1: %%%%%%%%%%%%%%%%%
2: \section{On strings in higher representations and ${1\over N}$ corrections
3: \la{sec:string_high_rep}}
4: %%%%%%%%%%%%%%%
5: We now leave the discussion of the adjoint-monopole-gas model and
6: discuss the properties of $k$-strings from the point of view of the
7: large-$N$ expansion.
8:
9: Standard arguments on large $N$ SU($N$) gauge theory~\cite{hooft,witten_baryons},
10: based on the planarity of Feynman diagrams and the (assumed) confinement of color,
11: imply that gauge invariant states have masses of order $N^0+N^{-2}$, with
12: a width of order $N^{-2}$. Also, no bound state of color singlet constituents
13: survives the large \n limit: the theory is expected to be a theory of
14: free `hadrons'.
15:
16: It is interesting to consider, at large but finite number of colors,
17: precisely those states whose wavefunction contains a
18: significant component which is a direct product of color singlet pieces.
19: Phenomenology provides a number of potential examples.
20: A classic example would be the deuteron, a very loosely bound state
21: lying only a few MeV under the nucleon-nucleon threshold. Another interesting
22: though less firmly established case is the $f_0(980)$ meson,
23: whose wavefunction has been discussed in terms of a mixture of a kaon-kaon molecule
24: and a four-quark state~(\cite{Close:2002zu} and ref. therein).
25: Again, the state is only a few MeV under the two-kaon threshold.
26:
27: At a more theoretical level, there are examples in the pure SU($N$) gauge theory
28: in three and four dimensions.
29: Consider the theory defined on a finite (but large: $L\gg 1/T_c$) hypertorus.
30: In addition to glueballs, the spectrum contains `torelon' states
31: (whose mass we denote by $m_k(L)$)
32: which transform non-trivially under the centre symmetry $Z_N$.
33: The sectors of different $\cal N$-ality are protected by this global symmetry.
34: Thus for $N\geq4$, one may ask whether two fundamental torelons can form
35: a bound state lying under the threshold $2m_{k=1}(L)$. That this is indeed
36: the case was first numerically demonstrated from first principles in
37: the work~\cite{lucini}. Further, at large $L$
38: the states are string-like and one can ask what the ratios of their string
39: tensions are (we may use the fundamental `$k=1$' string tension as the reference).
40: An alternative formulation of the problem would consider the strings to
41: be open and attached to static sources in the appropriate
42: representation~\cite{Deldar:1999vi}.
43:
44: For simplicity, we now focus on the $k=2$ sector;
45: for $N \geq 4$, the screening of the string is forbidden by the centre symmetry.
46: In our view, the first question to settle in the context of the large-$N$
47: expansion is, `What is the $1/N$ power of the leading correction to
48: the planar limit result $\sigma_2=2\sigma_1$?'.
49: Since $m_2(L)$ lies under the threshold $2m_{k=1}(L)$ at all $L$~\cite{lucini},
50: the question arises whether one should think of the $k=2$ torelon as a
51: weakly bound state of two $k=1$ torelons,
52: or if the colour structure gets completely rearranged
53: into a single `unfactorisable' color singlet piece.
54: In the nucleon-nucleon system, the analogous question is whether the deuteron is
55: primarily a bound state of two nucleons, or a 6-quark state.
56: In the first case, considered in~\cite{armoni},
57: the long-distance attractive force between the two $k=1$ strings
58: will be driven by the exchange of the lightest
59: ($0^{++}$) glueball, while the short distance force is essentially given by
60: two-gluon exchange. Both effects are indeed~\cite{armoni}
61: suppressed by $1/N^2$ with respect to the free propagation of two $k=1$ strings.
62: Regarding the second configuration, the simplest classical string configuration
63: is that of a single string winding twice around a cycle of the hypertorus.
64: At large \n, the energy of such a configuration is expected to approach
65: threshold from below at a $1/N^2$ rate.
66:
67: At finite \n and asymptotically large $L$ however,
68: we are in presence of two almost degenerate configurations lying near threshold.
69: It is therefore imperative to consider
70: the mixing effects between these two configurations.
71: To keep the discussion as simple as possible, we may keep the transverse spatial
72: dimensions $L_\perp$ finite, so as to separate two-torelon `scattering states' from
73: the weakly bound states we discuss by a finite, fixed gap. As \n is increased,
74: this transverse volume can be increased as well without affecting the validity of
75: our treatment of the $k=2$ sector as a 2-level system.
76:
77: We suppose, following~\cite{finivol}, that the Hamiltonian of the SU($N$)
78: gauge theory can be expanded in inverse powers of $1/N$:
79: \be
80: H(L,N)=\sum_{k=0}^\infty \frac{H_k(L)}{N^k}.
81: \ee
82: The existence of the t'Hooft limit implies that $H_o(L)$ has the same eigenvalues as
83: the Hamiltonian of the SU($\infty$) theory in the same spatial volume.
84: We consider now the $k=2$ flux tubes
85: winding around a cycle of the torus as a quantum mechanical two-state system,
86: as was done in~\cite{finivol} for the case of the scalar glueball --
87: adjoint Polyakov loop system in intermediate volume.
88: Consider on the one hand the state made of two $k=1$ non-interacting
89: closed fundamental strings, and on the other a single fundamental
90: string with winding number 2; in this basis
91: $H_o$ reads $H_o=2m_{k=1}{\bf I}_{2\times2}$.
92:
93: The `perturbation' describes the deviations from the planar limit. On the diagonal,
94: the corrections are ${\cal O}(1/N^2)$. Indeed, the attractive potential between
95: two fundamental torelons is suppressed by the product of two 3-vertices
96: each of which carries a $1/N$ factor. On the other hand, the amplitude of the
97: transition from one of our basis states to the other only contains one such vertex,
98: and therefore the off-diagonal element of our $2\times2$ hamiltonian is
99: ${\cal O}(1/N)$.
100: The perturbation hamiltonian in our basis reads:
101: \be
102: % \Delta H= \left(\begin{array}{c@{~}c} \frac{-\bar h_1}{N^2}&\frac{\bar h}{N}\\
103: % \frac{\bar h}{N} & \frac{-\bar h_2}{N^2} \end{array}\right)
104: \Delta H= \left(\begin{array}{c@{~}c} -\bar h_1/N^2 & \bar h/N\\
105: \bar h/N & -\bar h_2/N^2 \end{array}\right)
106: \la{eq:DH}
107: \ee
108: with $\bar h_1$, $\bar h_2$ and $\bar h$ of order $N^0$. It is clear that
109: to leading order in $1/N$, the resulting
110: energy eigenstates are now the symmetric and anti-symmetric linear combinations
111: of our basis states. The associated energies are
112: $E_A=2m_{k=1}-\frac{\bar h}{N}+{\cal O}(1/N^2)$ and
113: $E_S=2m_{k=1}+\frac{\bar h}{N}+{\cal O}(1/N^2)$. We thus reached the perhaps
114: surprising conclusion that the corrections to the mass of the lightest
115: $k=2$ string are of order $1/N$. There is one state below threshold and one above,
116: situated symmetrically about the threshold energy, up to $O(1/N)$ corrections.
117: We note that $\bar h$ (as well as the $\bar h_i$)
118: is expected to grow proportionally to $L$ at large $L$,
119: since the breaking of the string can occur at any point along the string,
120: so that the ratio in the torelon masses directly translates into the ratio of the
121: string tensions.
122:
123:
124: A caveat particularly relevant to Monte-Carlo simulations
125: is that in all the considerations above
126: we have supposed the strings to be long enough
127: to be able to identify the ratios of string tensions ratios
128: with the ratio of torelon masses. Since the $1/L$ string correction~\cite{luscher}
129: lowers the energy of the string, it induces a repulsive force between
130: two fundamental torelons at finite $L$~\cite{finivol}.
131: Since the binding energy of the strings
132: reduces at large \n, $L$ must be increased so that the condition
133: \be
134: \sigma_1 L^2 \gg N \la{eq:long_s_cond}
135: \ee
136: is satisfied to ensure that the ratio of torelon loops yields the correct
137: $N$-dependence of the string tension ratios.
138: If the large $N$ limit is taken at finite $L$, the ratio of
139: $k=2$ to $k=1$ torelon masses will approach 2 with $1/N^2$ corrections, because
140: a mixing amplitude only affects the energy spectrum at leading order if the
141: `unperturbed' states are degenerate (to that order). And indeed, the
142: two classical-string configurations we took as unperturbed states have different
143: $1/L$ corrections: if $ \gamma_1$ is the L\"uscher coefficient of the fundamental
144: string, the direct-product configuration of two $k=1$ strings
145: has a $2\gamma_1/L$ correction, while the fundamental string with winding number 2
146: admits a $\gamma_1/2L$ string correction.
147: As all lattice simulations so far~\cite{lucini,debbio,lucini04},
148: ours are done in the regime $N < \sigma_1 L^2 < N^2$.
149: The second inequality implies that the energy gap
150: $(\bar h_1-\bar h_2)/N^2$ is parametrically smaller than the string
151: vibrational excitations.
152:
153: \eq\ref{eq:long_s_cond} however also implies that the vibrational
154: excitations of the strings are separated by $4\pi/L$ gaps which are parametrically
155: much smaller than the mixing energy $\bar h/N$.
156: We have thus neglected the matrix elements of the Hamiltonian $H_1$
157: between the two states that we focused on and
158: the vibrationnally excited states, since we reduced the
159: diagonalisation problem from the full Hilbert space to the space spanned
160: by these two states. Although the neglected matrix elements
161: can modify the splitting pattern around the threshold, the
162: mixing between the string ground states will be enhanced relatively to the
163: other mixings if the condition $\sigma_1 L^2 \ll N^2$ holds. In any case,
164: the parametric size of these matrix elements is
165: $\sigma L/N$ and so we must expect the corrections to the $k=2$ string tension
166: to be of that order.
167:
168:
169: In summary, the predictions of the two-state mixing model are:
170: \begin{enumerate}
171: \item the energy eigenstates are the anti-symmetric and the symmetric linear
172: combinations of the direct-product configuration of two $k=1$ strings
173: and the fundamental string with winding number 2.
174: \item they are split symmetrically around the threshold energy $2m_1$ (up to
175: $O(1/N)$ corrections).
176: \item the splitting energy itself is of order $1/N$.
177: \end{enumerate}
178: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
179: \subsection{Static potentials}
180: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
181: If one considers open strings attached to static `quarks',
182: the argument takes a slightly different form. The relevant quantities
183: here are the static potentials between colour sources
184: in \emph{irreducible} representations of SU($N$).
185:
186: The k=1 string binds a quark and a distant antiquark together.
187: Similarly the k=2 configuration can be viewed as two (weakly interacting) strings
188: each joining one of the quarks to one of the antiquarks.
189: If we number the quarks by 1 and 2, and the antiquarks by $\bar 1 $ and $ \bar 2$,
190: then there are two classical string configurations which are exactly degenerate:
191: the configuration where $1$ is attached by a string to $\bar 1$ and
192: $2$ is attached to $\bar 2$, and the other where $1$ is attached to $\bar 2$
193: and $2$ to $\bar 1$. However,
194: the interaction between the strings can take one configuration into the other.
195: Therefore a splitting occurs between the symmetric and anti-symmetric
196: linear combinations, corresponding to the static potential
197: splitting between the $k=2$ symmetric
198: and anti-symmetric irreducible representations of SU($N$).
199: There is however general agreement that screening of
200: the static sources through virtual gluons implies that the string tension
201: obtained in either representation at large enough separations
202: is the same; although it can be difficult
203: to demonstrate this in Monte-Carlo simulations.
204:
205: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
206: \subsection{A caveat on the implications of factorisation at large $N$}
207: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
208: The standard way to extract the static potential for
209: fundamental charges, namely by measuring the expectation value
210: of a rectangular Wilson loop of size $R\times T$, $T\gg R$, can be generalised
211: to extract it for any representation~\cite{Deldar:1999vi}. In particular, the simplest way
212: to obtain a representation of ${\cal N}$-ality $k=2$ is to take the real part of
213: the square of $W(R,T)$, the trace of the fundamental Wilson loop.
214: At finite $R,~T$, the factorisation property of gauge invariant operators
215: (see for instance \cite{Das:1984nb})
216: then implies that the expectation value of this operator is given by
217: \be
218: \< W(R,T)^2 \>=\< W(R,T) \>^2~(1+{\cal O}(1/N^2)).\quad \mathrm {(factorisation)}
219: \ee
220: On the other hand, if we consider small separations $R$, asymptotic freedom
221: implies that the short-distance potential in an irreducible representation
222: ${\cal R}$ is given by $C_{\cal R}\alpha_s/R$.
223: The symmetric $k=2$ representation has $C_{\cal R}=C_S=2(N+2)C_F/(N+1)$, while
224: the $k=2$ anti-symmetric has $C_A=2(N-2)C_F/(N-1)$. In particular, for the
225: fundamental representation, it is
226: \be
227: W(R,T)=\exp{\left(-\frac{\bar \alpha T}{2R}\right)} ~+~ {\cal O}(1/N^2),
228: \qquad R\ll \sigma^{-1/2}.
229: \ee
230: The operator $W^2(R,T)$ belongs to a representation that can be reduced
231: into the symmetric and anti-symmetric. Therefore, if we take the
232: $T\rightarrow\infty$ limit, the potential energy of
233: the anti-symmetric representation dominates the expectation value of $W^2(R,T)$:
234: \be
235: \lim_{T\rightarrow\infty} \<W^2(R,T)\>
236: \propto e^{-C_A\alpha_s CT/R}=\<W(R,T)\>^2 ~e^{\frac{\bar\alpha T}{NR}}
237: ~(1+ {\cal O}(1/N^2)), ~~ R\ll \sigma^{-1/2}.
238: \ee
239: Thus tree-level perturbation
240: theory contradicts the large-$N$ counting rules concerning the leading corrections
241: to factorisation. The origin of the paradox lies in the straightforward
242: $T\rightarrow\infty$ limit necessary to filter out the ground state.
243: If the contribution from the symmetric representation is kept,
244: the large $N$ limit of the small-$R$, large-$T$ Wilson loop
245: $\<W^2(R,T)\>$ is given by
246: \be
247: \lim_{T\rightarrow\infty} \<W^2(R,T)\> \propto
248: \<W(R,T)\>^2 ~\cosh\left(\frac{\bar\alpha T}{NR}\right)
249: ~(1+ {\cal O}(1/N^2)),
250: \ee
251: which, at fixed $T$, has $1/N^2$ corrections to the planar limit result.
252:
253: What have we learnt? The large-$N$ factorisation property
254: does not necessarily imply that the lowest energy state of a `meson'
255: made of a static colour source in a certain representation and its anti-source
256: has ${\cal O}(1/N^2)$ corrections,
257: \emph{because other representations of same ${\cal N}$-ality become degenerate
258: with it in the $N\rightarrow\infty$ limit}. Since string tensions are extracted
259: from the lowest energy at large $R$, the same caveat applies to them.
260: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
261: \subsection{Strings in open and closed form\la{sec:open-closed}}
262: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
263: We finish with a remark on the relation between different representations
264: of same ${\cal N}$-ality and excited states in the open and the closed string sectors.
265: To that end it is useful to consider the correlator of Polyakov loops of length $L$
266: $\< P_{\cal R}(0) P_{{\bar{\cal R}}'}(\vec x)\>$. This expectation value is interpreted
267: (from the point of view of the transfer matrix along the dimension of size $L$)
268: as the free energy of the system in the presence of two static charges in the given
269: representations. When ${\cal R} = {\cal R'}=N$, the fundamental representation,
270: the heavy-heavy bound state can a priori be in the adjoint or the singlet representation
271: (in SU(3): $ 3\otimes \bar 3 = 8 \oplus 1$). Now, it is believed that only bound states
272: in the singlet representation have a finite free-energy in the confined phase.
273: That means that if the heavy charges themselves are not in the singlet representation,
274: virtual gluons will try and screen the chromoelectric field emanating
275: from this coloured bound state until it is a singlet again. Since the gluons are in the adjoint
276: representation, they can screen the configurations of the heavy-quark bound state that are
277: in the adjoint representation, albeit at a certain energy cost.
278: On the other hand they cannot screen a single heavy
279: quark, and the latter therefore has an infinite free energy.
280:
281: Suppose we want to determine the static potential for sources in all possible
282: representations of SU($N$) (not necessarily irreducible)
283: of a given ${\cal N}$-ality $k$ and up to a given size. Clearly
284: it is sufficient to determine the Polyakov loop correlators between the irreducible
285: representations obtained in the decomposition of the direct product representation of $k$
286: quarks. The question then arises whether the cross-correlations (i.e. for
287: ${\cal R}\neq {\cal R'}$) between Polyakov loops in these irreducible representations
288: vanish or not. If they do, it implies that the energy eigenstates are in definite
289: irreducible representations.
290:
291: Consider the $k=2$ case. The direct product of two fundamental representations
292: decompose into a symmetric and an anti-symmetric representation: $N\otimes N=A\oplus S$.
293: In the most familiar case of SU(3), the
294: anti-symmetric representation is nothing but the $\bar 3$ (anti-fundamental):
295: $3\otimes 3 = 6 +\bar 3$. So we are asking whether $\< P_{S}(0) P_{\bar A}(\vec x)\>$
296: has to vanish. We have $6\otimes 3 = 10 \oplus 8$, so that virtual gluons can
297: screen the adjoint piece, thus ensuring that the free energy of this system is finite.
298: So in general these cross-correlations do not vanish. It is easy to see (using Young
299: tableaux) that in SU($N$) the adjoint representation appears exactly once
300: in the decomposition of $S\otimes \bar A$.
301: However since gluons have to screen the heavy-heavy system, the
302: $\< P_{S}(0) P_{\bar A}(\vec x)\>$ are $1/N^2$ suppressed at large $N$.
303: Let us now see what this conclusion implies for the determination of the
304: $k$-string tensions in the open and the closed string sectors.
305:
306: In order to study the lightest open string, one may
307: in principle choose to immerse any one static source of the relevant ${\cal N}$-ality
308: in the system, since
309: for large enough $R$ the sources are expected to be screened down to the
310: representation with the smallest string tension. Once the linear behaviour of
311: $V(R)$ with the latter slope sets in, the differences between the static potentials
312: in irreducible representations of same ${\cal N}$-ality are expected to become
313: only weakly $R$-dependent (they correspond to `gluelump' masses~\cite{michael}).
314: For long enough strings, the lowest excitations of any of these static `mesons'
315: correspond to the lowest excitations of that string, which come in gaps of order $1/R$.
316: In short, there is at most one stable open string for a given ${\cal N}$-ality.
317:
318: It is also possible to interpret the Polyakov loop correlator
319: with a transfer matrix along the direction $\vec x$. One is then measuring the
320: spectrum of states of the gauge theory which carry a winding number with respect
321: to a cycle of the hypertorus of length $L$.
322: Of course, since the Polyakov loop correlator has a unique asymptotic area law,
323: the coefficient in front of the area defines both the string tension in the
324: open as in the closed string sector. Just as in the open string case,
325: there cannot be more than one stable string per $\cal N$-ality
326: because of the screening by gluons. A simple picture~\cite{lucini} is that
327: virtual gluons screen the unstable string down to the stable one
328: and propagate along it until they annihilate around the cycle of the torus.
329:
330: For long enough torelons, the lowest closed-string excitations are again
331: expected to be string-like, i.e. coming in $1/L$ gaps.
332: There can be resonant states of the torelons
333: (lying above the $k$-torelon threshold) whose energies grow linearly with $L$.
334: It is then natural to associate them with meta-stable strings.
335:
336: What we inferred about the cross-correlations between different irreducible
337: representations above tells us that the energy eigenstates do \emph{not} in
338: general belong to irreducible representations of SU($N$), although
339: the mixing between them is suppressed (at least in the $k=2$ case) by $1/N^2$.
340: %
341: %
342: %%%%%%%%%%%%%%%
343: \section{Lattice simulations\label{sec:lattice}}
344: %%%%%%%%%%%%%%%
345: We extract string tensions in the three-dimensional SU(8) gauge theory
346: from the masses of `torelons`, gauge
347: invariant states transforming non-trivially under the
348: $Z(N)$ symmetry of the action; they wind around one spatial cycle of a
349: the hypertorus. These masses are extracted from the exponential
350: decay of correlation functions at `large' Euclidean time.
351: To enhance the signal-to-noise ratio, we use fuzzing techniques
352: in the construction of our operators as described in~\cite{lucini04}.
353: The correlation
354: functions are measured on gauge configurations generated by a Monte-Carlo
355: program. We use the original Wilson action~\cite{Wilson:1974sk}.
356: The configuration is updated by sequences of `sweeps'.
357: One sweep consists of updating all links
358: by performing either a heat-bath (HB)~\cite{kenpen} or an over-relaxation
359: (OR)~\cite{adler} step on $N(N-1)/2$ of its $SU(2)$ subgroups~\cite{cm}.
360: The ratio of HB:OR is 1:3,
361: and we typically perform a sequence of 1 HB and 3 OR between measurements.
362: We use a 2-level algorithm~\cite{2leva} as described in~\cite{2levb}.
363: The latter reference also contains a detailed comparison of efficiency of the
364: ordinary 1-level and 2-level algorithms.
365: The number of measurements performed at fixed time-slices was
366: 800 at $\beta=115$, 200 at $\beta=138$ and 40 at $\beta=172.5$.
367: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
368: \subsection{String corrections\la{sec:string_corr}}
369: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
370: Consider the Euclidean gauge theory on a $L\times L\times T$
371: hypertorus, with cycles of length $L$. The gauge-invariant states with
372: winding number $k\neq 0$ around one spatial
373: cycle of the hypertorus are called torelons. In the Hamiltonian
374: language they are created by spatial Polyakov-loop operators
375: with ${\cal N}$-ality $k$; a description of the operators used
376: can be found in appendix C.
377: If the dynamics of a torelon state of length $L\sqrt{\sigma}\gg 1$
378: is described by an effective string action, then the expression
379: for its mass as a function of its length reads
380: \be
381: m(L)=\sigma L \left[1- \frac{\gamma}{\sigma L^2}+ {\cal O}\left(
382: \frac{1}{\sigma L^2}\right)^2\right], \la{eq:string_correction}
383: \ee
384: where $\gamma$ is a numerical coefficient of order one which
385: only depends on the universality class of the string~\cite{luscher}.
386: Recent accurate numerical results~\cite{Luscher:2002qv,lucini}
387: show that the flux-tube
388: in the fundamental representation belongs to the bosonic string class.
389: In the case of a torelon this implies that
390: \be
391: \gamma = \gamma_b \equiv \frac{(D-2)\pi}{6}. \la{eq:bosonic_value}
392: \ee
393:
394: In general, if \eq\ref{eq:string_correction} holds,
395: the ratio of the lightest $k$-torelon mass to the $k=1$ torelon mass
396: is given by
397: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
398: \be
399: \frac{m_k}{m_1}(L)=\frac{\sigma_k}{\sigma_1}+
400: \frac{\alpha_k}{\sigma_1L^2}+ {\cal O}\left(
401: \frac{1}{\sigma_1 L^2}\right)^2~,\qquad
402: \alpha_k=\frac{\gamma_1\sigma_k}{\sigma_1}-\gamma_k
403: \la{eq:ratio_formula}
404: \ee
405: The sign of $\alpha_k$ is of interest. If the $k$-string is a weakly bound
406: state of $k$ fundamental strings, then one would expect $\gamma_k=k\gamma_1$
407: and therefore $\alpha_k=-(k-\frac{\sigma_k}{\sigma_1})\gamma_1<0$.
408: If one the other hand the
409: fluctuations of the $k$ fundamental strings are `in phase', then
410: the number of degrees of freedom on the worldsheet of the $k$-string
411: is the same as for the fundamental string, hence $\gamma_k=\gamma_1$
412: and $\alpha_k=(\frac{\sigma_k}{\sigma_1}-1)\gamma_1 > 0$.
413:
414: At each lattice spacing, we measured the masses of the torelon states
415: of at least three different lengths. In practice,
416: we use two asymmetric lattices of the type $L_1\times L_2\times L_t$
417: and $L_2\times L_3\times L_t$. In this way, we obtain three different lengths of the
418: torelon, and we can also check for any dependence on the transverse size
419: of the lattice by
420: comparing the mass obtained for the torelon of length $L_2$ on the two lattices.
421: The $L_i$ range from 1.4fm to 3fm, if we set the scale by $\sqrt{\sigma_1}=440$MeV.
422: This is longer than what has been normally measured so far, and is made possible
423: by the use of the two-level algorithm. We use \eq\ref{eq:string_correction} to obtain
424: the fundamental string tension by fitting $m(L)/L$ with a linear function in $1/L^2$.
425: The intercept yields the string tension; the slope gives the L\"uscher coefficient.
426: Whether the functional form~(\ref{eq:string_correction}) successfully describes
427: the leading deviation from constant linear mass density is
428: controlled by the $\chi^2$ of the fit.
429:
430: Systematic errors play an important role in comparing the numerical
431: data to model predictions. In an attempt to get them under control
432: we propose two separate ways to extract the ratios of string tensions
433: (we refer to the first method as the `unconstrained' one,
434: and the second as the `constrained' one).
435: In practice, having learnt from the pros and contras of both data analyses,
436: we present our final, `educated' analysis in section~\ref{sec:final_anal}.
437:
438: \paragraph{1.}
439: Firstly the ratios of torelon masses $m_k(L)/m_1(L)$ are fitted according
440: to \eq\ref{eq:ratio_formula} with
441: a linear function in $1/L^2$, and the intercept gives us the
442: ratio $\frac{\sigma_k}{\sigma_1}$. In this way, we need make no assumption
443: about the values of the coefficients $\gamma$ corresponding to the
444: different representations; in particular, the different strings could
445: have different coefficients $\gamma_k$. Finally, these string ratios
446: are extrapolated to the continuum, $a\rightarrow 0$, in a standard way.
447:
448: \paragraph{2.}
449: The second analysis will \emph{assume} that all $k$-strings
450: belong to the bosonic class. Consequently, we can extract the string
451: tension ratio from \eq\ref{eq:ratio_formula} using the estimate
452: $\alpha_k\simeq \frac{m_k\gamma_1}{m_1}-\gamma_k$
453: with $\gamma_k=\gamma_1=\gamma_b$ at every $L$. The estimates of the
454: ratios obtained at different $L$ are then simply averaged,
455: as long as they are compatible with eachother, to produce the estimates
456: of the string-tension ratios. If the $\chi^2$ of the average is large,
457: we drop the smallest $L$ until an acceptable $\chi^2$ is reached.
458: The continuum limit is then taken.
459:
460: \paragraph{}
461: The multi-level algorithm allows us to
462: apply the variational method~\cite{mart_var} on the correlation matrices
463: at $t\geq2a$ of Euclidean time separation, an improvement
464: over the traditional where the method is usually unstable unless
465: $t=0$, although the method really finds its justification
466: when applied at large $t$.
467:
468: Having said that,
469: we note that this work constitutes the first attempt to extract the $k=4$
470: string tension from Monte-Carlo simulations, and should be regarded
471: as exploratory in that sector.
472: Indeed we found that the variational method~\cite{mart_var}
473: generally became unstable if all five operators listed in
474: appendix C were fed in the generalised eigenvalue problem.
475: As a consequence only three or four of the five types of measured operators
476: (at the `best' level of smearing-blocking) were finally employed.
477: This and the fact that we only have a
478: short range in Euclidean time to identify the mass plateau, due to the
479: rapid fall-off of the signal, means that the $k=4$ string tension
480: has a significant systematic error attached to it. For the lower $k$
481: states, these problems are less accute and we are much more confident
482: about their mass estimates.
483:
484: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
485: \subsection{Data analysis\la{sec:analysis}}
486: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
487: We give the masses of the lightest spatial torelons of each ${\cal N}$-ality
488: in \tab\ref{tab:m_T};
489: \tab\ref{tab:excited} gives estimates of the first-excited
490: torelon mass in the $k=2$ sector, that will be discussed below.
491: Within the range considered ($L\simeq1.9$fm,
492: $ 0.8L\leq L_\perp \leq 1.2 L$),
493: we certainly find no dependence of the $k=1$ torelon masses
494: on the transverse size. There is also no statistically significant
495: variation of the lightest higher-$k$ torelon masses.
496: Transverse size corrections are expected to be suppressed by a power
497: of $1/L$ varying continuously with $L_\perp$,
498: but greater than 3~\cite{Meyer:2005px}.
499:
500: We show on \fig\ref{fig:loc_eff_mass} the local effective mass of the
501: correlators in the $k=1$ and $k=2$ representations.
502: We emphasize that the variational method, which yields (quasi-)orthogonal states,
503: automatically picks out the symmetric and anti-symmetric linear combinations
504: (within very small fluctuations on the coefficients). We shall come back
505: to this point in the discussion below, section~\ref{sec:excited}.
506: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
507: \subsubsection{Setting the scale}
508: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
509: Although one could choose the (dimensionful) coupling to set the scale,
510: we prefer to use $\sqrt{\sigma_1}$ for this purpose. We extract the
511: fundamental string tension in lattice units
512: at each of our three lattice spacings by linearly extrapolating
513: the torelon mass per unit length, $m_T/g^4L$ as a function of $1/(g^4L^2)$,
514: to infinite $L$. This is illustrated by \fig\ref{fig:polyas} in the case
515: $\beta\equiv\frac{2N}{ag^2}=138$.
516: The resulting string tensions are given in \tab\ref{tab:ratios}.
517: We are able to extract the coefficient of the $1/R$ string correction with
518: moderate accuracy; it is also given in \tab\ref{tab:ratios}.
519: The coefficients we obtain are within 1.3 standard deviations of the bosonic
520: string value.
521:
522: Similarly, we can extract the $k=2$ string tension and its string correction
523: coefficient $\gamma_2$ (\fig\ref{fig:polyas}, bottom plot). It is clear however that the
524: accuracy of the data does not allow us to estimate $\gamma_2$.
525: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
526: \subsubsection{Unconstrained extrapolations\la{sec:unconstrained}}
527: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
528: In this analysis, for each lattice spacing
529: we extrapolate the ratios of $k$-torelon masses to $L=\infty$,
530: assuming $1/(g^2L)^2$ corrections.
531: %%% This is illustrated on \fig\ref{fig:ratios_sig_u}.
532: In most cases, we have three torelon lengths to extrapolate. For the
533: intermediate length, where we have two statistically independent and compatible
534: values obtained at different transverse sizes of the spatial lattice,
535: the average (weighted by the inverse square of the statistical error) of the two values
536: was taken, whilst keeping the smaller of the two errors.
537: In the ratio of the $k$-torelon to the $k=1$ torelon mass obtained in the same
538: simulation, we checked in several cases that the error bars obtained by assuming
539: statistical independence do not differ by more than $10\%$ from the jacknife
540: values of the error bars; the former are then used in the following.
541:
542: We note that the results of these extrapolations done at different lattice spacings
543: are in fact consistent within error bars (see \tab\ref{tab:ratios});
544: it appears that finite lattice spacing effects are much smaller than the finite
545: string-length effects in our data set.
546: The $\chi^2$ of each of these fits are good (smaller than 1),
547: except for the extrapolation of the $\sigma_2/\sigma_1$ at $\beta=172.5$,
548: where $\chi^2=3.0$. Since the $L=\infty$
549: extrapolated value is entirely consistent with that obtained at the other values
550: of $\beta$, we attribute this to a statistical fluctuation and, perhaps, a slight
551: underestimation of the error bars (due to the neglect of the sort of systematic
552: errors mentioned at the end of section~\ref{sec:string_corr}).
553:
554: Now extrapolating these string tension ratios
555: to the continuum (assuming ${\cal O}(\sigma_1 a^2)$ discretisation errors),
556: we obtain $\sigma_2/\sigma_1 = 1.701(77)$, $\sigma_3/\sigma_1 = 2.31(16)$
557: and $\sigma_4/\sigma_1 = 1.96(23)$.
558: % \ba
559: % \sigma_2/\sigma_1 &=& 1.701(77) \nn
560: % \sigma_3/\sigma_1 &=& 2.31(16) \qquad{\rm preliminary!} \la{eq:ratios} \\
561: % \sigma_4/\sigma_1 &=& 1.96(23).\nonumber
562: % \ea
563: The $\chi^2$ of these fits are smaller than 1. The final error bars have blown
564: up due to a somewhat small level-arm in the continuum extrapolation.
565: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
566: \subsubsection{Constrained analysis\la{sec:constrained}}
567: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
568: In this independent analysis, we assume the validity of \eq\ref{eq:string_correction}
569: with $\gamma$ given by the bosonic string value \eq\ref{eq:bosonic_value} to extract
570: the string tensions at finite $L$ (neglecting the ${\cal O}(1/L^4)$ terms);
571: see the string tension ratios in \tab\ref{tab:ratios_c},
572: where again statistical errors have been added in quadrature.
573: In most cases, these
574: ratios are consistent with being independent of $L$ for $L\geq 1.4$fm.
575: The exceptions concern the $k=2$ string at the two smaller lattice
576: spacings (due to the accuracy of the data),
577: where we drop the smallest $L$ in our average.
578: We note that, compared to the values of the unconstrained analysis
579: (\tab\ref{tab:ratios}), the ratios are systematically larger.
580: The ratios for the $k$-strings in the continuum limit now are:
581: $ \sigma_2/\sigma_1 = 1.776(33) $, $ \sigma_3/\sigma_1 = 2.210(50) $
582: and $ \sigma_4/\sigma_1 = 2.282(63)$.
583: %%%% (\fig\ref{fig:extrapol}):
584: % \ba
585: % \sigma_2/\sigma_1 &=& 1.776(33) \nn
586: % \sigma_3/\sigma_1 &=& 2.210(50) \qquad{\rm preliminary!} \la{eq:ratios_c}\\
587: % \sigma_4/\sigma_1 &=& 2.282(63). \nonumber
588: % \ea
589: The $\chi^2$ of the fits are again smaller than 1.
590: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
591: \subsubsection{Final `educated' analysis~\la{sec:final_anal}}
592: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
593: We consider the preceding analysis to be somewhat unsatifactory,
594: because it assumes a specific correction to the $k$-string energies
595: which we are not presently able to confirm directly (see \fig\ref{fig:polyas}),
596: and yet (in the $k=2$ case) is of the same order of magnitude as the
597: difference between two theoretical expectations we are to compare our data to.
598: Moreover we saw that the string tension ratios obtained in this way are systematically
599: higher than if we do not make any assumptions about the L\"uscher coefficients,
600: although the trend is at the one-standard-deviation level.
601:
602: The first analysis is well-principled but suffers from the succession
603: of extrapolations to $L=\infty$ and $a=0$, most of which are based on
604: three data points only and are therefore rather unstable.
605: Considering the large-$L$ extrapolation (\fig\ref{fig:ratios_sig_f},
606: in particular the $k=2$ plot), we see that while the coarsest lattice spacing data
607: still shows a difference with respect to the other two data sets, the latter two
608: essentially fall on a single curve. Therefore we drop the $\beta=115$ data and
609: combine the data at $\beta=138$ and $\beta=172.5$ to do a single extrapolation
610: to $L=\infty$. The result is:
611: \ba
612: \qquad\qquad {\rm lattice~(final)} & {\rm adj. monop.} \quad {\rm fund. monop.} & {\rm trigonometric} \nn
613: \sigma_2/\sigma_1 = 1.707(28) &1.714\qquad \qquad 2.105\qquad&\quad 1.848 \nn
614: \sigma_3/\sigma_1 = 2.182(55) &2.143\qquad \qquad 2.958\qquad&\quad 2.414 \la{eq:final_ratio}\\
615: \sigma_4/\sigma_1 = 2.203(82) &2.286\qquad \qquad 3.256\qquad&\quad 2.613 \nonumber.
616: \ea
617: (the $\chi^2$ are respectively $3.5/5$, $2.35/4$ and $2.6/4$).
618: One ought to associate a systematic error with this final result which is of the
619: same order as the statistical error, since evidence for the absence of scaling violations
620: was given only at that level of accuracy. We also note that the slope, which corresponds
621: to the quantity $\alpha_k$ defined in section~\ref{sec:string_high_rep}, is clearly positive,
622: clearly demonstrating that the central charge of a $k$-string is not $k$ times
623: that of the fundamental string.
624:
625: It is hoped that presenting different analysis strategies has given the reader a sense
626: of the challenge presented by these calculations to reduce the systematic errors on
627: the final string tension ratios.
628: Comparing our data to the theoretical predictions of various models
629: (\eq\ref{eq:final_ratio}) we find that our data
630: is consistent with the Casimir scaling predicted by the adjoint monopole model,
631: and rules out the sine formula by at least 3 standard deviations at all $k$
632: (even if we conservatively assign to the data a systematic error equal to the statistical one).
633: These conclusions agree with earlier results obtained for SU(4) and SU(6)~\cite{lucini},
634: although the accuracy was high enough for $N=4$ to see a (non unexpected)
635: small deviation from Casimir scaling.
636: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
637: \subsection{Excited $k=2$ strings\la{sec:excited}}
638: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
639: \fig\ref{fig:loc_eff_mass} shows the local effective masses, defined as
640: $m_{\rm eff}(t+\frac{a}{2}) \equiv \log\left(\frac{C(t)}{Ct+a)}\right)$,
641: of several of our operators; a plateau is the signature that
642: an energy eigenstate is saturating the correlator.
643: We show the local effective mass for our best $k=1$ operator.
644: The latter has been determined by a variational method~\cite{mart_var}
645: allowing to minimise the contributions from excited states to the correlator.
646: Although several levels of fuzzing were included in the variational basis,
647: the output wave function turned out to be dominated by a single level of fuzzing.
648: We note that its plateau extends out to $t\simeq\sigma^{-1/2}$,
649: giving as confidence in our mass extraction.
650: In the $k=2$ sector, we show local effective masses corresponding to
651: the same level of fuzzing that was optimal for the $k=1$ sector (the inclusion of
652: other fuzzing levels leads to imperceptible changes in the mass plateaux).
653: After the basis operators had been normalised in such a way that $\< O_i(0)O_i^*(t=0) \>=1~(i=1,2)$,
654: the variational procedure selected (within $O(1\%)$ error bars)
655: the anti-symmetric and the symmetric linear combinations of the operators $O_1\equiv\tr\{P^2\}$
656: and $O_2\equiv (\tr P)^2$ for respectively the lightest state and the first excited state.
657: Correspondingly these operators show quite convincing mass plateaux.
658: By comparison, the individual operators have a less good overlap onto the
659: lightest state, although the signal extends far enough in Euclidean time
660: to see that this overlap is not strongly suppressed: their local-effective-masses end up
661: being consistent with the plateau of the anti-symmetric combination. Remarkably
662: their whole correlators seem to agree at all $t$.
663:
664: Thus the theoretical expectation
665: that the energy eigenstates belong to irreducible representations of SU($N$)
666: up to $O(1/N^2)$ admixtures, which was motivated both in the two-state mixing model
667: and by more general arguments about the $N$-dependence of screening
668: (resp. sections~\ref{sec:string_high_rep} and~\ref{sec:open-closed}),
669: is indeed well verified.
670:
671: Another prediction of the two-state mixing model presented in section~\ref{sec:string_high_rep}
672: is that the lightest and the first-excited states should be split symmetrically
673: around the threshold energy of $2m_{k=1}$ (to leading order in $1/N$).
674: This is tested quantitatively in \tab\ref{tab:excited}, which directly compares $2m_{k=1}$
675: to $\frac{1}{2}(m_{k=2}+m^*_{k=2})$. The latter two quantities are remarkably close
676: for all string lengths and lattice spacings, and in many cases they are
677: compatible within the quite small error bars\footnote{As a technical aside,
678: we note that it is essential here to use the correlations between the
679: local effective mass of the lightest and the first-excited $k=2$ states,
680: as they seem to be strongly anti-correlated.}.
681:
682: As we discuss next, the numerical evidence obtained so far
683: favours a binding energy of $k$-strings of order $1/N$.
684: The three predictions that follow straightforwardly from the two-state mixing model
685: presented in section~\ref{sec:string_high_rep} have thus been verified quantitatively.
686: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
687: \subsection{$N$-dependence of the binding energy of $k$-strings}\label{sec:discuss}
688: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
689:
690: On \fig\ref{fig:all_n} (top)
691: we show the relative binding energy of $k$-strings per unit length,
692: $k\sigma_1-\sigma_k$, in units of $\sigma_1$ and rescaled by a factor $N$.
693: We do so by compiling our SU(8) lattice
694: data with the SU(4) and SU(6) data from~\cite{lucini}.
695: The predictions of Casimir scaling and of the Sine formula are also plotted.
696: The figure certainly suggests
697: that the $k=2$ binding energy scales as $1/N$,
698: with a coefficient of order one. By contrast, to account for the measured $N=8,~k=2$
699: binding energy in a $1/N^2$ expansion, the first coefficient would have to be about 20.
700: We further note that the numerical agreement between the Casimir scaling prediction
701: and the lattice data is quite remarkable. If anything, it lies
702: somewhat above the lattice data, indicating that the $k$-strings
703: are slightly less tightly bound that the Casimir formula suggests.
704:
705: On the bottom plot, we show the $k=N/2$ string tension, rescaled by a factor $2/N$,
706: as a function of $1/N$. The data is plausibly heading towards a finite value at
707: $N=\infty$. Here too, Casimir scaling offers a good description of the data.
708: Note that it predicts that the binding energy of the $k=N/2$ string is half
709: of the energy of $k$ non-interacting fundamental strings.
710: The case $k=N/2$ is special in that the relevant operators
711: (listed in appendix C) and their complex conjugate can mix through the
712: appearance of the baryonic vertex on the string. Pictorially it swaps the
713: oriented string from on orientation to the other. Naturally the eigenstates of
714: the Hamiltonian are also eigenstates of the charge conjugation operator, i.e.
715: the real and imaginary parts of the operators, which are respectively $C=+$ and $C=-$.
716: However the existence of a non-vanishing transition probability between the
717: strings of definite orientations means that there is a splitting between the
718: $C=+$ and the $C=-$ states of ${\cal N}$-ality $N/2$ (for $k<N/2$, the
719: center symmetry forces the degeneracy of these two sets of states). In a two-state
720: Hamiltonian formalism, the Casimir formula thus suggests that the
721: Hamiltonian matrix element (per unit length of the string)
722: associated with the baryon-vertex is $\frac{N}{2}\sigma$.
723:
724: It should be noted that the cost of computing the binding energy for a given $k$
725: naively increases as $N^5$ ($N^3$ for the cost of multiplying SU($N$) together in the
726: Monte-Carlo simulation and a $\propto N^2$ increase of the statistics to compensate for
727: the $1/N$ size of the binding energy). And this does not even
728: take into account the condition
729: $\sigma_1 L^2\gg N$ formulated in section~\ref{sec:string_high_rep}.
730: Therefore it could be useful to also
731: compute the string tension ratios for SU(5) and SU(7) before moving to even larger groups.
732:
733:
734:
735: