)=X^(0)(tau)\tilde{\tau}=X^{0}(\tau) (this is know as static gauge), so that
S=int d tausqrt(1- vec(X)^(˙)^(2))S=\int d \tau \sqrt{1-\dot{\vec{X}}^{2}}
where we relabeled tilde(tau)\tilde{\tau} as tau\tau. Lorentz-invariance is no longer manifest. Let's expand the static gauge action in the non-relativistic limit:
{:[S=int d tausqrt(1- vec(X)^(˙)^(2))],[=-m int d tau(1-(1)/(2) vec(X)^(˙)^(2)+dots)]:}\begin{aligned}
& S=\int d \tau \sqrt{1-\dot{\vec{X}}^{2}} \\
& =-m \int d \tau\left(1-\frac{1}{2} \dot{\vec{X}}^{2}+\ldots\right)
\end{aligned}
where ... indicates higher order terms in vec(X)^(˙)\dot{\vec{X}}. The first term is the rest-mass energy and the second term is the non-relativistic kinetic energy.
We will now compute the equations of motion and quantize using the manifestly Lorentz-invanaut action. Let's compute the equations of motion by extremising the action. Consider the variation X^(mu)(tau)rarrX^(mu)(tau)+deltaX^(mu)(tau)X^{\mu}(\tau) \rightarrow X^{\mu}(\tau)+\delta X^{\mu}(\tau) :
{:[delta S=int d tau deltaL(X^(˙)^(mu)(tau))],[=int d tau(delL)/(delX^(˙)^(mu))deltaX^(˙)^(mu)=int d tau[del_(tau)((delL)/(delX^(˙)^(mu))deltaX^(mu))-del_(tau)((delL)/(delX^(˙)^(mu)))deltaX^(mu)]],[=(delL)/(delX^(˙)^(mu))deltaX^(mu)|_(tau=tau_(i))^(tau=tau_(f))-int d taudel_(tau)((delL)/(delX^(˙)^(mu)))deltaX^(mu)],[=-int d taudel_(tau)((delL)/(delX^(˙)^(mu)))deltaX^(mu)]:}\begin{aligned}
& \delta S=\int d \tau \delta \mathcal{L}\left(\dot{X}^{\mu}(\tau)\right) \\
& =\int d \tau \frac{\partial \mathcal{L}}{\partial \dot{X}^{\mu}} \delta \dot{X}^{\mu}=\int d \tau\left[\partial_{\tau}\left(\frac{\partial \mathcal{L}}{\partial \dot{X}^{\mu}} \delta X^{\mu}\right)-\partial_{\tau}\left(\frac{\partial \mathcal{L}}{\partial \dot{X}^{\mu}}\right) \delta X^{\mu}\right] \\
& =\left.\frac{\partial \mathcal{L}}{\partial \dot{X}^{\mu}} \delta X^{\mu}\right|_{\tau=\tau_{i}} ^{\tau=\tau_{f}}-\int d \tau \partial_{\tau}\left(\frac{\partial \mathcal{L}}{\partial \dot{X}^{\mu}}\right) \delta X^{\mu} \\
& =-\int d \tau \partial_{\tau}\left(\frac{\partial \mathcal{L}}{\partial \dot{X}^{\mu}}\right) \delta X^{\mu}
\end{aligned}
where we take deltaX^(mu)(tau_(i))=deltaX^(mu)(tau_(f))=0\delta X^{\mu}\left(\tau_{i}\right)=\delta X^{\mu}\left(\tau_{f}\right)=0. Hence, we find that
where Pi_(mu)=(delL)/(delX^(˙)^(mu))\Pi_{\mu}=\frac{\partial \mathcal{L}}{\partial \dot{X}^{\mu}} is the canonical momentum.
In the present case L=-msqrt(-X^(˙)^(2)),X^(˙)^(2)=X^(˙)^(mu)X^(˙)^(nu)eta_(mu nu)\mathcal{L}=-m \sqrt{-\dot{X}^{2}}, \dot{X}^{2}=\dot{X}^{\mu} \dot{X}^{\nu} \eta_{\mu \nu}. Hence,
Hence, particles follow linear trajectories which satisfy the mass shell constraint, as expected.
To quantize, promote Pi_(mu)\Pi_{\mu} to the operator Pi_(mu)=-i del//delX^(mu)\Pi_{\mu}=-i \partial / \partial X^{\mu}. Then we have the canonical quantisation condition
where Psi\Psi is a wavefunction. Hence we recover the Klein-Gordon equation we saw previously for scalar fields. Note that Psi(x)\Psi(x) is not a field, however. In this formalism we can only define a single-particle state which is described by the wavefunction. In contrast, in QFT we can construct multiparticle states. Similarly, when we generalise this approach to string theory, we will only be able to construct single-panticle states. To define multi-particle states in string theory, we need string field theory, which is beyond the scope of this course.
4 Nambu - Goto Action
This will follow Tong section 1.2 and BBS section 2.2. Whereas a particle sweeps out a world line in spacetime, a string sweeps out a 2d2 \mathrm{~d} surface called a world sheet. Points on the worldsheet are parameterised by two worldsheet coordinates sigma^(alpha)=(sigma^(0),sigma^(1))=(tau,sigma).X^(mu)(tau,sigma)\sigma^{\alpha}=\left(\sigma^{0}, \sigma^{1}\right)=(\tau, \sigma) . X^{\mu}(\tau, \sigma) describes the embedding of the world sheet into spacetime.
Note that BBS takes sigma in[0,pi)\sigma \in[0, \pi) but we will follow the conventions of Tong.
Whereas the action for a relativistic point particle is proportional to the length of the worldline, the action of a string should be proportional to the area of the worldsheet. Hence, the classical string motion extremizes the area. To compute the area of the 2d2 \mathrm{~d} surface swept out by the string, first compute the induced metric on the surface:
" Area "=intd^(2)sigmasqrt(-det gamma)\text { Area }=\int d^{2} \sigma \sqrt{-\operatorname{det} \gamma}
Noting that quadgamma_(alpha beta)=([x^(˙)^(2),X^(˙)*X^(')],[X^(˙)*x^('),X^('2)])\quad \gamma_{\alpha \beta}=\left(\begin{array}{cc}\dot{x}^{2} & \dot{X} \cdot X^{\prime} \\ \dot{X} \cdot x^{\prime} & X^{\prime 2}\end{array}\right), where X^(˙)^(mu)=del_(tau)X^(mu),X^('mu)=del_(sigma)X^(mu)\dot{X}^{\mu}=\partial_{\tau} X^{\mu}, X^{\prime \mu}=\partial_{\sigma} X^{\mu} we see that
{:(4)" Area "=int d sigma d tausqrt(((X^(˙))*X^('))^(2)-X^(˙)^(2)X^('2)):}\begin{equation*}
\text { Area }=\int d \sigma d \tau \sqrt{\left(\dot{X} \cdot X^{\prime}\right)^{2}-\dot{X}^{2} X^{\prime 2}} \tag{4}
\end{equation*}
Let's verify this is the area when the target space in Euclidean (R^(D))\left(\mathbb{R}^{D}\right) rather than Minkowski space. Consider an infinitesimal region in sigma,tau\sigma, \tau space and let's compute the area of the corresponding region in the target space. The vectors tangent to the boundary of the region in the target space are:
d vec(l)_(1)=(del( vec(X)))/(del sigma)d sigma,quad d vec(l)_(2)=(del( vec(X)))/(del tau)d taud \vec{l}_{1}=\frac{\partial \vec{X}}{\partial \sigma} d \sigma, \quad d \vec{l}_{2}=\frac{\partial \vec{X}}{\partial \tau} d \tau
If the angle between the two vectors is theta\theta, then the area is
{:[d(" Area ")=|d vec(l_(1))||d vec(l)_(2)|sin theta=sqrt(d vec(l)_(1)^(2)d vec(l)_(2)^(2)(1-cos^(2)theta))=sqrt(d vec(l)_(1)^(2)d vec(l)_(2)^(2)-(d vec(l_(1))*d vec(l)_(2))^(2))],[=sqrt(X^(˙)^(2)X^('2)-((X^(˙))*X^('))^(2))d sigma d tau]:}\begin{aligned}
& d(\text { Area })=\left|d \overrightarrow{l_{1}}\right|\left|d \vec{l}_{2}\right| \sin \theta=\sqrt{d \vec{l}_{1}^{2} d \vec{l}_{2}^{2}\left(1-\cos ^{2} \theta\right)}=\sqrt{d \vec{l}_{1}^{2} d \vec{l}_{2}^{2}-\left(d \overrightarrow{l_{1}} \cdot d \vec{l}_{2}\right)^{2}} \\
& =\sqrt{\dot{X}^{2} X^{\prime 2}-\left(\dot{X} \cdot X^{\prime}\right)^{2}} d \sigma d \tau
\end{aligned}
which indeed takes the form of (4).
After this discussion we now present the Nambu-Goto action for a relativistic string:
S=-T int d sigma d tausqrt(((X^(˙))*X^('))^(2)-X^(˙)^(2)X^('2))S=-T \int d \sigma d \tau \sqrt{\left(\dot{X} \cdot X^{\prime}\right)^{2}-\dot{X}^{2} X^{\prime 2}}
where TT is the string tension. To understand physical interpretation of TT, let's take the non-relativistic limit. Using reparameteristion invariance sigma^(alpha)rarr\sigma^{\alpha} \rightarrowsigma^('alpha)(tau,sigma)\sigma^{\prime \alpha}(\tau, \sigma) (which you will prove in HW) we may choose "static gauge":
X^(0)=tau,X^(1)=sigmaX^{0}=\tau, X^{1}=\sigma
Note that in this gauge 0 < sigma < L0<\sigma<L, where LL is the spatial length of the string. Then
where I in{2,dots,D-1}I \in\{2, \ldots, D-1\} labels the transverse directions and ... indicate higherorder terms in X^(˙)^(I)\dot{X}^{I} or X^('I)X^{\prime I}, which we neglect in non-redativistic limit. In this limit,
{:[S=-T int d sigma d tausqrt(-det gamma)],[∼-T int d sigma d tausqrt(1-(X^(˙)^(I))^(2)+(X^('I))^(2))],[∼T int d tau d sigma[-1+(1)/(2)(X^(˙)^(I))^(2)-(1)/(2)(X^('I))^(2)]]:}\begin{aligned}
& S=-T \int d \sigma d \tau \sqrt{-\operatorname{det} \gamma} \\
& \sim-T \int d \sigma d \tau \sqrt{1-\left(\dot{X}^{I}\right)^{2}+\left(X^{\prime I}\right)^{2}} \\
& \sim T \int d \tau d \sigma\left[-1+\frac{1}{2}\left(\dot{X}^{I}\right)^{2}-\frac{1}{2}\left(X^{\prime I}\right)^{2}\right]
\end{aligned}
Let's compute the rest energy:
S_("rest ")=-T*int d tau d sigma=-m int d tau," where "m=LTS_{\text {rest }}=-T \cdot \int d \tau d \sigma=-m \int d \tau, \text { where } m=L T
This is the rest mass contribution to the action of a point-particle with mass LTL T. Hence, TT is mass per unit length and has dimensions ( length ) ^(-2){ }^{-2} (in units where ℏ=c=1)\hbar=c=1). It is conventional to define T=(1)/(2pialpha^('))T=\frac{1}{2 \pi \alpha^{\prime}}, where alpha^(')\alpha^{\prime} is known as "Regge slope" for reasons we will see later. It is also conventional to define alpha^(')=(1)/(2)l_(s)^(2)\alpha^{\prime}=\frac{1}{2} l_{s}^{2} where l_(s)l_{s} is the "string length".
5 Polyakov Action
This will mainly be based on BBS pg 26,27,30,3126,27,30,31. Although the Nambu-Goto action has a nice physical interpretation as the area of a string worldsheet, the
equations of motion are complicated and it is hard to quantize because of the square root. In practice, we use the Polyakov action:
Noting the deltasqrt(-h)=-(1)/(2)sqrt(-h)h_(alpha beta)deltah^(alpha beta)\delta \sqrt{-h}=-\frac{1}{2} \sqrt{-h} h_{\alpha \beta} \delta h^{\alpha \beta} (which you will provie on HW)
where T_(alpha beta)T_{\alpha \beta} is the "stress tensor". Hence action extremised (delta S=0)(\delta S=0) when T_(alpha beta)=0T_{\alpha \beta}=0.
In this case, we have
del_(alpha)X*del_(beta)X=(1)/(2)h_(alpha beta)h^(gamma delta)del_(gamma)X*del_(delta)X\partial_{\alpha} X \cdot \partial_{\beta} X=\frac{1}{2} h_{\alpha \beta} h^{\gamma \delta} \partial_{\gamma} X \cdot \partial_{\delta} X
Taking determinant of both sides (recall they are 2xx22 \times 2 matrices):
det(del_(alpha)X*del_(beta)X)=h((1)/(2)h^(gamma delta)del_(gamma)X*del_(delta)X)^(2)\operatorname{det}\left(\partial_{\alpha} X \cdot \partial_{\beta} X\right)=h\left(\frac{1}{2} h^{\gamma \delta} \partial_{\gamma} X \cdot \partial_{\delta} X\right)^{2}
Multiplying by (-1)(-1) and taking square rot:
sqrt(-det(del_(alpha)X*del_(beta)X))=(1)/(2)sqrt(-h)h^(gamma delta)del_(gamma)X*del_(delta)X\sqrt{-\operatorname{det}\left(\partial_{\alpha} X \cdot \partial_{\beta} X\right)}=\frac{1}{2} \sqrt{-h} h^{\gamma \delta} \partial_{\gamma} X \cdot \partial_{\delta} X
Recalling the Nambu-Goto action, we see that
{:[S=-T intd^(2)sigmasqrt(-det(del_(alpha)Xdel_(beta)X))],[=-(T)/(2)intd^(2)sigmasqrt(-h)h^(gamma)del_(gamma)X*del_(delta)X]:}\begin{aligned}
S & =-T \int d^{2} \sigma \sqrt{-\operatorname{det}\left(\partial_{\alpha} X \partial_{\beta} X\right)} \\
& =-\frac{T}{2} \int d^{2} \sigma \sqrt{-h} h^{\gamma} \partial_{\gamma} X \cdot \partial_{\delta} X
\end{aligned}
as claimed.
5.1 Symmetries and Gauge-fixing
The Polyakov action has the following symmeties:
Poincare transformations of target space: X^(mu)rarrX^(mu)=L_(nu)^(mu)X^(nu)+T^(mu)X^{\mu} \rightarrow X^{\mu}=L_{\nu}^{\mu} X^{\nu}+T^{\mu}
Note that h_(alpha beta)h_{\alpha \beta} is a symmetric 2xx22 \times 2 matrix so has three independent components. Since we have two reparameterisations (one for each sigma^(alpha)\sigma^{\alpha} ) we can fix two components of the metric, laving only one unfixed. Let's choose it to be conformally fiat:
This will be based on BBS pg 31-33. We already found the equations of motion for h_(alpha beta)h_{\alpha \beta}, which imply vanishing of stress tensor. In conformal gauge, this reduces to
{:[deltaS_(bdry)=-T int d tau d sigma[-del_(tau)(X^(˙)_(mu)deltaX^(mu))+del_(sigma)(X_(mu)^(')deltaX^(mu))]],[=-T int d tau d sigmadel_(sigma)(X_(mu)^(')deltaX^(mu))]:}\begin{aligned}
& \delta S_{b d r y}=-T \int d \tau d \sigma\left[-\partial_{\tau}\left(\dot{X}_{\mu} \delta X^{\mu}\right)+\partial_{\sigma}\left(X_{\mu}^{\prime} \delta X^{\mu}\right)\right] \\
& =-T \int d \tau d \sigma \partial_{\sigma}\left(X_{\mu}^{\prime} \delta X^{\mu}\right)
\end{aligned}
where we discarded boundary terms along the time direction.
Closed string:
deltaS_(bdry)=-T int d tau[X_(mu)^(')deltaX^(mu)|_(sigma=2pi)-X_(mu)^(')deltaX^(mu)|_(sigma=0)]=0\delta S_{b d r y}=-T \int d \tau\left[\left.X_{\mu}^{\prime} \delta X^{\mu}\right|_{\sigma=2 \pi}-\left.X_{\mu}^{\prime} \delta X^{\mu}\right|_{\sigma=0}\right]=0
since X^(mu)(tau,sigma)=X^(mu)(tau,sigma+2pi)X^{\mu}(\tau, \sigma)=X^{\mu}(\tau, \sigma+2 \pi).
Open string:
deltaS_(bdry)=-T int d tau[X_(mu)^(')deltaX^(mu)|_(sigma=pi)-X_(mu)^(')deltaX^(mu)|_(sigma=0)]\delta S_{b d r y}=-T \int d \tau\left[\left.X_{\mu}^{\prime} \delta X^{\mu}\right|_{\sigma=\pi}-\left.X_{\mu}^{\prime} \delta X^{\mu}\right|_{\sigma=0}\right]
In this case, boundary terms must vanish individually. This can be achieved with Dirichlet or Neumann boundary conditions:
Dirichlet: deltaX^(mu)=0\delta X^{\mu}=0 at sigma=0\sigma=0 or pi\pi
In other words, X^(mu)X^{\mu} is fixed to a constant at 0,pi0, \pi. This breaks Poincare invariance.
Neumann: X^(mu)=0X^{\mu}=0 at sigma=0\sigma=0 or pi\pi
This preserves Poincare invariance. We will only consider Neumann boundary conditions. Dirichlet boundary conditions will be considered next term.
5.3 Conserved Charges
This will be based on BBS pg 37,38, Srednicki Chapter 22. As shown earlier, Poincare transformations are symmetries of the string worldsheet theory:
For infinitesimal Lorentz transformations L^(mu)_(nu)=delta_(nu)^(mu)+deltaomega^(mu)_(nu)L^{\mu}{ }_{\nu}=\delta_{\nu}^{\mu}+\delta \omega^{\mu}{ }_{\nu}, we have
This symmetry implies that momentum and angular momentum of the string are conserved. The relation between symmetries and conserved quantities is given by Nether's theorem. Let's prove it for a general scalar field theory and then apply it to the string worldsheet theory.
Noether's theorem
Consider a set of scalar fields phi_(a)(x)\phi_{a}(x), with Lagrangian L(phi_(a),del_(mu)phi_(a))\mathcal{L}\left(\phi_{a}, \partial_{\mu} \phi_{a}\right). If L\mathcal{L} is invariant under the infinitesimal change deltaphi_(a)(x)\delta \phi_{a}(x), then the following current is classically conserved:
Q=ubrace(intd^(D-1)Xj^(-0)(x)ubrace)_("spatial intigral at fixed time ")Q=\underbrace{\int d^{D-1} X j^{-0}(x)}_{\text {spatial intigral at fixed time }}
is constant in time. This is known as the "Noether charge."
Proof
First let's compute the equations of motion. Consider an infinitesinal variation deltaphi_(a)\delta \phi_{a} (not necessarily a symmetry). Classical equations of motion determined by
demanding that delta S=0\delta S=0 :
where we used the Euler Lagrange equations to obtain the third equality. Hence, del_(mu)j^(mu)=0\partial_{\mu} j^{\mu}=0. This, proves Noether's theorem.
Now integrate del_(mu)j^(mu)\partial_{\mu} j^{\mu} over (D-1)(D-1)-dimensional space:
where we recall that deltaomega_(mu nu)=-deltaomega_(nu mu)\delta \omega_{\mu \nu}=-\delta \omega_{\nu \mu}. We have a Noether current for each independent component of deltaomega_(mu nu)\delta \omega_{\mu \nu} :
where we multiplied by 2 . The corresponding Noether charges are
J^(mu nu)=int d sigmaj_(0)^(mu nu)=T int d sigma(X^(mu)X^(˙)^(nu)-X^(nu)X^(˙)^(mu))J^{\mu \nu}=\int d \sigma j_{0}^{\mu \nu}=T \int d \sigma\left(X^{\mu} \dot{X}^{\nu}-X^{\nu} \dot{X}^{\mu}\right)
This encodes angular momentum of string.
6 Mode expansion
This will be based on Tong section 1.4, pg 52,53, and BBS pg 33-35,37,39,40. Recall the equations of motion
Let us introduce the lightcone cords: sigma^(+-)=tau+-sigma rarr tau=(1)/(2)(sigma^(+)+sigma^(-)),sigma=\sigma^{ \pm}=\tau \pm \sigma \rightarrow \tau=\frac{1}{2}\left(\sigma^{+}+\sigma^{-}\right), \sigma=(1)/(2)(sigma^(+)-sigma^(-))\frac{1}{2}\left(\sigma^{+}-\sigma^{-}\right). We then find
Hence, the general solution can be written as a sum of left and right-movers.
For closed strings, X^(mu)(tau,sigma)=X^(mu)(tau,sigma+2pi)X^{\mu}(\tau, \sigma)=X^{\mu}(\tau, \sigma+2 \pi). The most general periodic solution can be expanded in Fourier modes:
Recall that alpha^(')=(1)/(2)l_(s)^(2)\alpha^{\prime}=\frac{1}{2} l_{s}^{2}, as required by dimensional analysis. Normalisation of 1//n1 / n chosen for later convenience.
X_(L)^(mu),X_(R)^(mu)X_{L}^{\mu}, X_{R}^{\mu} not periodic because of terms linear in sigma ^(+-){ }^{ \pm}, but their sum gives (1)/(2)alpha^(')p^(mu)(sigma^(+)+sigma^(-))=alpha^(')p^(mu)tau\frac{1}{2} \alpha^{\prime} p^{\mu}\left(\sigma^{+}+\sigma^{-}\right)=\alpha^{\prime} p^{\mu} \tau, which is periodic.
p^(mu)p^{\mu} is the momentum:
T int d sigmaX^(˙)^(mu)(tau,sigma)=(1)/(2pialpha^('))(2pialpha^(')p^(mu))=p^(mu)T \int d \sigma \dot{X}^{\mu}(\tau, \sigma)=\frac{1}{2 \pi \alpha^{\prime}}\left(2 \pi \alpha^{\prime} p^{\mu}\right)=p^{\mu}
where we noted that oscillatory terms integrate to zero.
x_(0)^(mu)+alpha^(')p^(mu)taux_{0}^{\mu}+\alpha^{\prime} p^{\mu} \tau is the center of mass of the string:
(:X^(mu):)=(1)/(2pi)int d sigmaX^(mu)(tau,sigma)=x_(0)^(mu)+alpha^(')p^(mu)tau\left\langle X^{\mu}\right\rangle=\frac{1}{2 \pi} \int d \sigma X^{\mu}(\tau, \sigma)=x_{0}^{\mu}+\alpha^{\prime} p^{\mu} \tau
where we took n rarr-nn \rightarrow-n in the second line. This is equal to X^(mu)(sigma^(+))X^{\mu}\left(\sigma^{+}\right)iff (alpha_(-n)^(mu))^(**)=\left(\alpha_{-n}^{\mu}\right)^{*}=alpha_(n)^(mu)\alpha_{n}^{\mu}. Similar story for X_(R)X_{R}.
The solution must also satisfy the Virasoro constraints:
ds^(2)=-dtau^(2)+dsigma^(2)=-dsigma^(+)dsigma^(-)=-(1)/(2)(dsigma^(+)dsigma^(-)+dsigma^(-)dsigma^(+))d s^{2}=-d \tau^{2}+d \sigma^{2}=-d \sigma^{+} d \sigma^{-}=-\frac{1}{2}\left(d \sigma^{+} d \sigma^{-}+d \sigma^{-} d \sigma^{+}\right)
Hence, the vanishing of the energy momentum tensor becomes
{:[T_(++)=del_(+)X*del_(+)X=(1)/(4)((X^(˙))+X^(˙)^('))^(2)=0],[T_(-)=del_(-)X*del_(-)X=(1)/(4)((X^(˙))-X^('))^(2)=0]:}\begin{aligned}
& T_{++}=\partial_{+} X \cdot \partial_{+} X=\frac{1}{4}\left(\dot{X}+\dot{X}^{\prime}\right)^{2}=0 \\
& T_{-}=\partial_{-} X \cdot \partial_{-} X=\frac{1}{4}\left(\dot{X}-X^{\prime}\right)^{2}=0
\end{aligned}
while T_(+-)=T_(-+)=del_(+)X*del_(-)X-(1)/(2)h_(+-)(h^(+-)+h^(-+))del_(+)X*del_(-)X=0T_{+-}=T_{-+}=\partial_{+} X \cdot \partial_{-} X-\frac{1}{2} h_{+-}\left(h^{+-}+h^{-+}\right) \partial_{+} X \cdot \partial_{-} X=0 so T_(+-)=T_{+-}=T_(-+)T_{-+}trivially vanishes.
Let us compute Fourier expansion of T_(+-+-)T_{ \pm \pm}. First note that
where we replaced alpha_(-m)*alpha_(m)\alpha_{-m} \cdot \alpha_{m} with alpha_(m)*alpha_(-m)\alpha_{m} \cdot \alpha_{-m} in the second term of the second line. Note that this only holds classically. In quantum theory [alpha_(m)^(mu),alpha_(-m)^(mu)]!=0\left[\alpha_{m}^{\mu}, \alpha_{-m}^{\mu}\right] \neq 0 leading to an ordering ambiguity. Similarly we find
This will be based on BBS pg 40,41,58-62. After choosing conformal gauge h_(alpha beta)=n_(alpha beta)h_{\alpha \beta}=n_{\alpha \beta}, their is still a residual reparamaterisation symmetry. Let deltasigma^(alpha)=-xi^(alpha)\delta \sigma^{\alpha}=-\xi^{\alpha} be an infinitesimal reparametrisation and Lambda\Lambda be an infinitesimal parameter for a Weyl rescaling. The residual symmetries obey
This is known as the "conformal killing equation" and the solutions are known as conformal Killing vectors. In other words, we are free to make infinitesimal diffeomorphisms that correspond to Weyl transformations. These are known as "conformal transformations." Conformal transformations change distances, but preserve angles. Theories invariant under conformal transformations are known as "conformal field theories." The string worldsheet theory is a 2d CFT.
In D > 2D>2 spacetime dimensions, there are ((D+2)/(2))\binom{D+2}{2} solutions, which generate SO(D,2)S O(D, 2) corresponding to the "conformal group" in DD-dimensional Minkowski space. In DD-dimensional Euclidean space, the conformal group is SO(D+1,1)S O(D+1,1). In D=2D=2, the conformal killing equation has an infinite number of solutions, so the conformal group is infinite-dimensional.
Let's verify this. In lightone cords sigma^(+-)=sigma^(0)+-sigma^(1),xi^(+-)=xi^(0)+-xi^(1)\sigma^{ \pm}=\sigma^{0} \pm \sigma^{1}, \xi^{ \pm}=\xi^{0} \pm \xi^{1} the conformal Killing equations becomes
This is solved by xi_(+)(tau,sigma)=f_(+)(sigma^(-)),xi_(-)(tau,sigma)=f_(-)(sigma^(+))\xi_{+}(\tau, \sigma)=f_{+}\left(\sigma^{-}\right), \xi_{-}(\tau, \sigma)=f_{-}\left(\sigma^{+}\right), where f_(+-)f_{ \pm}are arbitrary functions respecting boundary conditions of worldsheet. Raising indices, xi^(+-)=\xi^{ \pm}=f^(+-)(sigma^(+-))f^{ \pm}\left(\sigma^{ \pm}\right). For closed strings a complete basis of solutions is given by
f_(n)^(+-)=e^(insigma^(+-)),n inZf_{n}^{ \pm}=e^{i n \sigma^{ \pm}}, n \in \mathbb{Z}
which are periodic in sigma rarr sigma+2pi\sigma \rightarrow \sigma+2 \pi. We may then define an infinite set of generators
which generate the diffeomorphisms deltasigma^(+-)=epsilon_(n)^(+-)V_(n)^(+-)(sigma^(+-))=epsilon_(n)^(+-)e^(insigma^(+-))\delta \sigma^{ \pm}=\epsilon_{n}^{ \pm} V_{n}^{ \pm}\left(\sigma^{ \pm}\right)=\epsilon_{n}^{ \pm} e^{i n \sigma^{ \pm}}, where epsilon^(+-)\epsilon^{ \pm} are infinitesimal parameters.
Let's Wick-rotate tau rarr-i tau\tau \rightarrow-i \tau so the worldsheet becomes Euclidean and then map it to complex plane:
Under this mapping, the cyclinder gets mapped to a plane. The infinite past on the cylinder corresponds to the origin of the plane. We also see that conformal transformations correspond to analytic functions on the complex plane:
which corresponds to two copies of the "classical Virasoro algebra." We will see later that it can get modified in quantum theory due to a "conformal anomaly." l_(0,+-1)l_{0, \pm 1} and bar(l)_(0,+-1)\bar{l}_{0, \pm 1} generate the finite dimensional subgroup SO(1,3)=S O(1,3)=SL(2,C)//Z_(2)S L(2, \mathbb{C}) / \mathbb{Z}_{2}, as you will show in the HW. This is the 2d2 d case of SO(D+1,1)S O(D+1,1)
which is the conformal group for D > 2D>2 Euclidean dimensions. In 2d2 d, this is konwn as the "restricted conformal group".
For open strings with Neumann boundary conditions, the residual symmetries in conformal gauge are generated by
V_(n)=e^(insigma^(+))(del)/(delsigma^(t))+e^(insigma^(-))(del)/(delsigma^(-))V_{n}=e^{i n \sigma^{+}} \frac{\partial}{\partial \sigma^{t}}+e^{i n \sigma^{-}} \frac{\partial}{\partial \sigma^{-}}
as you will show on HW. Hence, there is just one copy of Virasoro algebra. We can Wick-rotate to a Euclidean worldsheet and map to the upper half of the complex plane via z=e^(isigma^(-)), bar(z)=e^(isigma^(+))z=e^{i \sigma^{-}}, \bar{z}=e^{i \sigma^{+}}:
Note that the boundary of the string gets mapped to real axis. The Virasoro generators then become
where X^(+-)=(1)/(sqrt2)(X^(0)+-X^(D-1))X^{ \pm}=\frac{1}{\sqrt{2}}\left(X^{0} \pm X^{D-1}\right) and I in{1,dots,D-2}I \in\{1, \ldots, D-2\} labels transverse directions. Equivalently, we set x_(0)^(+)=alpha_(n)^(+)= tilde(alpha)_(n)^(+)=0,n!=0x_{0}^{+}=\alpha_{n}^{+}=\tilde{\alpha}_{n}^{+}=0, n \neq 0. This is known as "lightcone gauge." Note that we can't set terms linear in sigma^(+-)\sigma^{ \pm}to zero because they are not periodic in sigma\sigma.
In lightcone gauge, the mass formula and level-matching condition reduce to
where we noted that alpha_(-n)*alpha_(n)=-alpha_(-n)^(+)alpha_(n)^(-)-alpha_(-n)^(-)alpha_(n)^(+)+alpha_(-n)^(I)alpha_(n)^(I)=alpha_(-n)^(I)alpha_(n)^(I)\alpha_{-n} \cdot \alpha_{n}=-\alpha_{-n}^{+} \alpha_{n}^{-}-\alpha_{-n}^{-} \alpha_{n}^{+}+\alpha_{-n}^{I} \alpha_{n}^{I}=\alpha_{-n}^{I} \alpha_{n}^{I} in lightcone gauge. Hence in lightcone gauge, only transverse modes contribute to mass. Moreover, the modes of X^(-)X^{-}can be fixed in terms of transverse modes. This follows from the Virasoro constraints:
Recalling that alpha_(0)^(-)= tilde(alpha)_(0)^(-)=sqrt((alpha^('))/(2))p^(-)\alpha_{0}^{-}=\tilde{\alpha}_{0}^{-}=\sqrt{\frac{\alpha^{\prime}}{2}} p^{-}, for n=0n=0 these relations imply the mass
formula:
For an open string with Neumann boundary conditions, the story is similar except that there is only one set of oscillators and p^(mu)rarr2p^(mu)p^{\mu} \rightarrow 2 p^{\mu}. Here is a summary for open strings:
Mass formula: M^(2)=(1)/(alpha^('))sum_(n > 0)alpha_(-n)^(I)alpha_(n)^(I)M^{2}=\frac{1}{\alpha^{\prime}} \sum_{n>0} \alpha_{-n}^{I} \alpha_{n}^{I}
Hence, in lightcone gauge we express the theory in terms of transverse modes, which encode the independent physical degrees of freedom. To quantize the theory in light cone gauge, we must therefore quantize the transverse degrees of freedom.
9 Canonical Quantisation
Based on BBS pg 35-37, Tong pg 28-30. To quantize in a covariant way, we impose the canonical commutation relations:
Lot a_(m)^(mu)=(1)/(sqrtm)alpha_(m)^(mu),a_(m)^(mu†)=(1)/(sqrtm)alpha_(-m)^(mu),m > 0a_{m}^{\mu}=\frac{1}{\sqrt{m}} \alpha_{m}^{\mu}, a_{m}^{\mu \dagger}=\frac{1}{\sqrt{m}} \alpha_{-m}^{\mu}, m>0. Then
which is essentially the algebra of raising and lowering operators for harmonic oscillators. For open strings, there is only one set of oscillators.
Note that [a_(m)^(0),a_(m)^(0†)]=-1\left[a_{m}^{0}, a_{m}^{0 \dagger}\right]=-1. This gives negative norm states: Let |psi:)=|\psi\rangle=a_(m)^(0†)|0:),m > 0a_{m}^{0 \dagger}|0\rangle, m>0. Then
where the first term after the second equality vanishes and we used the commutation relations in the second term. Negative norm states can be removed by imposing additional constraints on the Hilbert space encoding which encode the Virasoro constraints (see chapter 24 of Zwiebach for more detials).
In lightcone gauge negative norm states do not appear because we eliminate oscillators of X^(+-)X^{ \pm}and the Hilbert space is then constructed using only transverse oscillators and is manifestly positive definite. We will therefore proceed with quantization in light cone gauge. Our main task will be to show that
where I in1,dots,D-2I \in 1, \ldots, D-2 and all other commutators vanish. In light cone gauge x_(0)^(-)x_{0}^{-}and p^(+)p^{+}are not eliminated, so we must also impose the commutation relation [x_(0)^(-),p^(+)]=-i\left[x_{0}^{-}, p^{+}\right]=-i.
Let's prove the commutation relations. Recall that for closed string
which vanishes after taking n rarr-nn \rightarrow-n in the first term of the second line.
This concludes the derivation of the canonic commutation relations for the closed string. For the open string there is only one set of oscillators, so we have
where I=1,dots,D-2I=1, \ldots, D-2 label transverse directions. For open strings, there are only one set of oscillators alpha_(m)^(I)\alpha_{m}^{I}. Let
We encountered this state when we quantised the relativistic point particle using the world line formalism, which only gives single-particle states. In string theory, we can act on this state with alpha_(-m)^(I)\alpha_{-m}^{I} and tilde(alpha)_(-m)^(I),m > 0\tilde{\alpha}_{-m}^{I}, m>0 to give an infinite tower of massive higher-spin states. Note that these are still single-particle states.
Since all the oscillators are transverse, the Hilbert space is manifestly positive definite. Hence lightcone gauge makes all the physical degrees of freedom manifest. Let's compute the spectrum, starting from the open string with Neumann boundary conditions.
Open string spectrum
Recall that in light cone gauge, the Virasoro constraints fix p^(-)p^{-}in terms of transverse oscillators implying the following mass formula:
where s inCs \in \mathbb{C}. This sum converges for Re(s) > 1\operatorname{Re}(s)>1, and has a pole at s=1s=1. It can be analytically continued around the pole to give zeta(-1)=-(1)/(12)\zeta(-1)=-\frac{1}{12} as you will show in the HW. Hence, we find that
a=(D-2)/(24)a=\frac{D-2}{24}
Let us consider the first excited state: quadalpha_(-1)^(I)∣0;k >\quad \alpha_{-1}^{I} \mid 0 ; k>. In this case we have N=1N=1 so M^(2)=(1)/(alpha^('))(1-a)M^{2}=\frac{1}{\alpha^{\prime}}(1-a). Note that the state is a (D-2)(D-2)-component vector and recall that Lorentz invariance implies that massive states must form represntations of SO(D-1)S O(D-1) and massless states must form representations of SO(D-2)S O(D-2). Since we cannot package the (D-2)(D-2) states into a representation of SO(D-1)S O(D-1), it must be a massless state. Indeed this is the structure of single
particle states that we found when quantizing Maxwell theory. Hence, we find that open strings contain massless spin- 1 particles, i.e. gauge fields, in their spectrum. It follows that
Hence, string theory predicts the dimension of spacetime! If D!=26D \neq 26, the theory would not be Lorentz-invariant.
A more rigorous (but much more tedious) way to derive a=1,D=26a=1, D=26 is to compute the commutator
[J^(I-),J^(K-)]\left[J^{I-}, J^{K-}\right]
where J^(mu nu)J^{\mu \nu} are Lorentz generators computed from Noether's theorem and expressed in terms of transverse oscillators after plugging in mode expansion and eliminating alpha_(n)^(-)\alpha_{n}^{-}oscillators in lightcone gauge. Note that generators of Lorentz group must satisfy the following algebra:
Hence [J^(I-),J^(K-)]=0\left[J^{I-}, J^{K-}\right]=0. After a tedious calculation one finds that this is only possible if a=1,D=26a=1, D=26. For more details, see section 2.4 of Tong and section 2.3.1 of GSW.
Let us return to the open string spectrum. The ground state |0;k:)|0 ; k\rangle is a scalar with momentum k^(mu)k^{\mu} and mass
The tachyon indicates an instability of the vacuum of the bosonic string. Recall that the mass squared of a field is the quadratic term of the potential:
so if M^(2) < 0M^{2}<0, we are expanding around a maximum of the potential. The tachyon can be removed by introducing supersymmetry, as you will see next term.
Finally, let us consider the first states with positive mass squared, correpsonding to N=2N=2 :
There are 24 states with a single index while the states with two indices correspond to a symmetric 24 xx2424 \times 24 matrix giving ((24)/(2))+24=300\binom{24}{2}+24=300 states. In total, there are 324 massive states at this level. Note that a symmetric traceless 25 xx2525 \times 25 matrix has ((25)/(2))+25-1=324\binom{25}{2}+25-1=324 independent components, so the states indeed form a representation of SO(D-1)=SO(25)S O(D-1)=S O(25), as expected.
Closed string spectrum
For the closed string in lightcone gauge, we construct states using two sets of transverse oscillators. Now the mass formula is
where NN and tilde(N)\tilde{N} are defined in terms of alpha_(-n)^(I)\alpha_{-n}^{I} and tilde(alpha)_(-n)^(I)\tilde{\alpha}_{-n}^{I}, repectively, and we the normal ordering constant to a=1a=1. We also have the level-matching constraint N= tilde(N)N=\tilde{N}.
Let's look at the first two mass levels:
N=0N=0 : Ground state |0;k:)|0 ; k\rangle has M^(2)=-(4)/(alpha^('))M^{2}=-\frac{4}{\alpha^{\prime}}, tachyon
N=1N=1 : A general state takes the following form:
This can be understood as the tensor product of two massless vectors, one left-moving and one night-moving. We can decompose this into a symmetric traceless part (graviton), an antisymmetric part (Kalb-Ramond), and the trace part (dilaton) as shown below:
We see that massless states indeed form reps of SO(D-2)S O(D-2). These massless fields will also play a rate in superstring.
Summary: String theory predicts gravity! Also unifies gravity with gauge theory since open strings give rise to gauge fields (like photons) while closed strings give rise to gravitons. String theory also predicts infinite tower of higher spin states with mass O(1//alpha^('))\mathcal{O}\left(1 / \alpha^{\prime}\right).
11 CFT: Virasoro from Stress Tensor
This is based on Zwiebach pg 254,258,259254,258,259, and BBS pg 63,64 . We will now develop ad CFT techniques for doing worldsheet calculations. This will be much more efficient than working with mode expansions and the oscillator algebra. In particular, we will use it to give a more elegant and rigorous proof of the critical dimension and to compute string amplitudes. We will not impose lightcone gauge.
First let us demonstrate that the modes of the stress tensor L_(m), tilde(L)_(m)L_{m}, \tilde{L}_{m} are Virasoro generators. To do that, we need the following identities (for the closed string):
Let us briefly comment on the factor of -i-i on appearing in these relations. Consider a scalar theory in DD spacetime dimensions. For symmetry delta phi\delta \phi with Noeither charge QQ, in the quantum theory, we have [Q,phi]=-i delta phi[Q, \phi]=-i \delta \phi.
Now let's map the worldsheet to the complex plane: tau rarr-i tau\tau \rightarrow-i \tau so that z=z=e^(isigma^(-))=e^(tau-i sigma)e^{i \sigma^{-}}=e^{\tau-i \sigma} and bar(z)=e^(isigma^(+))=e^(tau+i sigma)\bar{z}=e^{i \sigma^{+}}=e^{\tau+i \sigma}. In terms of z, bar(z)z, \bar{z} we then have
(1)/(2pi i)ointdzz^(n+1)T(z)=(1)/(2pi i)sum_(m)ointdzz^(n-m-1)L_(m)=(1)/(2pi i)sum_(m)2pi idelta_(m,n)L_(m)=L_(n)\frac{1}{2 \pi i} \oint d z z^{n+1} T(z)=\frac{1}{2 \pi i} \sum_{m} \oint d z z^{n-m-1} L_{m}=\frac{1}{2 \pi i} \sum_{m} 2 \pi i \delta_{m, n} L_{m}=L_{n}
where the contour encloses z=0z=0 and we used Cauchy's theorem: oint(dz)/(2pi i)z^(k)=\oint \frac{d z}{2 \pi i} z^{k}=delta_(k,-1)\delta_{k,-1}. Similarly, we find tilde(L)_(n)=(1)/(2pi i)ointd bar(z) bar(z)^(n+1) tilde(T)( bar(z))\tilde{L}_{n}=\frac{1}{2 \pi i} \oint d \bar{z} \bar{z}^{n+1} \tilde{T}(\bar{z}). More generally Cauchy's theorem can be stated as follows:
Cauchy's Theorem
(1)/(2pi i)oint_(Gamma)dzf(z)=sum_(i)Res[f(z_(i))]\frac{1}{2 \pi i} \oint_{\Gamma} d z f(z)=\sum_{i} \operatorname{Res}\left[f\left(z_{i}\right)\right]
where the sum runs over the simple poles z_(i)z_{i} of ff inside the contour Gamma\Gamma, and near each pole
Q_(epsilon)=(1)/(2pi i)ointdzT(z)epsilon(z),quadQ_( tilde(epsilon))=(1)/(2pi i)ointd bar(z) tilde(T)( bar(z)) tilde(epsilon)( bar(z))Q_{\epsilon}=\frac{1}{2 \pi i} \oint d z T(z) \epsilon(z), \quad Q_{\tilde{\epsilon}}=\frac{1}{2 \pi i} \oint d \bar{z} \tilde{T}(\bar{z}) \tilde{\epsilon}(\bar{z})
are the generators of conformal transformations. More generally, for delta z=epsilon(z)\delta z=\epsilon(z) and delta bar(z)= tilde(epsilon)( bar(z))\delta \bar{z}=\tilde{\epsilon}(\bar{z}) an operator Phi(w, bar(w))\Phi(w, \bar{w}) transforms as
This is based on BBS pg 64-66 and Tong sections 4.3.1,4.3.3. Under a conformal transformation delta z=epsilon(z)\delta z=\epsilon(z), an operator Phi\Phi transforms as
delta_(epsilon)Phi(w, bar(w))=(1)/(2pi i)ointdz epsilon(z)[T(z),Phi(w, bar(w))]\delta_{\epsilon} \Phi(w, \bar{w})=\frac{1}{2 \pi i} \oint d z \epsilon(z)[T(z), \Phi(w, \bar{w})]
The contour encluses the origin and is specified by "radial ordering": operators to the left should be located further from from the origin (recall that langer radius in the complex plane corresponds to larger tau\tau on the cylinder so this is essentially time ordering). Hence, the commutator gives a contour which encircles the point ww :
In general, we can express the product of local operators as a series of local operators located at one of their positions. This is known as an "Operator
Product Expansion" (OPE). Hence, the contour integral for delta_(epsilon)Phi\delta_{\epsilon} \Phi is just the residue of the pole in the OPE of TT and Phi\Phi :
under (z, bar(z))rarr(z^(')(z), bar(z)^(')(( bar(z))))(z, \bar{z}) \rightarrow\left(z^{\prime}(z), \bar{z}^{\prime}(\bar{z})\right). To see this, let z^(')=z+epsilon(z), bar(z)^(')= bar(z)+ tilde(epsilon)( bar(z))z^{\prime}=z+\epsilon(z), \bar{z}^{\prime}=\bar{z}+\tilde{\epsilon}(\bar{z}), where epsilon, tilde(epsilon)\epsilon, \tilde{\epsilon} infinitesimal:
where... are higher order terms in epsilon\epsilon and tilde(epsilon)\tilde{\epsilon}. Hence, delta Phi=delta_(epsilon)Phi+delta_(epsilon)Phi\delta \Phi=\delta_{\epsilon} \Phi+\delta_{\epsilon} \Phi found via the OPE is indeed an infinitesimal conformal transformation.
All of the OPE's we need in practice can be derived from the following OPE:
where we write X(z, bar(z))=X(z)+X( bar(z))X(z, \bar{z})=X(z)+X(\bar{z}) and dots\ldots are non-singular as z rarr wz \rightarrow w. The singular part of the OPE can be read of from the correlation function (:X^(mu)(( bar(z)),( bar(z)))X^(nu)(z^('), bar(z)^(')):)\left\langle X^{\mu}(\bar{z}, \bar{z}) X^{\nu}\left(z^{\prime}, \bar{z}^{\prime}\right)\right\rangle, which can in turn be deduced from a path integral:
0=intDX(delta)/(deltaX_(mu)(z,( bar(z))))[e^(-S)X^(nu)(z^('), bar(z)^('))]0=\int \mathcal{D} X \frac{\delta}{\delta X_{\mu}(z, \bar{z})}\left[e^{-S} X^{\nu}\left(z^{\prime}, \bar{z}^{\prime}\right)\right]
where the integral is over all field configurations and the integral vanishes because the integrand contains a total functional derviative. The action is given by
{:[S=(T)/(2)intd^(2)sigma(X^(˙)^(2)+X^('2))","quad tau rarr-i tau],[=(1)/(2pialpha^('))intd^(2)z quad del X* bar(del)X","quadd^(2)z=dzd bar(z)],[=-(1)/(2pialpha^('))intd^(2)zX*del bar(del)X]:}\begin{aligned}
S & =\frac{T}{2} \int d^{2} \sigma\left(\dot{X}^{2}+X^{\prime 2}\right), \quad \tau \rightarrow-i \tau \\
& =\frac{1}{2 \pi \alpha^{\prime}} \int d^{2} z \quad \partial X \cdot \bar{\partial} X, \quad d^{2} z=d z d \bar{z} \\
& =-\frac{1}{2 \pi \alpha^{\prime}} \int d^{2} z X \cdot \partial \bar{\partial} X
\end{aligned}
Let us denote (:X^(mu)(z)X^(nu)(w):)=X^(mu)(z)X^(nu)^(⏜)(w)\left\langle X^{\mu}(z) X^{\nu}(w)\right\rangle=\overparen{X^{\mu}(z) X^{\nu}}(w), which is known as a "Wick contraction." We can then compute OPE of more general operators by summing over all possible contractions. We also define "normal ordering":
:O(z_(1))dotsO(z_(n)):=O(z_(1))dotsO(z_(n))-" all contractions ": \mathcal{O}\left(z_{1}\right) \ldots \mathcal{O}\left(z_{n}\right):=\mathcal{O}\left(z_{1}\right) \ldots \mathcal{O}\left(z_{n}\right)-\text { all contractions }
Hence, these operator have no singular terms in their OPE. We an then define composite operators like TT and tilde(T)\tilde{T} more precisely as
Note that tilde(T)( bar(z))delX^(mu)(w)=dots\tilde{T}(\bar{z}) \partial X^{\mu}(w)=\ldots since (:X^(mu)(z)X^(nu)(w):)=0\left\langle X^{\mu}(z) X^{\nu}(w)\right\rangle=0.
: e^(ik*X(omega, bar(w)))e^{i k \cdot X(\omega, \bar{w})} is a primary operator with h= tilde(h)=alpha^(')k^(2)//4h=\tilde{h}=\alpha^{\prime} k^{2} / 4.
First compute the following OPE (suppressing Lorentz indices for simplicity)
{:[T(z):e^(ikX(w)):=-(1)/(alpha^(')):del X(z)del X(z)::e^(ikX(w)):],[=-(1)/(alpha^('))[(-(ialpha^(')k)/(2(z-w)))^(2):e^(ikX(w)):+2((-ialpha^(')k)/(2(z-w))):del X(z)e^(ikX(w)):]]:}\begin{aligned}
T(z): e^{i k X(w)}: & =-\frac{1}{\alpha^{\prime}}: \partial X(z) \partial X(z):: e^{i k X(w)}: \\
& =-\frac{1}{\alpha^{\prime}}\left[\left(-\frac{i \alpha^{\prime} k}{2(z-w)}\right)^{2}: e^{i k X(w)}:+2\left(\frac{-i \alpha^{\prime} k}{2(z-w)}\right): \partial X(z) e^{i k X(w)}:\right]
\end{aligned}
where the first term came from contracting both delX^(')\partial X^{\prime} 's in the stress tensor with the exponential and the second term came from contracting only one del X\partial X. This gives a factor of two since there are two possible ways to perform this contraction. We then obtain
{:[T(z):e^(ikX(w)):=(alpha^(')k^(2)//4)/((z-w)^(2)):e^(ikX(w)):+(ik:del X(z)e^(ikX(w)):)/(z-w)+dots],[=(alpha^(')k^(2)//4)/((z-w)^(2)):e^(ikX(w)):+(ik:(del X(w)+dots)e^(ikx(w)):)/(z-w)+dots],[=(alpha^(')k^(2)//4)/((z-w)^(2)):e^(ikX(w)):+(del_(w):e^(ikX(w)):)/(z-w)+dots]:}\begin{aligned}
T(z): e^{i k X(w)}: & =\frac{\alpha^{\prime} k^{2} / 4}{(z-w)^{2}}: e^{i k X(w)}:+\frac{i k: \partial X(z) e^{i k X(w)}:}{z-w}+\ldots \\
& =\frac{\alpha^{\prime} k^{2} / 4}{(z-w)^{2}}: e^{i k X(w)}:+\frac{i k:(\partial X(w)+\ldots) e^{i k x(w)}:}{z-w}+\ldots \\
& =\frac{\alpha^{\prime} k^{2} / 4}{(z-w)^{2}}: e^{i k X(w)}:+\frac{\partial_{w}: e^{i k X(w)}:}{z-w}+\ldots
\end{aligned}
Similarly
tilde(T)( bar(z)):e^(ikX( bar(w))):=(alpha^(')k^(2)//4)/((( bar(z))-( bar(w)))^(2)):e^(ikX( bar(w))):+(del_( bar(w)):e^(ikX( bar(w))):)/(( bar(z))-( bar(w)))+dots\tilde{T}(\bar{z}): e^{i k X(\bar{w})}:=\frac{\alpha^{\prime} k^{2} / 4}{(\bar{z}-\bar{w})^{2}}: e^{i k X(\bar{w})}:+\frac{\partial_{\bar{w}}: e^{i k X(\bar{w})}:}{\bar{z}-\bar{w}}+\ldots
13 Critical Dimension from CFT
This is based on Tong pg 82, sections 5.1, 5.2, BBS pg 73-76, GSW pg 121-127. We will revisit the calculation of critical dimension using 2d CFT. In lightcone
gauge, we found that Lorentz invariance was amalous unless D=26D=26. Now we will work covariantly and show that conformal symmetry is anomalous unless D=26D=26. The conformal anomaly shows up in central extension of the Virasoro algebra which can be derived from the OPE of the stress tensor with itself:
To obtain the second line, we wrote del X(z)=del X(w)+(z-w)del^(2)X(w)+dots\partial X(z)=\partial X(w)+(z-w) \partial^{2} X(w)+\ldots in the first line. Hence, we see that TT is not a primary operator because of the 1//(z-w)^(4)1 /(z-w)^{4} term.
where cc is the "conformal anomaly" or "central change". For the bosonic string c=Dc=D, ie. each X^(mu)(z)X^{\mu}(z) contributes 1 to the central charge. More generally, c counts the number of riigt-moving degrees of freddom. TT is only a primary it c=0c=0, in which case it has weight (2,0)(2,0). Similarly, if the left-moving central charge tilde(c)=0\tilde{c}=0, then tilde(T)( bar(z))\tilde{T}(\bar{z}) is a primary with weight (0,2)(0,2).
Let's verify that cc corresponds to central extension of Virasoro algebra. First recall that
Let us do the zz integral first, holding ww fixed. Using radial ordering, the commutator is computed by considering the zz integral along a small path encircling ww (which we will refer to as Gamma_(w)\Gamma_{w} ) so we can compute it from the OPE of T(z)T(w)T(z) T(w) :
Hence, we the Virasaro algebra recieves quantum correction proportional to the central charge cc.
Ghost fields
In order to have zero central charge, we need to add worldsheet fields with negative central charge. These violate spin statistics theorem and are known as "ghosts." In fact ghosts arise when gauge-fixing path integrals using the "Fadeev-Popov procedure." Let's sketch how this works for bosonic string. For details see GSW pg 121-127 and Tong sections 5.1 and 5.2. Recall that in light cone gauge the worldsheet metric is
Due to this gauge symmetry, path integral contains and infinite redundancy coming over the sum over physically equivalent field configurations related by a
gauge transformations. We can remove this redundancy by inserting
1=int dx delta(f(x))|f^(')(x)|," since "int dx delta(f(x))=(1)/(|f^(')(x_(i))|)," where "f(x_(i))=01=\int d x \delta(f(x))\left|f^{\prime}(x)\right|, \text { since } \int d x \delta(f(x))=\frac{1}{\left|f^{\prime}\left(x_{i}\right)\right|}, \text { where } f\left(x_{i}\right)=0
The other key fact we need is that determinants can be represented by path integrals over anticommuting scalar fields (ghosts), as you will show in the HW:
{:[det(deltah_(++)^(')//deltaxi_(+))=intDcDb exp(-(1)/(pi)intd^(2)sigma bdel_(+)c)],[det(deltah_(--)^(')//deltaxi_(-))=int D tilde(c)D tilde(b)exp(-(1)/(pi)intd^(2)sigma( tilde(b))del_(-)( tilde(c)))]:}\begin{aligned}
& \operatorname{det}\left(\delta h_{++}^{\prime} / \delta \xi_{+}\right)=\int \mathcal{D} c \mathcal{D} b \exp \left(-\frac{1}{\pi} \int d^{2} \sigma b \partial_{+} c\right) \\
& \operatorname{det}\left(\delta h_{--}^{\prime} / \delta \xi_{-}\right)=\int D \tilde{c} D \tilde{b} \exp \left(-\frac{1}{\pi} \int d^{2} \sigma \tilde{b} \partial_{-} \tilde{c}\right)
\end{aligned}
After converting to z, bar(z)z, \bar{z} coordinates the get
The equations of motion imply that the fields (b,c)(b, c) are holomorphic while ( tilde(b), tilde(c))(\tilde{b}, \tilde{c}) are antiholomprhic. Moreover, one can show that they have the following stress tensor:
T=-2:b del c:+:c del b:, tilde(T)=-2: tilde(b) bar(del) tilde(c):+: tilde(c) bar(del) tilde(b):T=-2: b \partial c:+: c \partial b:, \tilde{T}=-2: \tilde{b} \bar{\partial} \tilde{c}:+: \tilde{c} \bar{\partial} \tilde{b}:
under (z,w)rarr(w(z), bar(w)( bar(z)))(z, w) \rightarrow(w(z), \bar{w}(\bar{z})). Note that
dzd bar(z)rarr(del z)/(del w)(del( bar(z)))/(del( bar(w)))dwd bar(w)d z d \bar{z} \rightarrow \frac{\partial z}{\partial w} \frac{\partial \bar{z}}{\partial \bar{w}} d w d \bar{w}
so dzd bar(z)d z d \bar{z} has weight (-1,-1)(-1,-1) under a conformal rescaling. Hence
V=intd^(2)z hat(V)V=\int d^{2} z \hat{V}
will be conformal invariant if hat(V)\hat{V} is a primary of weight (+1,+1)(+1,+1). In that case, VV is "vertex operator." Vertex ops correspond to the physical states.
First note that the operator has no Lorentz indices, so it describes a scalar particle. In chapter 12 we showed that =e^(ip-X)=e^{i p-X} : is a primary with weight h= tilde(h)=alpha^(')p^(2)//4h=\tilde{h}=\alpha^{\prime} p^{2} / 4. If h= tilde(h)=+1h=\tilde{h}=+1, then M^(2)=-p^(2)=-(4)/(alpha^('))M^{2}=-p^{2}=-\frac{4}{\alpha^{\prime}}, which is indeed the tachyon mass found in chapter 10.
Hence, the operator has conformal weight h=1+alpha^(')p^(2)//4h=1+\alpha^{\prime} p^{2} / 4. Similarly, tilde(h)=\tilde{h}=1+alpha^(')p^(2)//41+\alpha^{\prime} p^{2} / 4. Hence h= tilde(h)=1rarrp^(2)=0h=\tilde{h}=1 \rightarrow p^{2}=0.
We still need to check that hat(V)\hat{V} is primary. In fact, there is a contribution to OPE which we haven't previously computed which can spoil this property. It comes from contracting one del X\partial X in AA with BB and the other del X\partial X in AA with CC :
where the factor of 2 in the first line comes from performing two equivalent contractions. Similarly, when we compute OPE tilde(T)\tilde{T} we find a cubic pole
as expected for massless spinning fields. Finally, we can decompose epsilon_(mu nu)\epsilon_{\mu \nu} into symmetric traceless, antisymmetric, and a trace piece, corresponding to tje graviton, Kalb-Ramond, and dilation fields respectively. Hence, we recover the massless states of the closed bosonic string.
This will be based on Tong sections 6.1, 6.2, and GSW section 1.1. Scattering amplitudes describe the probability for states to interact in a certain way. In QFT, they can be computed perturbatively by summing over Feynman diagrams. In string theory, one sums over all worldsheet topologies. For closed strings (which we will focus on) the number of loops corresponds to the number of handles or genus of the worlsheet. Note that many different Feynman diagrams can arise from a single string diagram at low energies because they correspond to different degenerations of the worldsheet. To compute string amplitudes in practice, we use conformal symmetry of the worldsheet to map infinite tubes corresponding to external legs to punctures:
At each puncture, we then place a vertex operator corresponding to an external state.
Let's consider the tree-level nn-point scattering of tachyons. This is given by the correlation function of nn tachyon vertex operators on sphere:
where p_(i)^(mu)p_{i}^{\mu} is the momentum of particle ii and we can use the global conformal group SL(2,C)S L(2, \mathbb{C}) to fix the location of three punctures. More explicitly, the correlator is given by
where J^(mu)(z, bar(z))=sum_(i=1)^(n)p_(i)^(mu)delta^(2)(z-z_(i))J^{\mu}(z, \bar{z})=\sum_{i=1}^{n} p_{i}^{\mu} \delta^{2}\left(z-z_{i}\right) and hat(D)=(1)/(2pialpha^('))del bar(del)\hat{D}=\frac{1}{2 \pi \alpha^{\prime}} \partial \bar{\partial}. This is a Gaussian integral. You worked out a finite-dimensional version of this in HW:
int_(-oo)^(oo)dxe^(ax^(2)+iJx)prope^(J^(2)//4a)\int_{-\infty}^{\infty} d x e^{a x^{2}+i J x} \propto e^{J^{2} / 4 a}
Similar manipulations of the functional integral give (see chapter 8 of Srednicki)
where c=1-a-bc=1-a-b (see see Tong section 6.5 for a derivation). Hence,
A_(4)prop(Gamma(-1-alpha^(')s//4)Gamma(-1-alpha^(')t//4)Gamma(-1-alpha^(')u//4))/(Gamma(2+alpha^(')s//4)Gamma(2+alpha^(')t//4)Gamma(2+alpha^(')u//4))\mathcal{A}_{4} \propto \frac{\Gamma\left(-1-\alpha^{\prime} s / 4\right) \Gamma\left(-1-\alpha^{\prime} t / 4\right) \Gamma\left(-1-\alpha^{\prime} u / 4\right)}{\Gamma\left(2+\alpha^{\prime} s / 4\right) \Gamma\left(2+\alpha^{\prime} t / 4\right) \Gamma\left(2+\alpha^{\prime} u / 4\right)}
This is known as the "Virasoro-Shapiro" amplitude. It is expressed in terms of the Mandelstam variables
Similarly, (alpha^('))/(2)p_(1)*p_(4)=-2-alpha^(')t//4\frac{\alpha^{\prime}}{2} p_{1} \cdot p_{4}=-2-\alpha^{\prime} t / 4 and (alpha^('))/(2)p_(1)*p_(3)=-2-alpha^(')u//4\frac{\alpha^{\prime}}{2} p_{1} \cdot p_{3}=-2-\alpha^{\prime} u / 4 Another useful identity is
and similarly, p_(3)*p_(4)=p_(1)*p_(2)p_{3} \cdot p_{4}=p_{1} \cdot p_{2}. Applying (6) to (5) and using the above identities, we see that the arguments of the Gamma functions are
{:[a=(alpha^('))/(2)p_(2)*p_(3)+1=(alpha^('))/(2)p_(1)*p_(4)+1=(-2-alpha^(')t//4)+1=-1-alpha^(')t//4],[b=(alpha^('))/(2)p_(3)*p_(4)+1=-1-alpha^(')s//4],[c=1-a-b=1+(1+alpha^(')t//4)+(1+alpha^(')s//4)],[=3+(alpha^('))/(4)(s+t)=-1-alpha^(')u//4]:}\begin{aligned}
& a=\frac{\alpha^{\prime}}{2} p_{2} \cdot p_{3}+1=\frac{\alpha^{\prime}}{2} p_{1} \cdot p_{4}+1=\left(-2-\alpha^{\prime} t / 4\right)+1=-1-\alpha^{\prime} t / 4 \\
& b=\frac{\alpha^{\prime}}{2} p_{3} \cdot p_{4}+1=-1-\alpha^{\prime} s / 4 \\
& c=1-a-b=1+\left(1+\alpha^{\prime} t / 4\right)+\left(1+\alpha^{\prime} s / 4\right) \\
& =3+\frac{\alpha^{\prime}}{4}(s+t)=-1-\alpha^{\prime} u / 4
\end{aligned}
Hence we obtain VS amplitude which describes the scattering of four tachyons in bosonic string theory. Let's analyse some of it's properties.
Poles and Residues
Recall that Gamma(z)\Gamma(z) has poles at z=0,-1,-2,dotsz=0,-1,-2, \ldots If we fix tt and vary ss, we see that Gamma(-1-alpha^(')s//4)\Gamma\left(-1-\alpha^{\prime} s / 4\right) has poles when
Hence, poles correspond to masses af exchanged particles. We can read off the spin of the exchanged particles from the residue of the pole. Near s=(4)/(alpha)(n-1)s=\frac{4}{\alpha}(n-1),
{:[A_(4) prop(1)/(s-M_(n)^(2))(Gamma(-1-alpha^(')t//4)Gamma(-1-alpha^(')u//4))/(Gamma(2+alpha^(')t//4)Gamma(2+alpha^(')u//4))],[=(1)/(s-M_(n)^(2))(Gamma(-1-alpha^(')t//4)Gamma(n+2+alpha^(')t//4))/(Gamma(-n-1-alpha^(')t//4)Gamma(2+alpha^(')t//4))]:}\begin{aligned}
\mathcal{A}_{4} & \propto \frac{1}{s-M_{n}^{2}} \frac{\Gamma\left(-1-\alpha^{\prime} t / 4\right) \Gamma\left(-1-\alpha^{\prime} u / 4\right)}{\Gamma\left(2+\alpha^{\prime} t / 4\right) \Gamma\left(2+\alpha^{\prime} u / 4\right)} \\
& =\frac{1}{s-M_{n}^{2}} \frac{\Gamma\left(-1-\alpha^{\prime} t / 4\right) \Gamma\left(n+2+\alpha^{\prime} t / 4\right)}{\Gamma\left(-n-1-\alpha^{\prime} t / 4\right) \Gamma\left(2+\alpha^{\prime} t / 4\right)}
\end{aligned}
where M_(n)^(2)=(4)/(alpha^('))(n-1)M_{n}^{2}=\frac{4}{\alpha^{\prime}}(n-1). Noting that
{:[(Gamma(n+2+alpha^(')t//4))/(Gamma(2+alpha^(')t//4))=(n+2+alpha^(')t//4)(n+1+alpha^(')t//4)dots(3+alpha^(')t//4)],[=(alpha^(')t//4)^(n)+dots],[(Gamma(-1-alpha^(')t//4))/(Gamma(-n-1-alpha^(')t//4))=(-1-alpha^(')t//4)(-2-alpha^(')t//4)dots(-n-alpha^(')t//4)],[=(-alpha^(')t//4)^(n)+dots]:}\begin{aligned}
\frac{\Gamma\left(n+2+\alpha^{\prime} t / 4\right)}{\Gamma\left(2+\alpha^{\prime} t / 4\right)} & =\left(n+2+\alpha^{\prime} t / 4\right)\left(n+1+\alpha^{\prime} t / 4\right) \ldots\left(3+\alpha^{\prime} t / 4\right) \\
& =\left(\alpha^{\prime} t / 4\right)^{n}+\ldots \\
\frac{\Gamma\left(-1-\alpha^{\prime} t / 4\right)}{\Gamma\left(-n-1-\alpha^{\prime} t / 4\right)} & =\left(-1-\alpha^{\prime} t / 4\right)\left(-2-\alpha^{\prime} t / 4\right) \ldots\left(-n-\alpha^{\prime} t / 4\right) \\
& =\left(-\alpha^{\prime} t / 4\right)^{n}+\ldots
\end{aligned}
where ... indicates lower order terms in tt, we see that near s=(4)/(alpha)(n-1)s=\frac{4}{\alpha}(n-1) the amplitude is
This indicates that the highest spin of a particle with mass M_(n)M_{n} is J=2nJ=2 n. To see this, note that coupling of a scalar field phi\phi to a spin- JJ field chi^(mu_(1)dotsmu_(J))\chi^{\mu_{1} \ldots \mu_{J}} is of the form
From this, we see that an s-channel Feynman diagram with four external scalars exchanging spin- JJ particle scales like t^(J)t^{J}. It follows from (7) that the maximum spin being exchanged is 2n2 n, which is precisely what we expect from the closed string spectrum. Indeed, recall that the spectrum is given by
Let's take alpha^(')s,alpha^(')t rarr oo\alpha^{\prime} s, \alpha^{\prime} t \rightarrow \infty holding s//ts / t fixed. In this limit, each term in (8) separately diverges for n > 0n>0. Hence exchange of spin-2 particle (gravity) diverges
like E^(2)E^{2} as large energy EE and the exchange of higher-spin particles gives rise to even worse divergences. Hence, if we truncate the sum in (8) the amplitude is ill-behaved at high energies. Remarkably, including all states in the sum yields a result that is well-behaved at high energies.
In the high energy limit, we can neglect the masses of the external particles and take the external momenta to be
Plugging this into the VS amplitude then gives a function of ss and theta\theta. Taking s rarr oos \rightarrow \infty holding theta\theta fixed and using Stirling's approximation Gamma(x)∼sqrt(2pi x)(x//e)^(x)\Gamma(x) \sim \sqrt{2 \pi x}(x / e)^{x}, I find that lim_(s rarr oo)ln A_(4)=-alpha^(')sF(theta)\lim _{s \rightarrow \infty} \ln \mathcal{A}_{4}=-\alpha^{\prime} s F(\theta) where