1: \documentclass[12pt]{article}
2: \usepackage{cite,epsfig}
3:
4: \textheight 9.3in
5: \textwidth 6.5in
6: \topmargin -0.9in
7: \evensidemargin 0.0in
8: \oddsidemargin 0.0in
9:
10: \def\BA{\begin{eqnarray}}
11: \def\BE{\begin{equation}}
12: \def\EA{\end{eqnarray}}
13: \def\EE{\end{equation}}
14:
15: \def\Address#1#2{$^{\rm#1}${\it\footnotesize#2}\\}
16: \def\Ref#1{(\ref{#1})}
17: \def\cf{{\it cf. }}
18: \def\eps{\varepsilon}
19: \def\Dtau{\Delta\tau}
20: \def\absP{\langle\vert P\vert\rangle}
21: \def\gtsim{\lower-0.45ex\hbox{$>$}\kern-0.77em\lower0.55ex\hbox{$\sim$}}
22: \def\ltsim{\lower-0.45ex\hbox{$<$}\kern-0.77em\lower0.55ex\hbox{$\sim$}}
23:
24: \begin{document}
25:
26: % ------ Title ---------------------------------------------
27: \title{\bf Transverse QCD Dynamics \\ Near the Light Cone}
28: \author{
29: E.--M.~Ilgenfritz$^{\rm a,c}$,
30: Yu.P.~Ivanov$^{\rm a,b,d}$,
31: H.J.~Pirner$^{\rm a,b}$\footnote{
32: Supported by the European Contract No. FMRX-CT96-0008.
33: }
34: \\
35: \\
36: \Address{a}{
37: Institut f\"ur Theoretische Physik der Universit\"at,
38: Philosophenweg 19, D-69120 Heidelberg, Germany
39: }
40: \Address{b}{
41: Max-Planck Institut f\"ur Kernphysik,
42: Postfach 103980, D-69029 Heidelberg, Germany
43: }
44: \Address{c}{
45: Research Center for Nuclear Physics,
46: Osaka University, Osaka 567-0047, Japan
47: }
48: \Address{d}{
49: Joint Institute for Nuclear Research,
50: Dubna, 141980 Moscow Region, Russia
51: }
52: }
53: \date{\today}
54: \maketitle
55:
56: % ---- Abstract --------------------------------------------
57: \begin{abstract}
58: Starting from the QCD Hamiltonian in near-light cone coordinates,
59: we study the dynamics of the gluonic zero modes. Euclidean 2+1
60: dimensional lattice simulations show that the gap at strong
61: coupling vanishes at intermediate coupling. This result opens
62: the possibility to synchronize the continuum limit with the
63: approach to the light cone.
64: \end{abstract}
65:
66: % ------ Introduction --------------------------------------
67: \section{Introduction}
68:
69: The solution of Quantum Chromodynamics {\em on} the light cone
70: is still an unsolved theoretical task the present status of which
71: is reviewed in ref. \cite{Brodsky}. In a recent paper \cite{NPFV}
72: a formulation of QCD in coordinates {\em near} the light cone
73: has been proposed which has the advantage of keeping a direct
74: link \cite{PRFR,LTLY,VFP} to equal time theories. The Cauchy
75: problem is well defined in near light cone coordinates, since
76: the initial data are given on a space like surface. This formulation
77: avoids the solution of constraint equations which, on the quantum
78: field theoretic level, may be very complicated. The problem of
79: a nontrivial vacuum appears in a solvable form related to the
80: transverse dynamics. This is physically very appealing, since
81: in high energy reactions the incoming particles propagate near the
82: light cone and interact mainly exchanging particles with transverse
83: momenta. We would like to connect successful models for the soft
84: transverse nonperturbative dynamics of high energy reactions
85: (\cf refs. \cite{DGKP,N}) to the underlying QCD Hamiltonian.
86: In diffractive reactions, fast Lorentz contracted hadrons
87: experience the confining forces of QCD in their transverse
88: extensions and interact with other fast moving hadrons via
89: soft interactions. In a theory of total cross sections the
90: nonperturbative infrared dynamics in the transverse plane is
91: essential. The objective of this paper is to investigate the
92: effective transverse Hamiltonian in QCD near the light cone
93: for small light cone momenta.
94:
95: Near light cone QCD has a nontrivial vacuum which we claim
96: cannot be neglected even in the light cone limit. We will
97: demonstrate the existence of massless excitations in the
98: zero mode theory. These excitations do not decouple in the
99: light cone limit. Genuine nonperturbative techniques must be
100: used to investigate the behavior of this limit. In principle
101: the additional parameter which labels the coordinate system
102: can be chosen arbitrarily. We will show that the zero mode
103: Hamiltonian depends on an effective coupling constant containing
104: this parameter and evolves towards an infrared fixed point.
105: Therefore, we propose a sophisticated choice of the frame
106: dependence which facilitates calculations in the near light
107: cone frame. We use the infrared fixed point of the zero mode
108: Hamiltonian to follow a trajectory in the space of couplings,
109: where high resolution is synchronized with the light cone limit.
110: Note, we consider the parameter associated with the frame
111: dependence as yet another coupling constant. As shown in
112: Ref. \cite{Pol} the zero mode sector is also relevant
113: to the definition of M-theory in light cone coordinates.
114:
115: We choose the following near light cone coordinates which
116: smoothly interpolate between the Lorentz and light front
117: coordinates:
118: \BA
119: x^t ~=~ x^{+} &=& \frac1{\sqrt2}
120: \left\{ \left(1 + \frac{\eta^2}{2} \right)
121: x^{0} + \left(1 - \frac{\eta^2}{2} \right) x^{3}
122: \right\} ~,\nonumber \\
123: x^{-} &=& \frac1{\sqrt2} \left( x^{0}-x^{3} \right)~.
124: \label{Coor}
125: \EA
126: The transverse coordinates $x^1$, $x^2$ are unchanged;
127: $x^t = x^{+}$ is the new time coordinate, $x^{-}$ is a
128: spatial coordinate. As finite quantization volume we will
129: take a torus and its extension in ``-'', as well as in ``1, 2''
130: direction is $L$. The scalar product of two 4-vectors $x$ and $y$
131: is given with $\vec{x}_\bot \vec{y}_\bot = x^1 y^1 + x^2 y^2$ as
132: \BA
133: x_\mu y^\mu & = & x^- y^+ + x^+ y^- - \eta^2 x^- y^-
134: - \vec{x}_\bot \vec{y}_\bot \nonumber \\
135: & = & x_- y_+ + x_+ y_- + \eta^2 x_+ y_+
136: - \vec{x}_\bot \vec{y}_\bot ~.
137: \label{scalpr}
138: \EA
139: Obviously, the light-cone is approached as the parameter $\eta$
140: goes to zero. For non-zero $\eta$, the transition to the coordinates
141: introduced above can be formally identified as a Lorentz-boost
142: combined with a linear transformation, which avoids time dependent
143: boundary conditions \cite{LTLY}. The boost parameter $\beta = v_3$
144: is given by
145: \BE
146: \beta = \frac{1 - \eta^2/2}{1+ \eta^2/2} ~,
147: \label{beta}
148: \EE
149: indicating that for $\eta^2 \rightarrow 0$ the relative velocity
150: $v_3 \rightarrow c\, (\equiv 1)$. Of course, this is connected to the
151: well-known interpretation of the 'tilted' light-cone frame in terms
152: of the infinite momentum frame.
153:
154: The choice of near light cone coordinates allows to quantize the
155: theory on a space-like finite interval of length $L$ in $x^-$ at
156: equal times, {\it i.e.} $\Delta x^+ = 0$. The invariant length
157: squared of this interval for $\Delta x_{\bot}^2=0$ is related
158: to the length of the compact $x^-$ dimension.
159: \BA
160: \Delta s^2 & = & \Delta x^- \Delta x^+ + \Delta x^+ \Delta x^-
161: - \eta ^2 (\Delta x^-)^2 - \Delta x_\bot^2 \nonumber \\
162: & = & -\eta^2 L^2 ~.
163: \EA
164: For simplicity we consider also the transverse dimensions periodic
165: in $L$. Previous work of the St.~Petersburg and Erlangen groups
166: \cite{PRFR,LTLY} assumed a fixed tilted coordinate system with
167: fixed transverse ultraviolet cut off. Our main purpose in this
168: paper is to consider the zero mode fields on a transverse lattice
169: with {\it varying} transverse lattice spacing $a$. We propose to
170: approach light-cone dynamics by synchronizing the continuum limit
171: $\Lambda=\pi/a\to\infty$ with the light-cone limit $x^+\to\frac{1}%
172: {\sqrt 2}(x^0+x^3)$. The vacuum fluctuations in the transverse
173: directions induce a second order phase transition which allows
174: to have a continuum limit of the lattice theory. Thereby we can
175: eliminate the cut off in a controlled way, preserving the
176: nontrivial vacuum structure.
177:
178: The gauge fixing procedure in the modified light-cone gauge
179: $\partial_-A_- = 0$ involves zero modes dependent on the transverse
180: coordinates. These zero mode fields carry zero linear momentum $p_-$
181: in near light cone coordinates, but finite amount of $p_0+p_3$.
182: They correspond to "wee" partons in the language of the original
183: parton model of Feynman. In $SU(2)$ the zero mode fields $a_-(x_\perp)$
184: are proportional to $\tau^3$, i.e. they can be chosen color diagonal.
185: The use of an axial gauge is very natural for the light-cone
186: Hamiltonian even more so than in the equal-time Hamiltonian.
187: The asymmetry of the background zero mode naturally coincides
188: with the asymmetry of the space coordinates on the light cone.
189: The zero mode fields describe disorder fields. Depending on the
190: effective coupling the zero mode transverse system will be in the
191: massive or massless phase. Its second order phase transition allows
192: to perform the continuum limit in the zero mode Hamiltonian. Our
193: main conjecture is that the continuum limit and light cone limit
194: can be realized simultaneously at this critical point. The evolution
195: of the coupling determines the approach to the light cone with
196: transverse resolution approaching zero. The resulting relation
197: is reminiscent of the simple behavior one gets from considering
198: naive scaling relations and dynamics in the infinite momentum
199: frame. We will also calculate the contribution of the two dimensional
200: zero modes to the ground state energy and show that it scales with the
201: three dimensional volume. This estimate demonstrates very simply the
202: relevance of modes with lower dimensionality to the full problem in
203: the case of critical second order behavior.
204:
205: The naive considerations for a simultaneous light cone and continuum
206: limit go as follows. We characterize the 'infinite' momentum frame
207: by giving the momenta of the proton and photon in usual coordinates.
208: The fast moving proton carries $(P,0_{\bot},P)$ and the photon
209: $q= (\frac{\nu M}{2 P},\sqrt{Q^2},-\frac{\nu M}{2 P})$, where
210: $\nu$ is the energy transfer in the laboratory. The dominant
211: contribution comes from energy conserving transitions, where
212: the energy of the quark in the final state equals the sum of
213: initial quark and photon energies. For collinear quarks with
214: momentum $p=(x_B P,0_{\bot},x_B P)$, where $x_B$ denotes the
215: momentum fraction, one explicitly finds
216: \BE
217: \sqrt{Q^2+\left(x_B P-\frac{\nu M}{2 P}\right)^2}=x_B P+\frac{\nu M}{2P},
218: \EE
219: which yields the ``scaling variable'' $x_B=\frac{Q^2}{2 M \nu}$
220: as long as $P\geq Q/x_B$, i.e. the ``infinite'' momentum has to be
221: large enough:
222: \BE
223: \gamma M = P \geq Q = 1/a.
224: \EE
225: At the same time, the photon has a wavelength $\propto 1/Q$ which
226: resolves transverse details of size $a$ in the hadronic wavefunction.
227: By eq. \Ref{beta}, the $\gamma$ factor of the ``infinite'' momentum
228: is related to $\eta$. Therefore we expect that the parameter $\eta$
229: approaches zero with the transverse cutoff $\Lambda=\pi/a\rightarrow\infty$
230: \BE
231: \eta=\frac1{\sqrt{2}\gamma} \approx \frac{1}{\sqrt{2}}\,aM \rightarrow 0.
232: \EE
233: One of the objectives of this paper is to derive the precise relation
234: between $\eta$ and $a$. We will simulate the zero mode dynamics
235: on a $(2+1)$ dimensional lattice to find the fixed point of the
236: effective coupling. The light-like limit is governed by an infrared
237: fixed point (\cf ref. \cite{Pol}). Most previous discussions of the
238: zero mode problem have been on the classical equations of motion level,
239: here we will work on the fully nonperturbative quantum level.
240: Our work is not directly related to M-theory, but some of the
241: consequences may be relevant for other studies with light like
242: compactification.
243:
244: % ------ Near Light Cone -----------------------------------
245: \section{Near Light Cone QCD Hamiltonian}
246:
247: In Ref. \cite{NPFV} the near light cone Hamiltonian has been derived.
248: A similar Hamiltonian has been obtained in equal time coordinates
249: \cite{LNOT,LNT}. Here we will sketch the derivation. We restrict
250: ourselves to the color gauge group $SU(2)$ and dynamical gluons;
251: only an external (fermionic) charge density $\rho_m$ is
252: considered here.
253:
254: Since the $A^a_+$ coordinates have no momenta conjugate to them,
255: the Weyl gauge $A^a_+=0$ is the starting point for a canonical
256: formulation. The canonical momenta of the dynamical fields
257: $A_-^a, A_i^a$ are given by
258: \BA
259: \Pi^a_- &=& \frac{\partial{\cal L}}{\partial F^a_{+-}} = F^a_{+-},
260: \nonumber \\
261: \Pi^a_i &=& \frac{\partial{\cal L}}{\partial F^a_{+i}} = F^a_{-i}
262: +\eta^2 F^a_{+i}.
263: \EA
264: >From this, we get the Weyl gauge Hamiltonian density
265: \BE
266: \label{Hweyl}
267: {\cal H}_W = \frac{1}{2} \Pi^a_- \Pi^a_-
268: + \frac{1}{2}F^a_{12}F^a_{12}
269: + \frac{1}{2\eta^2}\sum_{i=1,2}\left(\Pi^a_i-F^a_{-i}\right)^2 ~.
270: \EE
271: The Hamiltonian has to be supplemented by the original Euler--Lagrange
272: equation for $A_+$ as constraints on the physical states (Gauss' Law
273: constraints)
274: \BA
275: \label{Gauss}
276: G^a({x}_{\bot},x^-) |\Phi\rangle
277: &=& \left( D_-^{ab} \Pi^b_{-}
278: + D_{\bot}^{ab}\Pi^b_{\bot}
279: + g \rho_m^a
280: \right) | \Phi \rangle \nonumber \\
281: &=& \left(D_-^{ab}\Pi^b_{-} + G_{\bot}^{a}\right) |\Phi\rangle = 0~.
282: \EA
283:
284: In order to obtain a Hamiltonian formulated in terms of unconstrained
285: variables, one has to resolve the Gauss' Law constraint. Via unitary
286: gauge fixing transformations [9,10] a solution of Gauss' Law with
287: respect to components of the chromo--electric field $\Pi_{-}$ can be
288: accomplished. This gives a Hamiltonian independent of the conjugate
289: gauge fields $A_{-}$, {\it i.e.} the latter become cyclic variables.
290: Classically this would correspond to the light front gauge $A_{-}=0$.
291: However, this choice is not legitimate if we want to consider the
292: theory in a finite box. Instead, the (classical) Coulomb light front
293: gauge $\partial_{-}A_{-}=0$ is compatible with gauge invariance and
294: periodic boundary conditions. The reason is that $A_{-}$ carries
295: information on the (gauge invariant) eigenvalues of the spatial
296: Polyakov line matrix
297: \BE
298: \label{Polyakov}
299: \hat{\cal P}(x_\bot)=P\exp\left[ig\int dx^-A_-(x_\bot,x^-)\right] ~,
300: \EE
301: which can be written in terms of a diagonal matrix with the zero mode
302: field $a^3_-(x_\bot) \frac{\tau_3}{2} = a_-(x_\bot)$
303: \BE
304: \hat{\cal P}(x_\bot) =
305: V \exp\left[ig L a_-(x_\bot)\right] V^{\dagger} ~.
306: \EE
307: Obviously we have to keep these `zero modes' $a_-(x_\bot)$ as
308: dynamical variables, while the other components of $A_{-}$ are
309: eliminated. The zero mode degrees of freedom are independent of
310: $x^{-}$ and, therefore, correspond to quantities with zero
311: longitudinal momentum $p_{-}$.
312:
313: In order to eliminate the momentum $\Pi_-$, conjugate to $A_-$, by
314: means of Gauss' Law, one needs to `invert' the covariant derivative
315: $D_-$ which simplifies to $d_{-}=\partial_{-} - ig[a_{-},..]$. On
316: the space of physical states one can simply make the replacement%
317: \footnote{The inversion of $d_-$ can be explicitly performed in
318: terms of its eigenfunctions, \cf \Ref{Gauss}.}
319: \BE
320: \Pi_-(x_\bot,x_-) \rightarrow
321: p_-(x_\bot)-\left(d_-^{-1}\right) G_\bot(x_\bot,y^-) ~.
322: \EE
323: The operator $p_-({x}_{\bot})$ is also diagonal and $p^3_-({x}_{\bot})$
324: is the momentum conjugate to the zero mode $a^3_{- }(x_{\bot})$. It has
325: eigenvalue zero with respect to $d_{-}$, {\it i.e.} $d_- p_- = 0$, and
326: is therefore not constrained.
327:
328: The appearance of the zero modes implies a residual Gauss' Law, which
329: arises from the $x_{-}$ integration over eq. \Ref{Gauss} for the $a=3$
330: component using periodic boundary conditions in the $x_{-}$ direction.
331: This constraint on two dimensional fields can be handled in full analogy
332: to QED, since it concerns the diagonal part of color space. A further
333: Coulomb gauge fixing in the $SU(2)$ 3--direction eliminates the color
334: neutral, $x^-$--independent, two--dimensional longitudinal gauge field
335: in favor of a neutral chromo--electric field
336: \BE
337: e_\bot(x_\bot) = g \nabla_\bot \int dy^- dy_\bot
338: d(x_\bot - y_\bot)
339: \left\{ f^{3ab} A^a_\bot(y_{\bot}, y^-) \Pi^b_{\bot}(y_{\bot}, y^-)
340: + \rho^3_m(y_{\bot}, y^-) \right\} \frac{\tau^3}{2} ~.
341: \EE
342: Here we use the periodic Greens function of the two dimensional
343: Laplace operator
344: \BE
345: \label{Green2}
346: d(z_\bot) = - \frac1{L^2} \sum_{\vec{n} \neq \vec0 }
347: \frac1{p_n^2} e^{i p_n z_\bot} ~,
348: ~~~~~ p_n = \frac{2\pi}{L}\vec{n} ~,
349: \EE
350: where $\vec{n} = (n_1,n_2)$ and $n_1,n_2$ are integers.
351:
352: As a remnant of the local Gauss' Law constraints, a global condition
353: \BE
354: \label{Gaussgl}
355: Q^3 | \Phi' \rangle =
356: \int dy^- dy_\bot
357: \left\{ f^{3ab} A^a_\bot(y_\bot,y^-) \Pi^b_\bot(y_\bot, y^-)
358: + \rho^3_m(y_{\bot}, y^-) \right \}
359: | \Phi' \rangle = 0
360: \EE
361: emerges. The physical meaning of this equation is that the neutral
362: component of the total color charge, including external matter as
363: well as gluonic contributions, must vanish in the sector of physical
364: states. It can be shown (see \cite{NPFV}) that there must be
365: $\tilde{Q}_{12}(x_{\bot})=0$ everywhere in transverse space in
366: order to avoid an infinite Coulomb energy. The two conditions
367: together suggest that physical states have to be color singlets.
368:
369: The final Hamiltonian density in the physical sector explicitly reads
370: (in terms of unconstrained $A_{\bot}$ and $\Pi_{\bot}$ obtained after
371: a shift is made by subtracting averages) \cite {NPFV}
372: \BA
373: \label{Hamil}
374: {\cal H} &=& \mbox{tr}
375: \left[\partial_1 A_2 - \partial_2 A_1 - ig [A_1, A_2] \right]^2
376: + \frac1{\eta^2} \mbox{tr} \left[\Pi_\bot - \left(\partial_-A_\bot
377: - ig [a_-,A_\bot] \right) \right]^2 \nonumber \\
378: &+& \frac1{\eta^2} \mbox{tr}
379: \left[ \frac1L e_\bot - \nabla_\bot a_- \right]^2
380: + \frac1{2 L^2} p_-^{3\,\dagger}(x_\bot) p^3_-(x_\bot) \nonumber \\
381: &+& \frac1{L^2} \int_0^L dz^- \int_0^L dy^-
382: \sum_{p,q,n}\,^{'}
383: \frac{G_{\bot qp}(x_\bot,z^-)
384: G_{\bot pq}(x_\bot,y^-)}{\left[ \frac{2\pi n}{L}
385: + g(a_{-q}(x_\bot) - a_{-p}(x_\bot)) \right]^2}
386: e^{i 2\pi n(z^- - y^-)/L} ~,
387: \EA
388: where $p$ and $q$ are matrix labels for rows and columns, $a_{-q} = %
389: (a_-)_{qq}$ and the prime indicates that the summation is restricted
390: to $n \neq 0$ if $p = q$. The operator $G_\perp\left(x_\bot,x^-\right)$
391: is defined as
392: \BE
393: G_{\perp}=
394: \label{Gop}
395: \nabla_\perp \Pi_\perp
396: + gf^{abc} \frac{\tau^a}{2} A^b_\perp \left( \Pi^c_\perp
397: - \frac1L e^c_\bot\right)
398: + g\rho_{m} ~.
399: \EE
400:
401: The last two terms of the Weyl gauge Hamiltonian come from
402: the original term $\Pi_-^2$ with squared electric field
403: strengths in $x^-$ direction. After elimination of $\Pi_-$
404: the zero mode part and the light cone Coulomb energy in the
405: axial gauge remain. In the Coulomb term one sees the role of
406: the zero mode fields as infrared regulators of the spatial
407: momenta $p_-=2 \pi n/ L$ which are quantized due to the compact
408: interval $L$. Since the two dimensional theory has been Coulomb
409: gauge fixed, the electric field $\vec{e}_{\bot}$ replaces the
410: canonical momentum of the longitudinal, neutral gauge field
411: which has been eliminated.
412:
413: We note that the terms containing $\vec{e}_{\bot}$ and the
414: momentum $\vec{\Pi}_{\bot}$, have the pre-factor $1/\eta^2$.
415: Physically this pre-factor signals the increase of transverse
416: electric energies with the boost factor $\gamma = (\sqrt2\eta)^{-1}$.
417: The boost also couples transverse electric fields with transverse
418: magnetic fields. In the light-cone limit the pre-factor diverges
419: and the adjacent brackets become constraint equations. This
420: reflects the corresponding reduction of the number of degrees
421: of freedom if one goes exactly on the light cone. We do not
422: follow this procedure, but keep a finite $\eta$ as a kind of
423: Lagrange parameter.
424:
425: A characteristic feature of an exact light-cone formulation is
426: the triviality of the ground state. This may simplify explicit
427: calculations, {\it e.g.} of the hadron spectrum. However, the light
428: cone vacuum is definitely not trivial in the zero mode sector.
429: In fact, already in ref \cite{NPFV} we have shown strong and
430: weak coupling solutions of the zero mode Hamiltonian. Massless
431: modes influence the dynamics in the light cone limit, whereas
432: massive modes decouple when $\eta \to 0$. In contrast to earlier
433: work, we solve the zero mode Hamiltonian in this paper numerically
434: showing the transition from the massive phase to the massless phase.
435:
436: The zero mode degrees of freedom $a_-$ couple to the transverse
437: three dimensional gluon fields $A_i$ via the second magnetic term
438: in $\cal H$, the Coulomb term and directly via the electric field
439: $\vec{e}_{\bot}$. We remark that the transverse electric fields
440: $\Pi_\bot$ and $\vec{e}_\bot$ are dual to the magnetic fields
441: $\partial_-A_\bot$ and $\nabla_\bot a_-$. This duality is typical
442: for the light cone and is absent in the equal time case. Since
443: duality plays an important role in supersymmetric QCD its role
444: in light cone theories should deserves to be investigated in
445: greater detail.
446:
447: In the following we neglect the couplings between the three and
448: two dimensional fields and consider the pure zero mode Hamiltonian
449: \BE
450: \label{Hamil3}
451: h = \int d^2\!x \; \left[
452: \frac1{2 L} p_-^{3\dagger}(\vec{x}_\bot) p_-^3(\vec{x}_\bot)
453: + \frac{L}{2\eta^2} (\nabla_\bot a_-^3)^2 \right] ~.
454: \EE
455: The global constraint, eq. \Ref{Gaussgl}, does not contain $a_-$
456: and $p_-$ and, consequently, is irrelevant for the time being.
457: Even at this level of severe approximations the zero mode
458: Hamiltonian differs from the corresponding one in QED. The
459: reason is the hermiticity defect of the canonical momentum
460: $p^-$. In the Schr\"odinger representation eq. \Ref{Hamil3}
461: \BE
462: \label{Hamil4}
463: h = \int d^2x \; \left[
464: - \frac{1}{2L}\frac{1}{J(a_-^3(\vec{x}_\bot))}
465: \frac\partial{\partial a_-^3(\vec{x}_\bot)}
466: J\left(a_-^3(\vec{x}_\bot)\right)
467: \frac\partial{\partial a_-^3(\vec{x}_{\bot})}
468: + \frac{L^2}{2\eta^2} (\nabla_\bot a_-^3)^2 \right] ~,
469: \EE
470: contains the Jacobian $J(a_-)$ which equals the Haar measure of $SU(2)$
471: \BE
472: \label{Jacob}
473: J\left(a_-^3(\vec{x}_{\bot})\right) =
474: \sin^2 \left( \frac{gL}{2}a_-^3(\vec{x}_{\bot})\right) ~.
475: \EE
476: It stems from the gauge fixing procedure, effectively introducing
477: curvilinear coordinates. It also appears in the functional
478: integration volume element for calculating matrix elements.
479: It is convenient to introduce dimensionless variables
480: \BE
481: \varphi(\vec{x}_\bot) = \frac{gL a_-^3(\vec{x}_\bot)}{2} ~,
482: \EE
483: which vary in a compact domain $0 \leq \varphi \leq \pi$. We
484: regularize the above Hamiltonian $h$ by introducing a lattice
485: spacing $a$ between transversal lattice points $\vec{b}$.
486: Next we appeal to the physics of the infinite momentum frame and
487: factorize the reduced true energy from the Lorentz boost factor
488: $\gamma=\sqrt2/\eta$ and the cut off by defining $h_{\rm red}$
489: \BE
490: \label{hdef}
491: h = \frac1{2\eta a} h_{\rm red}
492: \EE
493: In the continuum limit of the transverse lattice theory we let
494: $a$ go to zero. For small lattice spacing we obtain the reduced
495: Hamiltonian
496: \BE
497: h_{\rm red} = \sum_{\vec{b}}\left\{ -g^2_{\rm eff}
498: \frac1J \frac\partial {\partial \varphi(\vec b)}
499: J \frac\partial {\partial \varphi(\vec b)}
500: + \frac1{g^2_{\rm eff}} \sum_{i=1,2}
501: \left(\varphi(\vec{b})-\varphi(\vec{b}+\vec{e}_i)\right)^2
502: \right\} ~.
503: \EE
504: with the effective coupling constant
505: \vskip-5mm
506: \BE
507: \label{g2eff}
508: g^2_{\rm eff} = \frac{g^2L\eta}{4a}
509: \EE
510: The first part of the Hamiltonian contains the kinetic (electric)
511: energy of the $SU(2)$ rotators on a half circle at each lattice
512: point and the second part gives the potential (magnetic) energy
513: of these rotators due to the differences of angles at nearest
514: neighbor sites. For further discussion we define these electric
515: and magnetic parts as:
516: \BA
517: h_{\rm red} & = & h_{\rm e} + h_{\rm m} \\
518: h_{\rm e} & = & \sum_{\vec{b}} h^{\rm e}_{\vec{b}} \nonumber \\
519: h_{\rm m} & = & \sum_{\vec{b}}\sum_{i=1,2}
520: h^{\rm m}_{\vec{b},\vec{b}+\vec{e}_i} \nonumber
521: \EA
522: In the strong coupling domain where we have analytical solutions
523: of the zero mode Hamiltonian, the numerical Hamiltonian lattice
524: theory agrees with the analytical solutions for the mass gap.
525: The real question concerns the continuum limit of the zero mode
526: system which occurs outside of the strong coupling region. For
527: this purpose we have to find the region of vanishing mass gap
528: for the lattice Hamiltonian.
529:
530: % ------ Lattice -------------------------------------------
531: \section{Lattice Calculation of the Zero Mode Hamiltonian}
532:
533: We solve the equivalent lattice theory in a Euclidean formulation.
534: The zero mode Hamiltonian represents a $2+1$ dimensional theory in
535: two spatial and one time direction. The lattice has $N_x a = N_y a$
536: extensions in transverse space and $N_T \Dtau = T$ extension in near
537: light cone time. To set up the density matrix one has to write down
538: the Trotter formula for the given Hamiltonian. Using $h_{\rm red} =%
539: h_{\rm e}+h_{\rm m}$ we have
540: \BE
541: \exp(-T h_{\rm red}) = \lim_{N_T\to\infty}
542: \left[ \exp\left(-\Dtau h_{\rm m}/2 \right)
543: \exp\left(-\Dtau h_{\rm e} \right)
544: \exp\left(-\Dtau h_{\rm m}/2 \right)
545: \right]^{N_T},
546: \EE
547: where each time evolution step $\Dtau$ can be separately done for
548: the electric and magnetic part of the Hamiltonian. For definiteness,
549: we choose $\Dtau=a/2$ in all the following. In the Appendix we show
550: that this choice of $\Dtau$ optimizes the updating procedure since
551: it generates approximately equal widths for the weight functions
552: resulting from the kinetic and potential energies. The different time
553: slices will labeled with the index $l$. The electric Hamiltonian can
554: be evaluated by inserting products of complete set single site
555: eigenfunctions $C_{n_l}$ (\cf \cite{NPFV}). Practically a maximal
556: number $N_{\rm max} = 100$ of eigenfunctions is fully sufficient
557: to reach convergence in the interval of couplings we need.
558: \BE
559: \label{sum_over_states}
560: \langle \varphi_{l+1} | h^{\rm e} | \varphi_l \rangle =
561: \!\!\sum_{n_{l+1},n_{l}}
562: \langle \varphi_{l+1} | n_{l+1} \rangle
563: \langle n_{l+1} | h^{\rm e} | n_l \rangle
564: \langle n_l | \varphi_l \rangle \approx
565: \!\!\sum_{n_{l}=0}^{N_{\rm max}}
566: C_{n_l}(\varphi_{l+1}) g_{\rm eff}^2~n_l(n_l+2)
567: C_{n_l}(\varphi_l)
568: \EE
569: with the (single site) eigenfunctions and eigenvalues given as:
570: \BA
571: \label{Basis}
572: C_{n_l}(\varphi_l) &=&
573: \sqrt{\frac2\pi} \left\{
574: \frac{\sin\left((n_l+1)\varphi_l\right)}{\sin\varphi_l}
575: \right\} ~, \\
576: h^{\rm e}~C_{n_l}(\varphi_l) &=&
577: g^2_{\rm eff}~n_l (n_l+2)~C_{n_l}(\varphi_l) ~.
578: \EA
579: The eigenfunctions form an orthonormal set with respect to a scalar
580: product which contains the Jacobian in the measure. The Jacobian
581: $J\left(\varphi(\vec b)\right)$ has been defined above \Ref{Jacob}.
582: The magnetic part is diagonal in $\{\varphi(\vec{b})\}$
583: {\it i.e.} local in time:
584: \BE
585: \langle \{\varphi_l(\vec{b})\}|h_{\rm m}|\{\varphi_l(\vec{b})\}\rangle
586: = h_{\rm m}(\{\varphi_l(\vec{b})\})
587: = \sum_{\vec{b}}\sum_{i=1,2}
588: h^{\rm m}_{\vec{b},\vec{b}+\vec{e}_i}
589: \left(\varphi_l(\vec{b})-\varphi_l(\vec{b}+\!\vec{e}_i\,)\right) ~.
590: \EE
591: The full partition function is given by an integral over all time slices
592: \BA
593: \!\!\!Z &=&
594: \mbox{tr}\,\exp\!\left[-T h_{\rm red}\right]
595: = \int \!\prod_{\vec{b}}
596: \left(J(\varphi(\vec{b})) d\varphi(\vec{b})\right) \,
597: \langle \{\varphi(\vec{b})\} | \exp\!\left[-T h_{\rm red}\right]
598: | \{\varphi(\vec{b})\}
599: \rangle \\
600: &=&
601: \int \!\prod_{\vec{b},l}
602: \left(J(\varphi(\vec{b})) d\varphi(\vec{b}) \,
603: \sum_{n_l=0}^{N_{\rm max}}
604: C_{n_l}\!(\varphi_{l+1}(\vec{b}))
605: \exp\left[ -g^2_{\rm eff} n_l (n_l\!+\!2) \Dtau \right]
606: C_{n_l}\!(\varphi_{l}(\vec{b})) \right)
607: \nonumber \\
608: &\times&
609: \prod_{l} \exp\!\left[ -h_{\rm m} (\{\varphi_l(\vec{b})\}) \Dtau \right] ~.
610: \nonumber
611: \EA
612: Because of the Jacobian the dominant contributions to the partition
613: function come from $\varphi_l$ values around $\frac{\pi}{2}$. The
614: Hamiltonian is invariant under reflections $\varphi_l-\frac{\pi}{2}
615: \rightarrow \frac{\pi}{2}-\varphi_l$. In the strong coupling limit,
616: where $g^2_{\rm eff}$ is large, the half rotators act almost independently
617: on each lattice site producing a large mass gap. The system is disordered.
618: For decreasing coupling constant $g^2_{\rm eff}$ the movement of the
619: individual rotators becomes locked from one site to the next. Long
620: range correlations develop. The order parameter (or ``magnetization'')
621: of the system is the expectation value of the trace of the Polyakov
622: line \Ref{Polyakov}, which on the lattice has the form
623: \BE
624: P = \overline{\frac{1}{2} tr \hat{\cal P}}
625: = \frac1{N^3}\sum\limits_{\vec b,\tau} \cos\varphi(\vec b,\tau) ~.
626: \label{OrderPar}
627: \EE
628:
629: The order parameter is odd under the above symmetry operation.
630: Operator expectation values are evaluated with the density matrix
631: defined by $h_{\rm red}$ and $T$:
632: \BE
633: \langle O \rangle =
634: \frac1Z\,\mbox{tr}\left\langle O\exp[-T h_{\rm red}]\right\rangle~.
635: \EE
636:
637: At each particular $\beta_g$
638: \BE
639: \beta_g \equiv 1/g_{\rm eff}^2
640: \EE
641: we simulate lattice sizes with equal number of sites in all directions
642: $N = N_x = N_y = N_T$ and $N = 4,6,8,12,16$. A Metropolis algorithm is
643: used for updating with the steps size and the number of hits adapted
644: to $\beta_g$. In order to tabulate the Boltzmann weights related to
645: the timelike links we also discretized the continuous angle variables
646: to a system of $N_\varphi=60$ orientations: $\varphi_j=j\pi/N_\varphi$,
647: $j=[0,N_\varphi]$.
648:
649: In this explorative investigation we generated between $5000$ and
650: $50000$ uncorrelated configurations depending on the $\beta_g$ and
651: lattice size. All our calculations were done on a cluster of
652: AlphaStations with the single site Metropolis updating algorithm
653: sketched above, so the accuracy can be improved using a more
654: powerful algorithms and/or more computing resources.
655:
656: We calculate the $\beta_g$ dependence of the following quantities:
657: average electric energy $\eps_{\rm e}$ and magnetic energy $\eps_{\rm m}$
658: per site, the average of the absolute value of the order parameter
659: $\absP$, the susceptibility $\chi$ and the normalized fourth cumulant
660: $g_r$ (which gives the deviation of the moments of the Polyakov
661: expectation value from a pure Gaussian behavior):
662: \vskip-7mm
663: \BA
664: \eps_{\rm e,m} &=& \frac1{N^2} \langle h_{\rm e,m} \rangle \\
665: \chi &=& N^3 \left(\langle P^2 \rangle
666: - \langle P \rangle^2\right) \label{chidef} \\
667: g_r &=& \frac{\langle P^4 \rangle}{\langle P^2 \rangle^2} - 3.
668: \EA
669:
670: Firstly we calculated the ground state energy for strong couplings
671: $g_{\rm eff}^2$ and compared with the exact calculation \cite{NPFV}.
672: The obtained results from the lattice agree with the analytical result.
673: For higher values of $\beta_g$ the lattice calculation agrees within
674: $10-20~\%$ with the previous effective double site calculation for
675: the energy per site\cite{NPFV}. The ultraviolet regularization
676: with $\Dtau$ influences the final lattice result. The energies are
677: measured with high accuracy and practically do not depend on the
678: infrared cutoff, i.e. on the total lattice size for $N \geq 8$.
679: The situation for other variables is different.
680:
681: In order to see the emergence of a massless phase, we investigated
682: the (connected) time correlation functions of the (trace of the)
683: Polyakov line operators:
684: \BA
685: K(0,\tau) = \frac1{N^4} \left\langle
686: \sum_{\vec{b},\vec{b'}}
687: \cos(\varphi(\vec b,0))
688: \cos(\varphi(\vec b',\tau))
689: \right\rangle - \langle |P| \rangle^2 ~.
690: \EA
691: The correlation masses are obtained from a fit of the time correlation
692: functions $K(0,\tau)\vert_{\tau=n\Dtau}$ to a parameterization taking
693: the periodicity in time into account
694: \BE
695: \label{mcorr}
696: K(0,n\Dtau) = c\,
697: \mbox{cosh}\!\left[m\Dtau\left(n-\frac{N_T}2\right)\right] ~.
698: \EE
699:
700: % ---- Fig. Masses
701: \begin{figure}[ht]
702: \centerline{
703: \scalebox{0.47}{\includegraphics{zm-mn.ps}}~~
704: \scalebox{0.47}{\includegraphics{zm-m0.ps}}
705: }
706: \caption{
707: \label{FigMasses}
708: Correlation masses $m$ for different effective couplings
709: $\beta_g = 1/g_{\rm eff}^2$: (a) -- for different lattice
710: sizes $N$ from a fit of the measured correlation functions
711: $K(0,\tau)$ with the parameterization \Ref{mcorr}; (b) --
712: in the limit $N \rightarrow \infty$ (see text). The solid line
713: on (b) is the fit curve given by expression \Ref{mbeta}.
714: }
715: \end{figure}
716:
717: The obtained masses for different lattice sizes are shown in
718: Fig.\ref{FigMasses}a. For each fixed $\beta_g$ we extrapolated
719: the mass to the limit $N\rightarrow\infty$ using a linear
720: $1/N$ parameterization
721: \BE
722: \label{mNfit}
723: m(\beta_g,N) = m_0(\beta_g) + \frac{C(\beta_g)}{N} ~.
724: \EE
725: Our data for $m(\beta_g,N)$ can be fitted with this linear form
726: \Ref{mNfit} for $\beta_g \gtsim 3$. The resulting dependence
727: of $m_0$ on $\beta_g$ is presented in Fig.\ref{FigMasses}b. The
728: mass vanishes in the infinite volume limit near $\beta\approx5$,
729: which is a first guess for the critical coupling $\beta^*_g$.
730:
731: In order to extract more exact information on the critical behavior
732: \BE
733: \label{mbeta}
734: m(\beta_g) \propto \left(\frac1{\beta_g}-\frac1{\beta_g^*}\right)^\nu ~.
735: \EE
736: from our data on small systems we have done a Finite Size Scaling
737: (FSS) \cite{Binder,Barber} analysis of the variables $\absP$, $\chi$
738: and $g_r$, searching for the critical coupling $\beta_g^*$ in the
739: interval $3 \leq \beta_g \leq 9$. We could determine the critical
740: indices $\beta$, $\gamma$ and $\nu$ using an approach employed for
741: $SU(2)$ gauge theory at the finite temperature transition by Engels
742: et al. \cite{Engels}. The general form of the scaling relations for
743: a variable $V$ ($V$ = $\absP$, $\chi$, $g_r$) is
744: \BE
745: \label{FSSgen}
746: V(t,N) = N^{\rho/\nu} F_V\left(t N^{1/\nu}, g_i N^{y_i}\right),
747: \EE
748: where $\rho$ is the corresponding critical index ($\beta$, $\gamma$, 0)
749: for the respective quantities and $t$ is the reduced inverse coupling
750: $\beta_g$:
751: \BE
752: \label{FSSt}
753: t = \frac{\beta_g^* - \beta_g}{\beta_g}.
754: \EE
755: In practice eq. \Ref{FSSgen} is computed near $t=0$. Expanding up to first
756: order in $t$ and taking into account only the largest irrelevant exponent
757: $y_1\equiv-\omega$ one obtains
758: \BE
759: \label{FSSone}
760: V(t,N) = N^{\rho/\nu}
761: \left[c_0+\left(c_1+c_2 N^{-\omega}\right)tN^{1/\nu}+c_3 N^{-\omega}\right].
762: \EE
763:
764: % ---- Fig. F.S.S.
765: \begin{figure}[ht]
766: \centerline{
767: \scalebox{0.47}{\includegraphics{zm-pa.ps}}~
768: \scalebox{0.47}{\includegraphics{zm-pc.ps}}
769: }
770: \caption{
771: \label{FigFSS}
772: Comparison of the theoretical parameterizations given in eqs.
773: (\ref{FSSt}-\ref{FSSone}) for $\absP$ and $\chi$ with the
774: lattice data results as a function of $\beta_g$ for different
775: lattice sizes $N$.
776: }
777: \end{figure}
778:
779: \begin{table}[ht]
780: \begin{center}
781: \begin{tabular}{|c|c|c|c|}\hline
782: & & & \\[-2mm]
783: $V$ & & Transverse QCD $(2+1)$ & Ising $(D=3)$ \\[-1mm]
784: & & & \\\hline
785: & & & \\[-3mm]
786: $<|P|>$ & $\beta$ & 0.21 $\pm$ 0.06 & 0.3267(10) \\
787: & $\nu$ & 0.57 $\pm$ 0.06 & 0.6289(8) \\
788: & $\beta_g^*$ & 6.3 $\pm$ 1.1 & \\[-3mm]
789: & & & \\\hline
790: & & & \\[-2mm]
791: $\chi$ & $\gamma$ & 1.13 $\pm$ 0.36 & 1.239(7) \\
792: & $\nu$ & 0.52 $\pm$ 0.10 & 0.6289(8) \\
793: & $\beta_g^*$ & 5.7 $\pm$ 1.2 & \\[-3mm]
794: & & & \\\hline
795: \end{tabular}
796: \end{center}
797: \caption{
798: \label{TabRes}
799: Critical parameters for the transverse QCD near the light cone.
800: The Ising indices are from ref.\cite{Ising}.
801: }
802: \end{table}
803:
804: With this parameterization we analyze our data for lattice sizes
805: $N=4,6,8,12,16$. The $\beta_g$ dependence in eq. \Ref{FSSone} is
806: valid near the critical point. The quality of the fit to the data
807: in different intervals on $\beta_g$ can serve as a guide to localize
808: the critical point $\beta_g^*$ \cite{Engels}. We find that in our case
809: this parameterization can be applied for $\absP$ and $\chi$ only for
810: $\beta_g~\gtsim~5$: for smaller $\beta_g$ the $\chi^2/D.F.$ becomes
811: too large. Finally we use the parameterization of eq. \Ref{FSSone}
812: in the interval $5 \leq \beta_g \leq 9$ (see Fig.\ref{FigFSS}) where
813: $\vert t\vert~\ltsim~0.3$ and an expansion of first order in $t$
814: still is applicable. In this $\beta_g$ interval the fits for $\absP$
815: and $\chi$ yield a $\chi^2/D.F\sim2$. To have a reliable error
816: estimate for the extracted parameters we use the jackknife method.
817: The results for $\beta_c^*$ and critical indices $\beta$, $\gamma$
818: and $\nu$ are presented in Table~1. We use $\omega=1$, but in fact
819: the results do not depend on $\omega$ in the interval
820: $0.8\leq\omega\leq1.2$.
821:
822: It should be noticed that the system frequently flips between the two
823: ordered states on a finite lattice. Therefore the expectation value
824: $\langle P \rangle$ in equation \Ref{chidef} vanishes with good accuracy.
825: Analogously to the treatment of the magnetization in the 3-dimensional
826: Ising model, we should - instead of the expression for $\chi$ in eq.
827: \Ref{chidef} - define the susceptibility in the broken phase as
828: $\chi_{\rm broken}=N^3 \left(\langle P^2 \rangle - \absP^2\right)$.
829: For the Ising case this susceptibility is supposed to converge towards
830: the correct infinite volume limit. Due to lack of statistics, however,
831: we did not use separate expressions for $\chi$ in the two phases.
832: Our data for $g_r$ are even much less accurate which excludes the
833: possibility to use them in the FSS analysis.
834:
835: The values found for the critical indices $\beta$, $\gamma$ and $\nu$
836: are not far from those of the Ising model ($D=3$) (see Table~1). They
837: are much less accurate but also in agreement with a high statistics
838: analysis \cite{Engels} of the $SU(2)$ deconfinement transition
839: at finite temperature. We interprete the behavior of the light
840: cone zero mode theory as a consequence of the underlying $Z(2)$
841: symmetry of the Hamiltonian. $Z(2)$ transformations correspond to
842: reflections of the angle variables $\varphi$ around $\frac{\pi}{2}$.
843: The Hamiltonian and the measure of eq. (17) are invariant under these
844: transformations. The resulting critical behavior is common to
845: Ising like models. The critical coupling itself is a nonuniversal
846: quantity and we have found a rough value. More extended work with
847: the lattice Hamiltonian near the light cone is needed to clarify
848: the continuous transition further.
849:
850: % ------- Conclusions --------------------------------------
851: \section {Conclusions}
852:
853: The scaling analysis gives indications that there is a second order
854: transition between a phase with massive excitations at strong coupling
855: and a phase with massless excitations in weak coupling. To reach a
856: higher accuracy large scale simulations of the Hamiltonian zero mode
857: system are needed. In this context it is advisable to treat the coupled
858: system of three and two dimensional modes together. A calculation in the
859: epsilon expansion \cite{LZ} gives the zero of the $\beta$-function
860: as an infrared stable fixed point. This shows that the limit of large
861: longitudinal dimensions $L$ is well defined. Using the running coupling
862: constant $g^2_{\rm eff}$ of the zero mode system we have (\cf \Ref{mbeta})
863: \BE
864: m\,a = \frac{1}{\zeta_0 g^{*\,2}_{\rm eff}}
865: \left(g^2_{\rm eff} - g^{*\,2}_{\rm eff}\right)^\nu
866: \EE
867: with $g^{*\,2}_{\rm eff} = 0.17\pm0.03$ and $\nu = 0.56\pm0.05$ from
868: our lattice calculations (\cf Table~\ref{TabRes}). The mass can be
869: interpreted as an inverse correlation length $ma = a/\xi$. When the
870: correlation length $\xi$ approaches the temporal size of the lattice
871: $\xi \approx N_T\, a/2$ in the infrared limit we can identify $ma$ with
872: $a/L$, where $L$ is our infrared length scale. Therefore the effective
873: coupling $g^2_{\rm eff} = g^2 \eta \frac{L}{4a}$ (\cf \Ref{g2eff}) runs
874: with $a/L$ in the following way
875: \BE
876: \label{gevol}
877: g^2_{\rm eff} =
878: g^{*\,2}_{\rm eff}
879: + \left(\zeta_0 g^{*\,2}_{\rm eff}\right)^{1/\nu}
880: \left(\frac{a}{L}\right)^{1/\nu} ~.
881: \EE
882: The coupling to the three dimensional modes produces the usual
883: evolution of $g^2$ in $SU(2)$ QCD, where the coupling $g^2(a)$
884: at the lattice scale $a$ is related to the coupling $g^2_0$
885: defined at the infrared scale $\approx L$ as follows
886: \BE
887: \label{gevola}
888: g^2(a)= \frac{g^2_0}{1
889: + \frac{g^2_0}{4\pi^2}\,\frac{22}{3}\,\log\frac{L}{a}}~.
890: \EE
891: Combining the eqs. \Ref{gevol} and \Ref{gevola} we can synchronize
892: the approach to the light cone, {\it i.e.} the limit $\eta\rightarrow0$
893: with the continuum limit $a\rightarrow0$. The condition that the
894: three-dimensional evolution of $g^2$ has to be compatible with the
895: two-dimensional evolution of $g^2_{\rm eff}$ towards $g^{*\,2}_{\rm eff}$
896: yields that for $a\to0$ the light cone parameter $\eta$ approaches zero as
897: \BE
898: \eta(a) \sim
899: \frac{g^{*\,2}_{\rm eff}}{\pi^2} \,
900: \frac{22}{3} \, \frac{a}{L} \, \log\frac{L}{a} ~.
901: \EE
902:
903: This relation is similar to the naive scaling result deduced in section
904: $1$ from physics arguments besides logarithmic modification. Although at
905: the outset we had as parameters $N_L=L/a$ and $\eta$ in the Hamiltonian
906: we reduce this multiparameter problem to a problem with one single coupling
907: constant which can be chosen as $g$.
908:
909: We can use the energy and the critical behavior of the Euclidean $2+1$
910: dimensional system to obtain its contribution to the ground state energy
911: of the system in $3+1$ dimensions. In our calculations the energy of the
912: zero mode Hamiltonian $\langle h \rangle$ near the critical coupling is
913: proportional to $N^2_\bot/\eta$. Using the scaling of $\eta$ determined
914: from the compatibility of 2-dimensional and 3-dimensional dynamics, we
915: can eliminate $\eta$ in favor of $N_L = L/a$ and get a zero mode energy
916: which grows linearly with the 3-dimensional volume. That means the zero
917: mode dynamics becomes relevant for the full problem, {\it i.e.}
918: \BE
919: \langle h a \rangle \sim N_L N^2_\bot ~.
920: \label{meanha}
921: \EE
922: In reality the zero mode system is coupled to the $3+1$ dimensional system
923: the extension of which is large. So we expect that the couplings of the
924: $2+1$ dimensional system can get modified. Because of the universal dynamics
925: near the infrared fixed point this will not change the qualitative form of
926: the volume dependence obtained in eq. \Ref{meanha}. The behavior of the
927: zero mode Hamiltonian has to be taken into account when we try to solve
928: for the infinite momentum frame solutions of the complete Hamiltonian.
929: The discussion \cite {VFP, HMV} until now often centers on the order of
930: limits. In order to guarantee a meaningful limit of the near light cone
931: theory, one first has to take $L \rightarrow \infty$ and then one can
932: take $\eta \rightarrow 0$. This is a workable procedure in solvable $1+1$
933: dimensional models. In QCD in $3+1$ dimensions it hardly can serve as a
934: prescription. Numerical solutions have to be obtained with a finite cutoff
935: and the continuum limit can only be reached approximately via scaling
936: relations. It is in this case that the role of zero mass excitations will
937: crucially enter the physics. As one sees above the synchronization of the
938: continuum limit with the light cone limit leads to results which are
939: independent of the near light cone parameter $\eta$. The independence of
940: the ground state energy near the critical coupling gives us back Lorentz
941: invariance which states that the physics should not depend on the reference
942: frame. The vacuum energy is not allowed to depend on $\eta$. There are
943: dynamical zero modes in near light cone QCD which contain the physics of
944: the nonperturbative QCD vacuum, the physics of the gluon and quark
945: condensates. The nontrivial structure of the nonabelian gauge theory
946: enters in our calculation through the kinetic operator of the rotators
947: which at finite resolution have a finite mass gap vanishing only in the
948: continuum limit.
949:
950: The near light cone description of QCD based on a modified axial gauge
951: $\partial_-A_-=0$ yields a very natural formulation of high energy scattering.
952: Polyakov variables near the light cone have already entered the calculations
953: of high energy cross sections in two different ways: For the geometrical size
954: of cross sections in ref.\cite{DGKP,N} the correlations between Polyakov
955: lines along the projectile and target directions are important. For this
956: situation an interpolating gauge may be useful. For the energy dependence
957: of cross sections an evolution equation of Polyakov line correlators
958: approaching the light cone has been discussed in ref. \cite{Bal} which
959: generalizes the BFKL- equation. This situation is close to the treatment
960: in this paper. Now the decisive step is to back up our theoretical work
961: on effective transverse QCD dynamics near the light cone with
962: phenomenological consequences for high energy scattering.
963:
964: % ------- Appendix -----------------------------------------
965: \section*{Appendix}
966:
967: A choice of $\Dtau$ is proposed by comparing Monte-Carlo weights
968: $W_t$ and $W_s$ which link nearest neighbors in time and space
969: respectively
970: \BA
971: W_t(\varphi_1,\varphi_2) &=&
972: \langle \varphi_1(x,y,\tau+1) \vert e^{-\Dtau h^{\rm e}} \vert
973: \varphi_2(x,y,\tau ) \rangle \,
974: \sqrt{J(\varphi_1) J(\varphi_2)} ~, \\
975: W_s(\varphi_1,\varphi_2) &=&
976: \langle \varphi_1(x+1,y,\tau) \vert e^{-\Dtau h^{\rm m}} \vert
977: \varphi_2(x ,y,\tau) \rangle ~.
978: \EA
979: \begin{figure}[ht]
980: \vskip -15mm
981: \centerline{
982: \scalebox{0.3}{\includegraphics{zm-wa.ps}}~~
983: \scalebox{0.3}{\includegraphics{zm-wb.ps}}~~
984: \scalebox{0.3}{\includegraphics{zm-wc.ps}}
985: }
986: \vskip -7mm
987: \caption{
988: \label{FigWlinks}
989: Weights $W_t(\varphi_1,\varphi_2)$ of time-like and
990: $W_t(\varphi_1,\varphi_2)$ of space-like links
991: for different $\Dtau$.
992: Figures (a,b,c) correspond to $\Dtau = a/10, a/2, a$.
993: }
994: \end{figure}
995:
996: The probability distribution for the update of the field
997: $\varphi(x,y,\tau)$ is given by the product of weights for all
998: links connecting it to neighbors in time and space. So the
999: updating procedure for $\varphi$ is more effective when $W_t$
1000: and $W_s$ have similar distributions near the angle $\varphi=%
1001: \pi/2$ where the Jacobian has its maximum. For $\Dtau=a/10$
1002: (\cf Fig.~\ref{FigWlinks}a) the spatial weights $W_s$ have
1003: a much wider distribution than the timelike weights $W_t$,
1004: whereas for $\Dtau=a$ (\cf Fig.~\ref{FigWlinks}c) the
1005: situation is reverse. For the intermediate case $\Dtau=a/2$
1006: (\cf Fig.~\ref{FigWlinks}b) the widths of the distributions
1007: are very similar, therefore the updating procedure is optimal.
1008:
1009: % ------- Bibliography -------------------------------------
1010: \begin{thebibliography}{99}
1011: \bibitem{Brodsky} S.~Brodsky, H.-C.~Pauli, S.~Pinsky,
1012: Phys.~Rept.~{\bf 301} 299 (1998);
1013: hep-ph/9705477.
1014: \bibitem{NPFV} R.~Naus, H.~J.~Pirner, T.~J.~Fields and J.~P.~Vary,
1015: Phys.~Rev.~{\bf D 56} 8062 (1997).
1016: \bibitem{PRFR} E.~V.~Prokhvatilov and V.~A.~Franke,
1017: Sov.~J.~Nucl.~Phys.~{\bf 49} 688 (1989).
1018: \bibitem{LTLY} F.~Lenz, M.~Thies, S.~Levit and K.~Yazaki,
1019: Ann.~Phys.~(N.Y.)~{\bf 208} 1 (1991).
1020: \bibitem{VFP} J.~P.~Vary, T.~J.~Fields and H.~J.~Pirner,
1021: Phys.~Rev.~{\bf D 53} 7231 (1996).
1022: \bibitem{DGKP} H.~G.~Dosch, T.~Gousset and H.~J.~Pirner,
1023: Phys.~Rev.~{\bf D 57} 1666 (1998); \\
1024: U.~D'Alesio, A.~Metz and H.~J.~Pirner,
1025: Eur.~Phys.~J.~{\bf C 9} 601 (1999).
1026: \bibitem{N} O.~Nachtmann,
1027: Ann.~Phys.~(N.Y.)~{\bf 209} 436 (1991); \\
1028: O.~Nachtmann, Schladming Lectures 1996,
1029: {\it Perturbative and nonperturbative aspects of
1030: quantum field theory} pp.~49-138.
1031: \bibitem{Pol} S.~Hellerman and J.~Polchinski,
1032: Phys.~Rev.~{\bf D 59} 125002 (1999); \\
1033: J.~Polchinski,
1034: Prog.~Theor.~Phys.~Suppl. {\bf 134} 158 (1999);
1035: hep-th/9903165.
1036: \bibitem{LNOT} F.~Lenz, H.~W.~L.~Naus, K.~Ohta and M.~Thies,
1037: Ann.~Phys.~(N.Y.)~{\bf 233} 17 (1994).
1038: \bibitem{LNT} F.~Lenz, H.~W.~L.~Naus and M.~Thies,
1039: Ann.~Phys.~(N.Y.)~{\bf 233} 317 (1994).
1040: \bibitem{LZ} J.~C.~Le~Guillou and J.~Zinn~Justin,
1041: Phys.~Rev.~Lett.~{\bf 39} 95 (1977).
1042: \bibitem{Binder} K.~Binder,
1043: Z.~Phys.~{\bf B 43} 119 (1981).
1044: \bibitem{Barber} M.~N.~Barber,
1045: {\it Phase Transitions and Critical Phenomena},
1046: vol.~{\bf 8}, p.~146 (1983).
1047: \bibitem{Engels} J.~Engels, J.~Fingberg and V.~K.~Mitrjushkin,
1048: Phys.~Lett.~{\bf B 298} 154 (1993); \\
1049: J.~Engels, S.~Mashkevich, T.~Scheideler and G.~Zinovev,
1050: Phys.~Lett.~{\bf B 365} 219 1996.
1051: \bibitem {Ising} A.~M.~Ferrenberg and D.~P.~Landau,
1052: Phys.~Rev.~{\bf B 44} 5081 (1991),
1053: D.~P.~Landau,
1054: Physica {\bf A~205} 41 (1994).
1055: \bibitem {HMV} A.~Harindranath, L.~Martinovic and J.~P.~Vary,
1056: hep-th/9912085.
1057: \bibitem {Bal} I.~I.~Balitsky, hep-ph/9706411.
1058: \end{thebibliography}
1059: \end{document}
1060: