nucl-th0309068/talk.tex
1: \documentclass{aipproc}
2: 
3: %\usepackage{epsf}
4: \usepackage{graphicx}
5: 
6: \layoutstyle{6x9}
7: 
8: \bibliographystyle{aipproc}
9: 
10: \begin{document}
11: 
12: \title{Simulating the scalar glueball on the lattice}
13: \author{Colin Morningstar}{address={
14: Department of Physics, Carnegie Mellon University,
15: Pittsburgh, PA 15213-3890}}
16: \author{Mike Peardon}{address={School of Mathematics, Trinity College, Dublin
17: 2, Ireland.}}
18: 
19: \begin{abstract}
20: Techniques for efficient computation of the scalar glueball mass on the lattice
21: are described. Directions and physics goals of proposed future calculations
22: will be outlined. 
23: \end{abstract}
24: \maketitle 
25: 
26: \section{Introduction}
27: 
28: Along with the mesons of the quark model, QCD allows for the existence of
29: bound-states of gluons, the glueballs. 
30: The scalar glueball is the lightest state in the spectrum of strongly
31: interacting gluons (described by the $SU(3)$ Yang-Mills theory) and in this
32: consistent, confining quantum field theory, it is stable. The world of gluons
33: alone 
34: thus provides an attractive starting point for investigations of glueballs.
35: Confinement is not described at all within QCD perturbation theory and so
36: computing properties of the glueballs requires a non-perturbative technique for
37: studying strong interactions.  
38: As a result, no
39: unambiguous identification of the scalar glueball has been made, although 
40: there is broad agreement that an extra state in the family of scalar 
41: resonances, unaccounted for by the quark model, exists between 1.5 and 
42: 1.8 GeV. 
43: In QCD, the glueball states will mix with the isoscalar resonances of the 
44: quark model, and will also decay strongly so 
45: to claim a complete understanding of the scalar mesons below 2 GeV 
46: therefore requires reliable knowledge of the masses and other properties of 
47: glueballs. 
48: 
49: In order to disentangle the complex picture presented by experiments much
50: more detailed and precise theoretical prediction from QCD must be made. At 
51: present, the only practical {\it ab initio} method for computing the 
52: non-perturbative properties of strongly interacting field theories such as QCD
53: is through Monte Carlo simulation of the lattice regularisation of the theory.
54: The glueballs of $SU(3)$ Yang-Mills theory have been studied
55: extensively in lattice simulations \cite{Michael:1989jr, Bali:1993fb, 
56: Morningstar:1999rf}, and this spectrum is becoming increasingly
57: accurately determined. This is the first step in understanding the
58: appearance of glueballs among the light scalar resonances of QCD. In spite of 
59: the limitation of these calculations, their output is proving to be very 
60: valuable in the construction and validation of phenomenological models.
61: 
62: In this article, we describe the state-of-the-art calculation of the
63: glueball spectrum of the Yang-Mills theory. Investigations of further 
64: properties of these fascinating states, including their behaviour in QCD (with 
65: the dynamics of quark fields included) are then outlined.
66: 
67: \section{Lattice QCD}
68: 
69: Lattice QCD solves two key problems arising from strongly interacting gauge
70: theories at a stroke; it provides both a non-perturbative regularisation of QCD
71: and a means by which observables in the theory can be predicted by computer.
72: This inherently gauge invariant regularisation is manifest through 
73: a direct cut-off of momenta at the Brillouin zone boundaries of the lattice. 
74: 
75: To begin, the path-integral formulation of the theory in Euclidean space is
76: taken. The four dimensions of space and time are discretised with a regular
77: lattice of points separated by a spacing, $a$. In the gauge-invariant 
78: formulation defined by Wilson nearly 30 years ago \cite{Wilson:1974sk}, the quark fields of QCD are defined
79: at sites of the lattice, while the gluonic degrees of freedom are represented
80: by parallel transporters (usually in the fundamental representation of the
81: gauge group) connecting nearest-neighbouring sites. A definition of the
82: Yang-Mills action is constructed from the trace of path-ordered products of the
83: link variables around small loops on the lattice. The smallest non-trivial loop
84: on the lattice circumnavigates a square of side-length $a$, usually called a
85: plaquette.  Similarly, a number of different definitions of the quark bilinear 
86: action, coupled to the gluons, can be made. 
87: 
88: Path integral expectation values can then be estimated by numerical methods.
89: If the theory is considered for space and time inside a finite-sized box, the
90: path integral at non-zero lattice spacing is reduced to a very-high-dimensional
91: ordinary integral. This type of problem can normally only be addressed by
92: Monte Carlo methods, where a stochastic sampling of points inside the phase
93: space to be integrated over is made.
94: 
95: \subsection{Stochastic estimation methods for quantum field theory}
96: 
97: After Wick rotation, the path integral for field theories such as QCD (at least 
98: at zero chemical potential) can be regarded as a statistical mechanical
99: partition function. The Boltzmann weight for a particular field configuration
100: is real and positive definite, and so can be interpreted as a statistical 
101: probability of the configuration appearing. In these circumstances, the 
102: natural Monte Carlo method to employ is importance sampling. An update 
103: algorithm is defined which generates configurations of the gluon fields, $U$ 
104: with probability density given by the Boltzmann weight, $\exp\{ -S(U)\}$. The 
105: method is employed on the computer to generate an ensemble of gluon field 
106: configurations. Expectation values are then estimated from ensemble averages. 
107: 
108: To compute the physical properties of the theory appropriate observation
109: functions on the underlying fields are defined with the required quantum 
110: numbers, and these observables are measured on all members of the stochastic 
111: ensemble.  Careful analysis allows a reliable determination of statistical 
112: uncertainties.
113: 
114: \subsection{Controlling simulation artefacts}
115: 
116: In order to make contact with the continuum field theory, a number of artefacts 
117: of the method must be controlled. The most obvious of these is the need to use
118: a finite-sized box for numerical work. In practical simulations, boxes with
119: side lengths of the order of 2 fm are feasible. 
120: 
121: In many cases, the qualitative behaviour of states in increasingly 
122: large-but-finite volumes can be predicted. Simulations are run for a range of 
123: box sizes, and the data matched to this predicted form to allow safe 
124: extrapolation to the infinite volume limit. In most studies of glueballs these 
125: effects are now extremely small and do not constitute the dominant systematic 
126: error. 
127: 
128: More difficult is the need to extrapolate data to the limit of zero lattice
129: spacing (usually called the continuum limit). Again, the most natural procedure
130: is to compute the physics of interest on a set of lattices with different grid
131: spacings, and extrapolate. In many cases, the expected leading-order behaviour
132: of these finite lattice spacing effects can be predicted, and this can be used 
133: to control data fitting. The problem with this direct approach is that the cost 
134: of computer simulations rises rapidly as the lattice spacing is diminished, at
135: least in proportion to $a^{-4}$ and generally worse than this. Since the
136: continuum limit exists at a critical point corresponding to a second order
137: phase transition, where the correlation length of the system diverges in units
138: of the lattice spacing, critical slowing down makes this cost higher. It is
139: clear that the cheapest computer simulations are those run on the coarsest
140: grids, which are furthest from the continuum. 
141: 
142: \subsection{Symanzik improvement}
143: 
144: The effects of a finite grid spacing can be understood in the language of
145: quantum field theory as arising from irrelevant operators appearing in the
146: lattice action. These operators are of higher dimension than the relevant 
147: operator of the continuum theory, Tr $F_{\mu\nu} F_{\mu\nu}$ and are 
148: multiplied in the action by dimensionful couplings proportional to (positive)
149: powers of the lattice spacing. For the simplest description of the gauge
150: action, proposed by Wilson and consisting of the trace of the plaquette, the
151: leading irrelevant operator appears at ${\cal O}(a^2)$. Odd powers are
152: prohibited by exact parity symmetry, which is preserved by the discretisation. 
153: 
154: Universality suggests that there is a broad class of lattice
155: representations of the continuum action, whose members all have the identical
156: continuum limit: QCD. This implies that different lattice actions can be 
157: engineered which have smaller contributions from higher-dimensional operators. 
158: For asymptotically free theories, like QCD and the $SU(3)$ Yang-Mills theory of
159: gluons, these discretisations of the action can be designed within the 
160: framework of perturbation theory. This concept is termed Symanzik improvement
161: \cite{Symanzik:1983dc}. 
162: 
163: The idea has been widely exploited in lattice calculations, both in
164: discretisations of gluon fields as well as the quark fields. For the gluon
165: action, taking appropriate linear combinations of traces of larger closed
166: loops, such as the $2\times 1$ rectangle means terms at ${\cal O}(a^2)$ can be
167: eliminated \cite{Luscher:1985zq, Lepage:1996ph}. This allows lattice 
168: investigations to be carried out with grid spacings as coarse as 
169: 0.25 fm, while remaining sufficiently close to the continuum limit to 
170: explore the physics reliably. 
171: 
172: \subsection{The anisotropic lattice}
173: 
174: At the point in the calculation where measurements are made, these coarse 
175: lattices present new disadvantages. The mass of a state is extracted from the 
176: decay of a two-point correlation function in Euclidean space
177: \begin{equation}
178:     C(t) = \langle \Phi(t) \Phi(0) \rangle \propto e^{-M t}
179: \end{equation}
180: where $\Phi(t)$ is a lattice operator that creates or annihiliates the state 
181: of mass $M$ at time $t$. This correlator falls exponentially as the
182: two operators are moved further apart, and for a heavy state (such as the
183: glueballs) the fall-off is rapid. Also, since the correlation function is being
184: estimated in a Monte Carlo simulation, the statistical variance of the
185: operators determines the accuracy to which these correlations can be measured.
186: For the glueballs, these operators are the traces of closed Wilson loops on the
187: underlying gauge fields and have rather large vacuum fluctuations, so a
188: critical signal-to-noise problem arises.  Since the operators can only be
189: measured on points of the grid, little information can be extracted from coarse
190: lattice measurements before the signal is lost in the noisy vacuum.
191: 
192: A solution that combines the economy of the coarse lattice with the good
193: resolution of a fine grid is to use an anisotropic lattice, where the temporal
194: lattice spacing is made small, while the three spatial dimensions are
195: discretised more coarsely. Note that time-like correlation functions are used
196: in extracting the masses of states. This method is particularly efficient for
197: glueball simulations \cite{Morningstar:1997ff}. The two natural length scales 
198: in the problem, the mass and size of the glueball set the optimal temporal and
199: spatial lattice spacings for numerical investigation.
200: 
201: Fig.~\ref{fig:meff} describes the output from just such an investigation. The 
202: figure shows the effective mass,
203: \begin{equation}
204:    a_t M_{\rm eff} = \frac{\ln\; C(t)}{\ln\; C(t+1)}
205: \end{equation}
206: for the scalar glueball measured on a highly anisotropic lattice, where the
207: ratio of scales is $a_s/a_t=6$. The operator, $\Phi$ is constructed from a
208: large basis set of the trace of closed Wilson loops built from smoothed link
209: variables. A single operator for each state is made by taking linear
210: combinations from the basis set of operators, with this combination chosen to
211: optimise the overlap with the state of interest. The graph shows data for the 
212: ground-state scalar and the first excited state with the same quantum numbers.
213: It illustrates the method is capable of precise determinations of the energies 
214: of both these levels. In the simulation presented in Fig.~\ref{fig:meff} these
215: energies are measured to $1\%$ statistical precision.
216: \begin{center}
217: \begin{figure}
218: %\setlength{\epsfxsize}{12cm}\epsfbox{scalar_meff.eps}
219: \includegraphics[width=12cm]{scalar_meff}
220: \caption{The effective mass plot of the scalar glueball channel, including data
221: for the first excited state.\label{fig:meff}}
222: \end{figure}
223: \end{center}
224: 
225: In Fig.~\ref{fig:dip}, the dependence on the lattice cut-off of the scalar
226: glueball mass in units of a physical scale $r_0$, derived from the static
227: potential \cite{Sommer:1994ce}, is presented. The data points
228: represented with crosses are from a range of simulations performed using
229: the Wilson plaquette action \cite{Michael:1989jr, Bali:1993fb}. The circles 
230: are from simulations with an 
231: anisotropic lattice action, improved according to the Symanzik programme with 
232: parameters determined from tree-level perturbation theory after the tadpole 
233: graphs in the lattice weak-coupling expansion have been resummed 
234: \cite{Morningstar:1997ff, Morningstar:1999rf}. The quality 
235: of a lattice action should be judged by how strongly the mass (in physical 
236: units) depends on the cut-off; ``better'' actions should show a weaker 
237: dependence on the lattice spacing. By this assessment, the Symanzik improved 
238: action is superior to the simplest action, however a strong dependence on the 
239: cut-off still remains. The scalar glueball mass (in units of $r_0^{-1}$) falls 
240: as the lattice spacing is increased to reach a minimum from which it begins to 
241: rise again.  We term this effect the ``scalar dip.''
242: 
243: \subsection{Curing the ``scalar dip''}
244: 
245: The lattice theory is not necessarily QCD; the physics of this theory is only 
246: recovered once the discretised version is ``close'' to the critical point on 
247: which QCD exists. If the space of lattice theories contains other critical 
248: points, simulations performed near to these points will be probing the 
249: properties of other continuum quantum field theories. Remarkably, this 
250: apparently abstract problem in non-perturbative renormalisation seems to arise 
251: in the simulation of $SU(3)$ Yang-Mills. 
252: 
253: It has been recognised for some time that there is a line of ``bulk'' 
254: first-order phase transitions in the two-dimensional plane of lattice theories 
255: in which both a coupling to the fundamental and adjoint representations of the 
256: link variables is made \cite{Heller:1995bz}. This line ends in a critical 
257: point at which some
258: unknown continuum theory resides. If the physics of the lattice theory close
259: to this critical point is investigated, it will be predominantly governed by
260: the properties of this other theory. One suggestion is that the continuum
261: theory at this critical point is a free scalar one, and if this was the case,
262: the mass of the scalar particle in the lattice theory would be artificially
263: light in comparison to the higher spin states. Precisely this artefact is
264: observed. 
265: 
266: The data represented by squares in Fig.~\ref{fig:dip} are from a new
267: discretisation of the gluon action on an anisotropic lattice. In this
268: prescription, the lattice action includes terms that trace over the 
269: link variables in the adjoint representation, as well as the fundamental. 
270: The links stored on the computer are in the fundamental representation, but the
271: trace in the adjoint can be computed using the identity
272: \begin{equation}
273:   \mbox{Tr }A(g) = \mbox{Tr }U(g) \;\mbox{Tr }U^\dagger(g) - 1,
274: \end{equation}
275: with $A(g)$ the adjoint representation of an element, $g \in SU(3)$ and $U(g)$
276: its corresponding fundamental representation. 
277: \begin{center}
278: \begin{figure}
279: % \setlength{\epsfxsize}{12cm}\epsfbox{scalar_dip.eps}
280: \includegraphics[width=12cm]{scalar_dip}
281: \caption{Extrapolating simulation data at finite lattice spacing to the
282: continuum limit.\label{fig:dip}}
283: \end{figure}
284: \end{center}
285: 
286: These simulation results are extremely encouraging and clearly show very weak 
287: lattice spacing dependence out to coarse spatial lattices ($a\approx 0.25$ fm).
288: A very reliable extrabolation to the continuum limit can be made.
289: 
290: \subsection{Spin on the lattice}
291: 
292: In a scattering experiment, the spin of resonances is determined from a
293: partial wave analysis. Revealing the spin of states in a lattice
294: calculation requires similar care.
295: Putting quarks and gluons onto a lattice breaks the continuum rotation group 
296: $SO(3)$ to the discrete cubic point group, $O_h$. As a result, states are no
297: longer classified according to a spin quantum number which describes the
298: irreducible representation (irrep) of $SO(3)$ they transform under.
299: They have instead one of five labels corresponding to the five irreps of $O_h$:
300: $A_1,A_2,E,T_1$ and $T_2$. 
301: Table~\ref{tab:subduce}
302: describes how the representation of $O_h$ subduced from spin irreps of 
303: $SO(3)$ subsequently decompose into irreps of $O_h$.
304: \begin{table}[h]
305: \begin{tabular}{c|ccccc}
306:   J & $A_1$ & $A_2$ & $E$ & $T_1$ & $T_2$ \\
307:   \hline
308:   0 &   1   &       &     &       &       \\
309:   1 &       &       &     &   1   &       \\
310:   2 &       &       &  1  &       &  1    \\
311:   3 &       &   1   &     &   1   &  1    \\
312:   4 &   1   &       &  1  &   1   &  1    \\
313:   \hline
314: \end{tabular}
315: \caption{The irreducible content of the representations of $O_h$ subduced from
316: the spin irreps of $SO(3)$}\label{tab:subduce}
317: \end{table}
318: To identify the spin of the state in the continuum,
319: degeneracies across lattice states must be found and compared to the table. For
320: example, tagging a spin 2 state on the lattice requires finding degenerate
321: states in the $E$ and $T_2$ channels (and no others) in the continuum limit. 
322: For the
323: scalar state, this procedure is quite straightforward; a state in the trivial
324: representation, $A_1$, which is not degenerate with any other state in the
325: spectrum can be uniquely identified with a continuum $J=0$ scalar. This spin
326: analysis was carried out in Ref.~\cite{Morningstar:1999rf} for the glueball 
327: spectrum, and the spins of many continuum states were identified. 
328: 
329: \section{The $SU(3)$ Yang-Mills spectrum}
330: 
331: Following the procedures outlined in the previous sections, the simulations of 
332: the Yang-Mills theory of strongly interacting gluons (without quarks) have 
333: reached high levels of precision. Fig.~\ref{fig:spectrum} shows the lightest
334: six states in this spectrum of gluon bound-states. Note that both the 
335: ground-state and first-excited state in the scalar channel has been determined 
336: more reliably after removing the artefacts of the scalar dip and performing a
337: continuum extrapolation. The masses of the scalar glueballs are consistent
338: within errors with earlier determinations. 
339: \begin{center}
340: \begin{figure}[h!]
341: %\setlength{\epsfxsize}{12cm}\epsfbox{spectrum_2.eps}
342: \includegraphics[width=12cm]{spectrum_2}
343: \caption{The lightest six states in the spectrum of the $SU(3)$ Yang-Mills
344: theory.\label{fig:spectrum}}
345: \end{figure}
346: \end{center}
347: 
348: \section{Current and future directions}
349: 
350: While the simulation of the Yang-Mills theory is now well understood, the
351: inclusion of quantum fluctuations of quark fields is significantly more
352: difficult. Since the Grassmann algebra of fermions in a path integral can not 
353: be handled directly by computers, the quark bilinear in the QCD action must be 
354: integrated out analytically, leaving a non-local effective action in the 
355: gluons alone. The non-local nature of these interactions means the simpler 
356: Monte Carlo algorithms used to generate ensembles of gauge field configurations 
357: for the Yang-Mills theories become inefficient. More sophisticated techniques 
358: are used, but they add a large numerical overhead; QCD simulations including 
359: quark dynamics are a factor of 100-1000 times more expensive than those of the 
360: Yang-Mills theory. 
361: 
362: This issue of including the quark field dynamics remains a topic of a good deal
363: of research in the lattice community and optimal simulation strategies are
364: still actively under investigation. It seems likely that within the near
365: future, the physics of glueballs that can decay into and mix with quark mesons
366: will be investigated. Progress in this direction has begun by other lattice
367: groups \cite{Hart:2002sp} although lattice calculations involving unstable 
368: particles are in their infancy and remaind a challenging topic of research. 
369: Also, extending the anisotropic lattice technology, 
370: which proved so useful for scanning the pure gauge theory, to QCD simulations 
371: with dynamical quarks is under investigation.
372: 
373: \section{Conclusions}
374: 
375: Scalar mesons seem to constantly challenge theorists and phenomenologists by
376: turn. At this meeting we heard of the difficulties in finding a convincing
377: picture of the broad range experimental data and in this article, some of the 
378: quite distinct problems presented by the scalar glueball to lattice theorists 
379: have been outlined. Significant progress in this field is still being made and 
380: the Yang-Mills theory has now been mapped out. Full QCD dynamics, with the quark
381: fields playing their role, is a major challenge under investigation by
382: many members of the lattice community \cite{Lat03}. At the same time, more 
383: detailed properties of the gluonic theory are now being examined, including 
384: calculations of the chromoelectric and magnetic matrix elements 
385: \cite{Dong:Lat03} and the structure form factors. 
386: 
387: \bibliography{trinlat_bib/trinlat}
388: 
389: \end{document}
390: